1: \documentclass[10pt]{article}
2: \usepackage{graphicx}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \newcommand{\ds }{ \displaystyle }
5: \newcommand{\bd }[1]{ {\mbox{\boldmath $#1$}} }
6: \newcommand{\spc}{ \,\,\,\, }
7: \newcommand{\R}{ {I\!\!R} }
8: \newcommand{\Z}{ {Z} }
9: \newcommand{\N}{ {I\!\!N} }
10: \newcommand{\fns} { \footnotesize }
11: \newcommand{\scs} { \scriptsize }
12: \newcommand{\U}[3] { u^{#1}_{(#2,#3)} }
13: \newcommand{\mP} {\mathcal{P}}
14: \newcommand{\mA} {\mathcal{A}}
15: \newcommand{\mB} {\mathcal{B}}
16: \newcommand{\ta} {\gamma}
17: \newcommand{\infnm}[1] { \| #1 \|_{\infty} }
18: \newcommand{\pdeT}[1]{\frac{\partial #1}{\partial t}}
19: \newcommand{\fTd}{ D_{\ast}^{\beta} }
20: \newcommand{\fSd}{ D_{0}^{\alpha} }
21: \newcommand{\uxt}{ u(\bd{x},t) }
22: \newtheorem{theorem}{Theorem}
23: \newenvironment{remark}
24: {\noindent \slshape Remark: \begin{quote} \small \itshape}{\end{quote}}
25:
26: \DeclareGraphicsExtensions{.jpg, .eps}
27: \DeclareGraphicsRule{.jpg}{eps}{.jpg.bb}{`jpeg2ps -h -r 600 #1}
28: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
29: \begin{document}
30:
31:
32:
33: \begin{center}
34: Monte Carlo Random Walk Simulations Based on Distributed Order
35: Differential Equations
36: \vspace*{0.5cm}
37:
38: {\small
39: Erik Andries \\
40: Department of Pathology \& Department of Mathematics and Statistics\\
41: The University of New Mexico, Albuquerque, New Mexico 87131\\
42: andriese@unm.edu\\~\\
43: Sabir Umarov \\
44: Department of Mathematics and Mechanics\\
45: The National University of Uzbekistan, Tashkent, Uzbekistan\\
46: sabir@math.unm.edu\\~\\
47: Stanly Steinberg
48: \footnote{Partially supported by NIH grant P20 GMO67594.}\\
49: Department of Mathematics and Statistics\\
50: The University of New Mexico, Albuquerque, New Mexico 87131\\
51: stanly@math.unm.edu
52: }
53: \end{center}
54: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
55: \begin{abstract}
56: In this paper the multi-dimensional random walk models governed by
57: distributed fractional order differential equations and multi-term fractional order
58: differential equations are constructed. The scaling limits of
59: these random walks to a diffusion process in the sense of distributions is proved.
60: Simulations based upon multi-term fractional order
61: differential equations are performed.
62: \end{abstract}
63:
64: \vskip 6pt
65:
66: {\it Mathematics Subject Classification}: 65C05, 60G50, 39A10, 92C37
67:
68: {\it Key Words and Phrases}: Random walk,
69: distributed order differential equation, Monte-Carlo simulation, Markovian jumps, non-Markovian jumps
70:
71: \vskip 3pt
72:
73: \section{Introduction}
74:
75: \subsection{Motivation.}
76: In this paper we study simulation models based on distributed
77: order differential equations, which we will call DODE simulations.
78: This type of simulation reflects the rich structure of diffusion media, in
79: which a several diffusion modes are possible. Diffusion processes with
80: complex and changing modes are ubiquitous in nature (see,
81: \cite{BouchaudGeorges90,ChechkinGonchar,MK,UchaykinZolotarev,Zaslavsky} and references
82: therein). One of the motivations for conducting
83: DODE simulations is to model the
84: movement of proteins on the cell membrane. Numerous experiments
85: \cite{Edidin,GhoshWebb,Kusumi,Saxton01,SaxtonJacobson97} show that
86: macromolecule movement through the cell membrane is distinct from Brownian
87: motion. Saxton and Jacobson \cite{SaxtonJacobson97} noted that practically
88: all experimental results show apparent transitions among modes of motion.
89:
90: The governing
91: equation, which we take as a basis for our simulation models,
92: in general form, is
93: distributed space fractional order differential equation
94: \begin{equation}
95: \label{dode0}
96: D_{\ast}^{\beta} u(t,x) = \int_0^2 a(\alpha)
97: D_0^{\alpha} u(t,x) d \alpha, \, t>0, \, x \in {\R}^N,
98: \end{equation}
99: where $0<\beta \leq 1$, $D_{\ast}^{\beta}$ is the Caputo fractional order
100: derivative \cite{Caputo,GLU}, $D_0^{\alpha}=(-\Delta)^{{\alpha \over 2}}$ is the space fractional
101: order (pseudo-differential) operator with the symbol $|\xi|^{\alpha}$.
102: Note that $D_0^{\alpha}$ can be written in the form of hypersingular
103: integral as well \cite{SKM}. The function $a(\alpha)$ is a positive
104: integrable function (or positively defined distribution). Depending on
105: $a(\alpha)$, (\ref{dode0}) may become a multi-term fractional order
106: differential equation, which can possibly describe the existence of
107: a finite number of diffusion regimes.
108: Although, the distributed order differential operators were first mentioned by
109: \cite{Caputo1,Caputo2} in the 1960s, the intensive study of models
110: based on the distributed order differential
111: equations has been started recently
112: \cite{BagleyTorvic,Chechkin,Diethelm,LorenzoHartley,MMM,UG05-2,UmarovSteinberg}.
113:
114: The present report is organized as follows. In Section 2, we briefly recall the
115: theoretic platform of the construction of the DODE simulation models announced
116: in \cite{UmarovSteinberg}. In Section 3 we analyze
117: the difference schemes associated with the
118: DODE models, and in Sections 4 and 5 we
119: construct random walk models and simulations based on the transition
120: probabilities introduced in the previous sections.
121:
122: \subsection{Notation.}
123: In this paper, ${\R^N}$ is the $N$-dimensional Euclidean space with
124: coordinates ${x} =(x_1,...,x_N)$ while ${\Z^N}$ is the $N$-dimensional
125: integer-valued lattice with the lattice nodes being given by the multi-index
126: notation $j=(j_1,...,j_N)$. The letters $i$, $j$ and $k$ will be exclusively
127: used for the multi-indexing of lattice nodes. We denote by $x_j =
128: (h_{j_1},...,h_{j_N}), j \in {\Z}^N$, the nodes of the uniform $h$-lattice
129: ${\Z}_h^N$ which is defined as $(h {\Z})^N$ with $h$ being the distance
130: between any two lattice nodes. We introduce a spatial grid $\{x_j=jh, j \in
131: {\Z}^N \}$, with $ h>0$ and a temporal grid $\{t_n=n \tau, n=0,1,2,... \}$
132: with a fixed stepsize $\tau>0$. Furthermore, let $u^n_j$ denote the
133: discretization of the function
134: $u(t,x)$ on the spatial and temporal grid at $x=x_j$ and $t=t_n$, i.e
135: $u^n_j = u(t_n,x_j)$.
136:
137:
138: \section{Markovian random walks associated with the DODE}
139: \subsection{Particle jumps.}
140: Assume ${\bf X}$ to be a N-dimensional random vector \cite{MS} whose values
141: range in ${\Z}^N$. Let a sequence of random vectors ${\bf X}_1, {\bf X}_2,
142: ...$ also be N-dimensional independent identically distributed random
143: vectors, all having the same probability distribution. Consider the sequence
144: of random vectors
145: %-------------------------------------------------------------------
146: \[
147: {\bf S}_n = h{\bf X}_1+h{\bf X}_2+...+h{\bf X}_n, n=1,2,...
148: \]
149: %-------------------------------------------------------------------
150: taking ${\bf S}_0 ={\bf 0} =(0,\ldots,0) \in {\Z}^N_h$ for convenience.
151: We interpret ${\bf X}_1, {\bf X}_2, ...$,
152: as a sequence of particle jumps
153: starting time $t=t_0=0$.
154: At time $t=t_n$, the particle takes a jump $h{\bf X}_n$ from
155: ${\bf S}_{n-1}$ to ${\bf S}_n$.
156: If $u^n_j=u(t_n,x_j)$ is the probability of
157: a particle being at location $x_j$ at time $t_n$ and,
158: taking into account the
159: recursion ${\bf S}_{n+1}={\bf S}_n + h{\bf X}_{n+1}$, we have
160: %-------------------------------------------------------------------
161: \begin{equation}
162: \label{recursion}
163: u^{n+1}_j = \sum_{k \in {\Z}^N}p_k u^n_{j-k}, j \in {\Z}^N, \,\,
164: n=0,1,...
165: \end{equation}
166: %-------------------------------------------------------------------
167: where the coefficients $p_k, \, k \in {\Z}^N$ are called the transition
168: probabilities.
169: The convergence of the sequence ${\bf S}_n$ when
170: $n \rightarrow \infty$
171: means convergence of the discrete probability law (probability mass
172: function)
173: $(u^n_j)_{j\in {\Z}^N}$, properly rescaled as explained
174: below, to the probability law with a density $u(t,x)$ in the
175: sense of distributions (in law). This is equivalent to the locally
176: uniform convergence of the corresponding characteristic functions
177: (see for details \cite{MS}). This idea is used in \cite{UG05-1,UmarovSteinberg}
178: to prove the convergence of the
179: sequence of characteristic functions of the
180: corresponding random walks to the
181: fundamental solution of distributed order diffusion equations.
182:
183: \subsection{Markovian transition probabilities.}
184: Let the transition probabilities in Eq.(\ref{recursion}) take the form
185: %-------------------------------------------------------------------
186: \begin{equation}
187: p_{k} = \tau q_k(\alpha,h), \,\, k \neq 0,
188: \label{eq:p_k}
189: \end{equation}
190: %-------------------------------------------------------------------
191: where
192: %-------------------------------------------------------------------
193: \begin{equation}
194: q_k(\alpha,h) = \int_0^2
195: \left[
196: \frac{ a(\alpha) b(\alpha) }
197: { \vert k \vert^{N+\alpha} h^{\alpha} }
198: \right] d \alpha,
199: \spc \mbox{and} \spc
200: b(\alpha) =
201: \frac{
202: \left[
203: \Gamma \left( 1 + \frac{\alpha}{2} \right)
204: \right]^2
205: sin \left( \frac{\alpha}{2} \pi \right)
206: }{
207: \pi^2 2^{N-\alpha-1}
208: }.
209: \label{eq:q_k}
210: \end{equation}
211: %-------------------------------------------------------------------
212: The transition probability $p_0$ can then be defined as
213: %-------------------------------------------------------------------
214: \begin{equation}
215: p_0 = 1 - \sum_{k \neq 0} p_k = 1 - \tau q_0(\alpha,h),
216: \label{eq:p_0}
217: \end{equation}
218: %-------------------------------------------------------------------
219: where
220: %-------------------------------------------------------------------
221: \begin{equation}
222: q_0(\alpha,h) =
223: \sum_{k \neq 0} q_k(\alpha,h) =
224: \sum_{k \neq 0} \int_0^2 \left[
225: \frac{ a(\alpha) b(\alpha) }{ \vert k \vert^{N+\alpha} h^{\alpha} }
226: \right] d \alpha,
227: \label{eq:q_0}
228: \end{equation}
229: %-------------------------------------------------------------------
230: Assuming that the condition
231: $0 < \tau q_0(\alpha,h) \leq 1$ is fulfilled,
232: the transition probabilities then satisfy the following properties:
233: %-------------------------------------------------------------------
234: \begin{enumerate}
235: \item ${\displaystyle \sum_{k \in {\Z}^N} p_k} = 1;$
236: \item $p_k \geq 0,{k \in {\Z}^N}.$
237: \end{enumerate}
238: %-------------------------------------------------------------------
239: Note that the non-negativity condition\footnote{This condition
240: is equivalent to the
241: stability condition of finite-difference schemes giving the
242: usual stability condition if $a(\alpha)=\delta(\alpha-2)$.}
243: in property 2 is linked with the Riemann zeta-function.
244: Indeed, introduce the function
245: %-------------------------------------------------------------------
246: \begin{equation}
247: {\mathcal{R}} (\alpha) =
248: \sum_{k \neq 0} \frac{1}{|k|^{N+\alpha}} =
249: \sum_{m=1}^{\infty} \frac{M_m}{m^{N+\alpha}}, \spc 0<\alpha \leq 2,
250: \end{equation}
251: %-------------------------------------------------------------------
252: where $M_m = \sum_{|k|=m}1.$ In the one-dimensional case
253: ${\mathcal{R}}(\alpha)= 2 \zeta (1+\alpha)$, where $\zeta (z)$ is the
254: Riemann zeta-function.
255: Then the nonnegativity condition $0 < p_0 \leq 1$ can be rewritten as
256: %-------------------------------------------------------------------
257: \begin{equation}
258: \label{eq:condfortrpr1}
259: \tau q_0(\alpha, h) = \tau \int_0^2 \left[
260: \frac{a(\alpha)b(\alpha){\mathcal{R}}(\alpha)}{h^{\alpha}}
261: \right] d \alpha \leq 1.
262: \end{equation}
263: %-------------------------------------------------------------------
264: It follows from this condition that $h \rightarrow 0$ yields
265: $\tau \rightarrow 0$. This, in turn,
266: yields $t/\tau \rightarrow \infty$ for any finite $t.$
267:
268: %%%%%%%%%THEOREM-1%%%%%%%%%%%
269: \begin{theorem}
270: \label{theorem:dode}
271: Let ${\bf X}$ be a random vector with the transition probabilities
272: $p_{k} = P({\bf X}=x_k), k \in {\Z}^N,$ defined in
273: Eq.(\ref{eq:p_k}) and Eq.(\ref{eq:p_0})
274: which satisfy properties 1 and 2.
275: Then the sequence of random vectors
276: ${\bf S}_{n}=h{\bf X}_{1}+ ... + h{\bf X}_{n},$
277: converges as $n \rightarrow \infty$ in law to
278: the random vector whose probability density function is the
279: fundamental solution of the distributed space fractional order
280: differential equation (\ref{dode0}) with $\beta=1$.
281: %(\ref{cc}), i.e. $G(t,x)$ defined in (\ref{fs}).
282: \end{theorem}
283: %%%%%%%%%%%%%%%%%
284:
285: Note, for the simulations used in this paper, it is important to use
286: the multi-term analog of this theorem. Assuming that
287: %-------------------------------------------------------------------
288: \begin{equation}
289: \label{discret}
290: a(\alpha)=\sum_{m=1}^{M} a_m \delta (\alpha - \alpha_m),
291: \spc
292: 0 < \alpha_1 < \cdots < \alpha_M \leq 2,
293: \end{equation}
294: %-------------------------------------------------------------------
295: with positive constants $a_m$, we get a multiterm DODE
296: %-------------------------------------------------------------------
297: \begin{equation}
298: \label{multiterm}
299: D_{\ast}^{\beta} u(t,x) = \sum_{m=1}^M a_m
300: D_0^{\alpha_m}u(t,x), \spc
301: t>0, \, x \in {\R}^N.
302: \end{equation}
303: %-------------------------------------------------------------------
304: Also note that the coefficients $q_k(\alpha,h)$ in Eq.(\ref{eq:q_k})
305: and Eq.(\ref{eq:q_0}) become multi-term as well:
306: %-------------------------------------------------------------------
307: \[
308: q_k(\alpha,h) =
309: {\ds \sum_{m=1}^M
310: \left[
311: \frac{ a_m b(\alpha_m) }
312: { \vert k \vert^{N+\alpha_m} h^{\alpha_m} }
313: \right]}, \,\, k \neq 0,
314: \quad
315: q_0 = {\ds \sum_{k \neq 0} q_k}.
316: \]
317: %-------------------------------------------------------------------
318:
319:
320: %%%%%%%%%%%%%%%( THEOREM 2
321: \begin{theorem}
322: \label{theorem:multiterm}
323: Let the transition probabilities
324: $p_{k} = P({\bf X}=x_k), k \in {\Z}^N,$ of the
325: random vector ${\bf X}$ be given as follows:
326: %----------------------------------------------------------
327: \begin{equation}
328: p_k = \tau q_k(\alpha,h)
329: \spc \mbox{and} \spc
330: p_0 = 1 - \tau q_0(\alpha,h)
331: \label{eq:markovian_transprob}
332: \end{equation}
333: %----------------------------------------------------------
334: where $a(\alpha) = \sum_{m=1}^M a_m \delta(\alpha-\alpha_m)$.
335: Assume
336: %----------------------------------------------------------
337: \[
338: \tau
339: \sum_{m=1}^M
340: \frac{ a_m b(\alpha_m){\mathcal{R}}(\alpha_m) }
341: { h^{\alpha_m} }
342: \leq 1.
343: \]
344: %----------------------------------------------------------
345: Then the sequence of random vectors
346: ${\bf S}_{n}=h{\bf X}_{1}+ ... + h{\bf X}_{n},$
347: converges as $n \rightarrow \infty$ in law to
348: the random vector whose probability density function is the
349: fundamental solution of the multiterm fractional order differential
350: equation (\ref{multiterm}) with $\beta=1$.
351: \end{theorem}
352: %%%%%%%%%%%%%%%( END OF THEOREM 2
353:
354: \begin{remark}
355: As we noted above these results were announced in \cite{UmarovSteinberg}. The more general case of
356: these theorems corresponding to a fractional $\beta \in (0,1)$ can be
357: obtained introducing a positive waiting time distribution and corresponding
358: iid random variables \cite{GM,MMM}. We do not describe this case in this
359: paper. We note only that the general case is studied by applying
360: a general finite-difference approach and that this
361: general difference scheme is stable under some
362: condition and has a unique solution.
363: \end{remark}
364: \vspace{0.5cm}
365:
366:
367: \section{Generalized Transition Probabilities for the DODE}
368:
369: The set of grid points in $\Z^N_h$
370: used to update $u$ at time $t=t_{n+1}=(n+1) \tau$
371: is called the stencil.
372: In this section, we start from stating the values of the transition
373: probabilities associated with the stencil for the
374: discretization of the particular space-time-fractional
375: differential equation,
376: %----------------------------------------------------------------
377: \begin{equation}
378: D_{\ast}^{\beta} u(t,x)
379: =
380: D_{0}^{\alpha} u(t,x), \spc
381: t>0, \,
382: x \in \R^N, \,
383: 0 < \beta \leq 1, \, 0 < \alpha \leq 2,
384: \label{eq:frac_diff_eqn}
385: \end{equation}
386: %----------------------------------------------------------------
387: and then generalize it to distributed order differential equations.
388: %----------------------------------------------------------------
389: \vspace{0.3cm}
390:
391:
392: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
393: \subsection{Discretization of the time-fractional derivative.}
394: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
395: Using the Caputo time-fractional derivative \cite{Caputo},
396: the left-hand-side of (\ref{eq:frac_diff_eqn}) becomes
397: %----------------------------------------------------------------
398: \begin{equation}
399: D_{\ast}^{\beta}u(t,x) =
400: \frac{1}{\Gamma(1-\beta)} \int_{0}^t
401: \left[ \frac{\partial u(s,x)}{\partial s} \right]
402: \frac{ds}{(t-s)^{\beta}}, \,\,
403: 0 < \beta < 1.
404: \label{eq:time_frac_der}
405: \end{equation}
406: %----------------------------------------------------------------
407: Note that when $\beta=1$,
408: $D_{\ast}^{\beta}{u(t,x)} = \partial u/ \partial t$.
409: When $0 < \beta < 1$, we will use the following
410: discretization (see \cite{LSAT_05} for the derivation):
411: %----------------------------------------------------------------
412: \begin{eqnarray}
413: D_{\ast}^{\beta} u^n_j
414: & \approx &
415: \frac{1}{\Gamma(1-\beta)}
416: \sum_{m=0}^n \int_{t_n}^{t_{n+1}}
417: \frac{ u^{'}_j (t_{n+1}-s) }{ s^{\beta} } ds
418: \nonumber \\
419: & = & \frac{1}{\nu \tau^{\beta}}
420: \left(
421: u^{n+1}_j - \sum_{m=1}^n c_m u^{n+1-m}_j - \gamma_n u^0_j
422: \right)
423: \label{eq:time_frac_der_discr_caputo}
424: \end{eqnarray}
425: %----------------------------------------------------------------
426: where
427: %----------------------------------------------------------------
428: \[
429: \gamma_m = (m+1)^{1-\beta} - m^{1-\beta}, \,\, m=0,1,\ldots,n,
430: \quad
431: c_m = \gamma_{m-1} - \gamma_m, \,\, m=1,\ldots,n
432: \]
433: %----------------------------------------------------------------
434: and $\nu = \Gamma(2-\beta)$.
435: The formulas for the coefficients
436: $c_m$ and $\gamma_m$ and the scalar $\nu$
437: that were used in (\ref{eq:time_frac_der_discr_caputo}),
438: which were based upon the Caputo time-fractional derivative, easily
439: generalize to other definitions of the
440: time-fractional derivative. For
441: example, in the case of the
442: Grunwald-Letnikov time-fractional derivative,
443: $\nu = 1$ and $\gamma_m$ and $c_m$ are
444: re-defined as the following \cite{Cies}:
445: %----------------------------------------------------------------
446: \[
447: c_m = {\ds \left|
448: \left( \begin{array}{c} \beta \\ m \end{array} \right)
449: \right|,
450: } \, k=1,\ldots,n,
451: \quad
452: \gamma_{m} = {\ds 1-\sum_{i=1}^m c_i}, \, m=0,\ldots,n.
453: \]
454: %----------------------------------------------------------------
455: For simplicity of notation, we will now set
456: %----------------------------------------------------------------
457: \[ \begin{array}{rcl}
458: w_0 & = & \gamma_n \\
459: w_i & = & c_{n+i-1}, \,\, i=1,\ldots,n.
460: \end{array} \]
461: %----------------------------------------------------------------
462: and, as a result, (\ref{eq:time_frac_der_discr_caputo}) can be
463: rewritten as
464: %----------------------------------------------------------------
465: \begin{equation}
466: D_{\ast}^{\beta} u^n_j =
467: \frac{1}{\nu \tau^{\beta}}
468: \left( u^{n+1}_j - \sum_{m=0}^n w_m u^m_j \right).
469: \label{eq:space_time_frac_eqn_w}
470: \end{equation}
471: %----------------------------------------------------------------
472: Note that for $\beta=1$, $\nu=\Gamma(2-\beta)=1$ and
473: $w_0=\cdots=w_{n-1}=0$ with $w_n=1$. In this case,
474: (\ref{eq:time_frac_der_discr_caputo}) reduces to the
475: standard forward-time discretization for $\partial u/\partial t$:
476: %----------------------------------------------------------------
477: \[
478: D_{\ast}^{1} u^n_j =
479: \frac{\partial u}{\partial t} \approx
480: \frac{u^{n+1}_j - u^n_j}{\tau}.
481: \]
482: %----------------------------------------------------------------
483:
484:
485: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
486: \subsection{Discretization of the space-fractional derivative.}
487: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
488: Just as the discretization for the time-fractional derivative assumes a simple
489: form when $\beta = 1$, the discretization
490: for the space-fractional
491: derivative, based upon centered differences,
492: assumes a simple form when $\alpha=2$.
493: For example, when $\alpha=2$ and the $N=2$,
494: %----------------------------------------------------------------
495: \[
496: D_{0}^{\alpha} u^n_j =
497: \Delta u^n_j \approx
498: \frac{1}{h^2} \left(
499: u^n_{(j_1+1,j_2)} +
500: u^n_{(j_1-1,j_2)} +
501: u^n_{(j_1,j_2+1)} +
502: u^n_{(j_1,j_2-1)} - 4 u^n_{(j_1,j_2)}.
503: \right)
504: \]
505: %----------------------------------------------------------------
506: In $N$-dimensions, the stencil consists of $j=(j_1,\ldots,j_N)$
507: and its nearest $2N$ neighbors with each nearest neighbor
508: being $h$ units away from $j$.
509: When $\alpha=\{\alpha_1,\ldots,\alpha_M\} \neq 2$, the
510: space-fractional derivative is given by \cite{UmarovSteinberg}:
511: %----------------------------------------------------------------
512: \begin{equation}
513: D_{0}^{\alpha} u^n_j \approx -q_0(\alpha,h) u^n_j +
514: \sum_{k \neq 0} q_k(\alpha,h) u^n_{j-k}
515: \label{eq:space_frac_discr}
516: \end{equation}
517: %----------------------------------------------------------------
518: where the coefficients $q_0(\alpha,h)$ and $q_k(\alpha,h)$
519: are defined in (\ref{eq:q_k}) and (\ref{eq:q_0})
520: using the multiterm definition for $a(\alpha)$.
521: The geometric consequence
522: of changing $\alpha$ from $\alpha=2$ to
523: $\alpha=\{\alpha_1,\ldots,\alpha_M\} \neq 2$ is
524: that the stencil gets enlarged from
525: $2N+1$ grid points to
526: all of the lattice points in $\Z^N_h$.
527: \vspace{0.3cm}
528:
529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
530: \subsection{Construction of the explicit finite difference scheme.}
531: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
532: Setting the
533: discretizations for the time and space-fractional derivatives equal to each
534: other in (\ref{eq:time_frac_der_discr_caputo}) and
535: (\ref{eq:space_frac_discr}), we get
536: %----------------------------------------------------------------
537: \begin{equation}
538: \frac{1}{\nu \tau^{\beta}}
539: \left( u^{n+1}_j - \sum_{m=0}^n w_m u^{m}_j \right) =
540: -q_0(\alpha,h) u^n_j +
541: \sum_{k \neq 0} q_k(\alpha,h) u^n_{j-k}.
542: \end{equation}
543: %----------------------------------------------------------------
544: Solving for $u^{n+1}_j$, the following
545: explicit finite-difference scheme is constructed:
546: %----------------------------------------------------------------
547: \begin{equation}
548: u^{n+1}_{j}
549: =
550: {\ds \sum_{m=0}^{n-1} w_m u_j^{m} +
551: \sum_{k \in \Z^N} p_k u^n_{j-k}},
552: \label{eq:time_space_frac_discr_eqn}
553: \end{equation}
554: %----------------------------------------------------------------
555: where
556: %----------------------------------------------------------------
557: \[
558: p_k = \nu \tau^{\beta} Q_k(\alpha,h), \, k \neq 0
559: \quad \mbox{and} \quad
560: p_0 = w_n - \nu \tau^{\beta} q_0(\alpha,h).
561: \]
562: %----------------------------------------------------------------
563: When $\beta = 1$, the coefficients $p_k$ are equivalent to the
564: transition probabilities $p_k$ in (\ref{eq:markovian_transprob}).
565: Furthermore, since all the transition probabilities are non-negative
566: and taking into account that $w_n = c_1 = 2-2^{1-\beta}$ and
567: $\nu = Gamma(2-\beta)$, we have an upper bound for the stepsize $\tau$:
568: %----------------------------------------------------------------
569: \[
570: p_0 \geq 0
571: \quad \Rightarrow \quad
572: 0 < \tau \leq
573: \left(
574: \frac{2 - 2^{1-\beta}} {\Gamma(2-\beta) q_0(\alpha,h)}.
575: \right)^{1/\beta}.
576: \]
577: %----------------------------------------------------------------
578:
579: The update $u^{n+1}_j$ in (\ref{eq:time_space_frac_discr_eqn})
580: is determined by Markovian contributions
581: (those values of $u$ at time $t=t_n$) and non-Markovian
582: contributions (those values of $u$ at times
583: $t=\{t_0,t_1,\ldots,t_{n-1}\}$).
584: The order of the time fractional derivative $\beta$
585: determines the effect that the non-Markovian transition
586: probabilities ($w_0,\ldots,w_{n-1}$)
587: has on $u^{n+1}_j$. This effect can be measured by
588: examining the
589: sum of all of the transition probabilities in
590: (\ref{eq:time_space_frac_discr_eqn}):
591: %----------------------------------------------------------------
592: \begin{equation}
593: {\ds \sum_{m=0}^{n-1} w_m +
594: \sum_{k \in \Z^N} p_{k} = 1},
595: \quad
596: \left\{ \begin{array}{rcl}
597: {\ds \sum_{m=0}^{n-1} w_m} & = & 1 - w_n \\
598: & & \\
599: {\ds \sum_{k \in \Z^N} p_k} & = & w_n.
600: \end{array} \right.
601: \label{eq:transprob_sum}
602: \end{equation}
603: %----------------------------------------------------------------
604: Recall that when $\beta=1$, $w_n = 1$ and
605: $w_0 = \cdots = w_{n-1} = 0$. In this case,
606: the first term in (\ref{eq:transprob_sum}) vanishes and
607: $p_0 = 1 - \tau q_0(\alpha,h)$.
608:
609: When $0<\beta <1$, the values of $u^n_j$
610: associated with $t \in \{t_0,\ldots,t_{n-1}\}$
611: are weighted by the coefficients
612: $\{w_0,w_1,\ldots,w_{n-1}\}$.
613: Figure \ref{fig:coef} plots $w_m$ for
614: $m=0,1,\ldots,n$ where $n=100$ and $\beta=0.9$.
615: It is well-known that the sequence
616: $\{w_m\}_{m=1}^{n}$ are monotone increasing \cite{Cies}, i.e.
617: $w_1 < w_2 < \ldots < w_{n-1} < w_n$.
618: However, it is not true $w_0 < w_1$. In fact,
619: in Figure \ref{fig:coef}, $w_9 < w_0 < w_8$. Hence, the
620: contribution of $u^0_j$ to $u^{101}_j$ is quite
621: large relative to the other intermediate values of $u^n_j$.
622: We will see later on that this will have important consequences in
623: non-Markovian random walk numerical simulations.
624: %----------------------------------------------------------------
625: \begin{figure}[htbp]
626: \centering
627: \includegraphics[width=5.0in]{coef.eps}
628: \caption{\small The weight $w_m$ associated with the density
629: $u_j^m$ is plotted as a function of $m$ for both the Caputo
630: and Gr\"{u}nwald-Letnikov (G.L.) time-fractional derivatives
631: and $\beta=0.9$.
632: The lower dotted horizontal line corresponds to the value of
633: $w_0 \approx 0.005$ while the upper two dotted lines correspond
634: to $w_n=c_1$ for both the Gr\"{u}nwald-Letnikov
635: ($w_{100}=0.8$) and Caputo derivatives ($w_{100} \approx 0.851$).}
636: \label{fig:coef}
637: \end{figure}
638: %----------------------------------------------------------------
639:
640: \section{Monte Carlo Protocol for the Random Walk}
641:
642: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
643: \subsection{General Framework.}
644: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
645: The random walk model corresponding to the governing equation in
646: (\ref{eq:frac_diff_eqn}) uses the
647: non-Markovian transition probabilities,
648: $w_m$ the the Markovian transition probabilities $p_k$ to assign
649: where in the $\Z_h^N$ lattice
650: a particle will jump to. This jump can be
651: based upon a partitioning of the unit interval
652: ${\mathcal{P}}=[0,1)$ into two disjoint subintervals
653: ${\mP}_1$ and ${\mP}_2$ such that
654: ${\mP} = {\mathcal{P}}_1 \cup {\mathcal{P}}_2$
655: where ${\mathcal{P}}_1 = [0,1-w_n)$ and
656: ${\mathcal{P}}_2=[1-w_n,1)$.
657:
658: We will use a two-dimensional
659: walk for illustration purposes.
660: The random walk process begins by
661: generating a uniformly distributed random number $r$
662: in the unit interval and observing what subinterval
663: (${\mathcal{P}}_1$ or ${\mathcal{P}}_2$) it falls into.
664: If $r \in {\mathcal{P}}_1 = [0,1-w_n)$, then the particle
665: will do a \emph{non-Markovian} jump, i.e. the jump will be determined
666: by transition probabilities $w_m, \, m=0,\ldots,n-1$.
667: Otherwise, if $r \in {\mathcal{P}}=[1-w_n,1)$, then the
668: particle will undergo a
669: \emph{Markovian jump}, i.e. the jump will be determined by transition
670: probabilities $p_k$.
671: In effect,
672: the random walk interpretation presented here is a two-dimensional extension
673: of the one-dimensional random walk interpretation given in \cite{GMMPP}.
674:
675:
676: \vspace{0.3cm}
677: %
678: \subsection{Non-Markovian Jumps.}
679: If $0 < \beta < 1$ and
680: $r \in {\mathcal{P}}_1$, then the jump
681: that the particle takes will be determined by
682: $w_m$, $m=0,\ldots,n-1$.
683: Let
684: ${\mathcal{A}}=\{{\mathcal{A}}_0,{\mathcal{A}}_1,\ldots,{\mathcal{A}}_{n-1}\}$
685: be an $n$-element set such that ${\mathcal{A}}_i=w_i$, $i=0,\ldots,n-1$.
686: Furthermore, let the interval
687: ${\mathcal{P}}_1$ be refined in the following way:
688: %-----------------------------------------------------------------
689: \[
690: {\mathcal{P}}_1 = [{\mathcal{B}}_0,{\mathcal{B}}_1,\ldots,
691: {\mathcal{B}}_n),
692: \]
693: %-----------------------------------------------------------------
694: such that ${\mathcal{B}}_0=0$ and
695: ${\mathcal{B}}_j = \sum_{i=0}^{j-1} {\mathcal{A}}_{i}$,
696: $j=1,\ldots,n$.
697: If $r \in [{\mathcal{B}}_0,{\mathcal{B}}_1)=[0,w_0)$, then
698: the position of the particle at $t=t_{n+1}$ is given by
699: $\bd{S}_{n+1}=\bd{S}_0$ (the origin).
700: Otherwise,
701: if $r \in [ {\mathcal{B}}_{j-1}, {\mathcal{B}}_{j} )$, $j=1,\ldots,n$,
702: then the particle will jump back to the position that
703: it had visited at time $t=t_j$, i.e.
704: $\bd{S}_{n+1}=\bd{S}_j$.
705:
706: \vspace{0.3cm}
707: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
708: \subsection{Markovian Jumps when $\alpha=2$.}
709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710: If $r \in {\mP}_2=[1-w_n,1)$ and $\alpha=2$ then
711: the jump will only be to adjacent lattice grid points.
712: Let ${\mathcal{P}}_2$ be partitioned in the following manner:
713: %-----------------------------------------------------------------
714: \[
715: {\mP}_1= [{\mB}_0,{\mB}_1,\ldots,{\mB}_5)
716: \]
717: %-----------------------------------------------------------------
718: where ${\mB}_0 = 1-w_n$ and
719: ${\mathcal{B}}_j = {\mB}_0 + \sum_{i=0}^{j-1} {\mathcal{A}}_{i}$
720: $(j=1,\ldots,5)$.
721: Here, ${\mA} = \{{\mA}_0,{\mA}_1,{\mA}_2,{\mA}_3,{\mA}_4\}$ where
722: ${\mA}_0 = w_n-4 \eta$ and
723: ${\mA}_i = \eta = \nu \tau^{\beta}/h^{\alpha}$, $i=1,2,3,4$.
724: If $r \in [{\mB}_0,{\mB}_1)$, then the particle remains at the
725: current position, otherwise if
726: $r \in
727: \{ [{\mB}_1,{\mB}_2),
728: [{\mB}_2,{\mB}_3),
729: [{\mB}_3,{\mB}_4),
730: [{\mB}_4,{\mB}_5) \}$
731: then the particle will move left, right, up or down,
732: respectively,
733: one lattice position.
734:
735: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
736: \subsection{Markovian Jumps when
737: $\alpha=\{\alpha_1,\ldots,\alpha_M\} \neq 2$.}
738: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
739: If $r \in {\mP}_2=[1-w_n,1)$
740: and $\alpha=\{\alpha_1,\ldots,\alpha_M\} \neq 2$,
741: then the jump
742: will be determined by an {\em infinite}
743: partition refinement of ${\mP}_2$.
744: Let \[
745: {\mA}=\{{\mA}_0,{\mA}_1,\ldots\}; \spc
746: {\mP}_1= [{\mB}_0,{\mB}_1,\ldots)
747: \]
748: such that ${\mathcal{B}}_0 = 1-w_n$ and
749: ${\mathcal{B}}_j = {\mB}_0 + \sum_{i=0}^{j-1} {\mathcal{A}}_{i}$
750: $(j=1,\ldots)$.
751: In this case, the set $\mA$ consists of all of the
752: transition probabilities $p_k$, $k \in \Z^2$, with
753: ${\mA}_0=p_0$.
754: If $r \in [{\mB}_0,{\mB}_1)=[1-w_n,(1-w_n)+p_0)$,
755: then the particle will remain at the current position. Otherwise,
756: if $r \in [{\mB}_s,{\mB}_{s+1})$,
757: then there exists a unique $k=(k_1,k_2) \in \Z^2$
758: associated with $s \in \N$
759: such that
760: the particle will jump from
761: $\bd{S}_n$ to $\bd{S}_{n+1} = \bd{S}_n + (k_1 h, k_2 h)$.
762:
763: \section{Simulations}
764:
765:
766: Our motivation of the numerical simulations presented here is to see how DODE
767: simulations of biomolecular motion of particles on a cell surface differ from
768: those based upon classical Brownian motion. Although the DODE random walk
769: models are described theoretically for multivariate case in $N$-dimensions,
770: nevertheless all our simulations are conducted in the two dimensional case
771: since we are interested in the diffusion of proteins on a cell membrane
772: surface, which can be locally approximated by a two-dimensional membrane
773: sheet. In \cite{Kusumi2}, simulated particle motion is based upon the
774: classical Brownian motion scenario (where $\alpha = 2$ and $\beta = 1$) in
775: which the particle is confined within cytoskeletal barriers (see Figure
776: \ref{fig:kusumi}). In these single particle tracking studies, particle
777: appears to be spatially and temporarily confined within \emph{transient
778: confinement zones}. Although the barriers are never directly observed, it is
779: postulated that the cytoskeletal barriers are the reason for the transient
780: spatial confinement of particle. In principle, DODE simulations provide an
781: alternative explanation for the observed trajectories in single particle
782: tracking studies that does not necessarily require the existence of
783: cytoskeletal barriers to explain transient confinement.
784:
785: In \cite{Kusumi2}, the authors use the mean-squared-displacement
786: formula $4 a \tau = h^2$ in which the parameters $a$ (the diffusion
787: coefficient) , $\tau$ (the timestep) and $h$ (the lattice width),
788: respectively, are given using the following values:
789: $h=6$ nanometers and $\tau = 1 \mu s$ (microseconds, or
790: $\tau=10^{-6}$ seconds). Since the
791: mean-squared displacement formula implicitly assumes that
792: %------------------------------------------------------------------------
793: \[
794: p_0 = 1 - 4 a \frac{\tau^{\beta}}{h^{\alpha}} =
795: 1 - 4a \frac{\tau}{h} = 0,
796: \]
797: %------------------------------------------------------------------------
798: the diffusion coefficient is then computed as
799: $a = h^2/(4\tau) = 9 \times 10^{-12} \mbox{m}^2/s$.
800: To facilitate a comparison of our DODE simulations with the simulations
801: of \cite{Kusumi2,Kusumi}, we will also use the same diffusion coefficient
802: ($a_1 = \cdots = a_M = a = 9 \times 10^{-12}
803: \mbox{m}^2/s$) and the same lattice width
804: ($h=6$ nanometers).
805: Using the fact that the transition probabilities sum to 1,
806: %------------------------------------------------------------------------
807: \[
808: 1 = \sum_{m=0}^{n-1} w_m + \sum_{k} p_k =
809: (1-w_n) + p_0 +
810: \nu \tau^{\beta} q_0(\alpha,h)
811: \]
812: %------------------------------------------------------------------------
813: we can now solve for $\tau$ in terms of
814: $\alpha$, $\beta$ and $p_0$,
815: %------------------------------------------------------------------------
816: \[
817: \tau = \tau(\alpha,\beta,p_0) =
818: \left( \frac{c_1 - p_0}{\nu q_0(\alpha,h)}
819: \right)^{1/\beta} =
820: \left(
821: \frac{(2-2^{1-\beta}) - p_0}{\Gamma(2-\beta) q_0(\alpha,h)}
822: \right)^{1/\beta}.
823: \]
824: %------------------------------------------------------------------------
825: As in \cite{Kusumi2}, we set
826: $p_0 = 0$.
827: However, due to the dependence of $\tau$ on $\alpha$ and $\beta$,
828: the relative
829: size of the timestep (from $\tau=10^{-6}s$ in the case of $\alpha=2$ and
830: $\beta=1$)
831: will change as $\alpha$ and $\beta$ vary.
832: Instead if fixing the simulations to have the same stepsize $\tau$, we will fix
833: the duration of the overall walk to be the same,
834: Let $T$ denote the overall duration of the random walk
835: simulation. In all of our DODE simulations, $T$ is set to
836: $T = \frac{1}{30}$ seconds. This is equivalent to 1 frame at video
837: rate where video rate is
838: measured as 30 frames per second. All simulations were performed in
839: MATLAB\cite{matlab}.
840:
841:
842: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
843: \begin{figure}[htbp]
844: \centering
845: \includegraphics[width=3.0in]{kusumi.eps}
846: \caption{\small This random walk simulation depicts
847: classical Brownian motion confined to rectangular
848: cytosketetal barriers. The parameters used in this simulation
849: are as follows:
850: $h=6$ nanometers, $\tau=10^{-6}s$ and
851: $a=9 \times 10^{-12} \mbox{m}^2/s$. The
852: barriers are spaced out every 66 nanometers and the the probability
853: of escape is $p=0.01$ when a particle encounters a barrier.}
854: \label{fig:kusumi}
855: \end{figure}
856: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
857:
858: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
859: \begin{figure}[ht]
860: \centering
861: \includegraphics[width=5.5in]{fig_1.eps}
862: \caption{\fns The first three subplots in the top row correspond to
863: Markovian DODE simulations ($\beta=1$) with different values of
864: $\alpha$: $\alpha=2$, $\alpha=1.5$ and $\alpha=\{1.5,2\}$ for
865: the left, middle and right plots. The bottom plot
866: superimposes all of the top three simulations on one graph.}
867: \label{fig:plot1}
868: \end{figure}
869: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
870:
871: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
872: \begin{figure}[ht]
873: \centering
874: \includegraphics[width=5.5in]{fig_2.eps}
875: \caption{\fns The first three subplots in the top row correspond to
876: non-Markovian DODE simulations ($\beta=0.999$) with different
877: values
878: of $\alpha$: $\alpha=2$, $\alpha=1.5$ and $\alpha=\{1.5,2\}$
879: for
880: the left, middle and right plots. The dark shaded lines
881: correspond to non-Markovian walks while the white lines
882: indicate non-Markovian jumps to previously visited positions.
883: The bottom plot
884: superimposes all of the top three simulations on one graph.
885: }
886: \label{fig:plot2}
887: \end{figure}
888: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
889:
890: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
891: \begin{figure}[hb]
892: \centering
893: \includegraphics[width=5.5in]{fig_3.eps}
894: \caption{\fns The first three subplots in the top row correspond to
895: non-Markovian DODE simulations with $\alpha=\{0.8,1.3,1.8\}$
896: and different values of $\beta$:
897: $\beta=.999$, $\beta=.99$ and $\beta=.999$
898: for
899: the left, middle and right plots. The dark shaded lines
900: correspond to non-Markovian walks while the white lines
901: indicate non-Markovian jumps to previously visited positions.
902: The bottom plot
903: superimposes all of the top three simulations on one graph.
904: }
905: \label{fig:plot3}
906: \end{figure}
907: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
908:
909: Figure \ref{fig:plot1} shows various Markovian DODE
910: simulations ($\beta=1$) across
911: various values of $\alpha$. The left, middle and right plots in the
912: top row show DODE simulations
913: for $\alpha = \{ 2 \}$, $\alpha=\{ 1.5 \}$ and
914: $\alpha = \{ 1.5, 2 \}$, respectively.
915: The first two DODE simulations
916: are actually monofractal DODE simulations with $M=1$ while the last one
917: ($\alpha=\{1.5,2\}$) is a multi-fractal case with $M=2$.
918: The large white dots indicate the first and last positions
919: of the random walk and the starting position is
920: always the origin $(0,0)$.
921: It is clear that for these DODE simulations with
922: $\alpha \not= \{2\}$ that the particle
923: travels much longer distances since
924: the probability of jumping to faraway lattice sites is greater
925: than what would be expected for $\alpha=2$.
926:
927: Figure \ref{fig:plot2} shows
928: various non-Markovian DODE simulations ($\beta=0.999$) using the same values of
929: $\alpha$ as in Figure \ref{fig:plot1}. The bottom plot in both Figures
930: \ref{fig:plot1} and \ref{fig:plot2} show the plots on top row superimposed
931: on one graph. The dark shaded lines correspond to Markovian jumps
932: ($r \in {\mP}_1$)
933: while the
934: white lines correspond to non-Markovian
935: jumps ($r \in {\mP}_2$).
936: The frequency of the non-Markovian jumps are given by the size of the
937: ${\mP}_1$ interval. For $\beta=0.999$,
938: ${\mP}_1 \approx [0,1-w_n) = [0,0.00069339)$.
939: Hence, the probability at every
940: timestep of doing a non-Markovian jump is $0.00069339$.
941: The bottom plot in Figure \ref{fig:plot2} shows the superposition all three
942: non-Markovian DODE simulations on the same graph.
943:
944: For Figure \ref{fig:plot3}, we have non-Markovian DODE simulations for
945: a fixed set of $\alpha$ values ($\alpha=\{0.8,1.3,1.8\}$) with
946: $\beta$ varying. The left, middle and right plots correspond to
947: $\beta=0.999$, $\beta=0.99$ and $\beta=0.9$, respectively.
948: The probability of taking a non-Markovian per timestep for these graphs
949: is $0.00069339$ (left), $0.0070$ (middle) and $0.0718$ (right).
950: For example, roughly 7\% of all jumps for the right subplot on the top row
951: are non-Markovian jumps. The effect of decreasing $\beta$ is clear: the
952: overall
953: distances that the particle traverses is decreased since motion is
954: constrained by jumps to previously visited positions.
955:
956: The average jump sizes
957: associated with Figures \ref{fig:plot1}, \ref{fig:plot2}
958: and \ref{fig:plot3} are shown in
959: Table \ref{tab:avg_jump}. The numbers in the brackets before the colon
960: correspond to the $(\alpha,\beta)$ pair used in the DODE simulation
961: while the number after the colon corresponds to the average jump size.
962: For the non-Markovian walks, the average jump
963: length is larger when, for a fixed set of $\alpha$ values, $\beta$ is decreased
964: from 1. This is a consequence of the non-Markovian nature of the random walks for
965: $0 < \beta < 1$.
966: Since the particle is allowed to jump back to any previously visited
967: position, the jump size can be quite large if the previously visited position was
968: spatially remote from the particle's
969: current position (see Figure \ref{fig:plot3}).
970: In particular, in Figure \ref{fig:coef}, the probability of the particle to
971: jump back to the origin is disproportionately larger than for other previously
972: visited sites. In Figures \ref{fig:plot2} and \ref{fig:plot3}, one can
973: observe evidence of this phenomenon.
974:
975: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
976: \begin{table}[h]
977: \centering
978: \caption{This table reports the average jump size (after the colon)
979: for all of the DODE
980: simulations in Figures \ref{fig:plot1}, \ref{fig:plot2} and
981: \ref{fig:plot3}. The numbers before the colon indicates values of the
982: $(\alpha,\beta)$-pair used in the DODE simulation.}
983: \begin{tabular}{|l|l|l|l|} \hline
984: &
985: {\fns Left Plot} &
986: {\fns Middle Plot} &
987: {\fns Right Plot} \\ \hline
988: {\fns Figure \ref{fig:plot1}} &
989: {\tiny (2,1): 6.0000} &
990: {\tiny (1.5,1): 10.9770} &
991: {\tiny (\{.5,2\},1): 7.3320} \\ \hline
992: {\fns Figure \ref{fig:plot2}} &
993: {\tiny (2,0.999): 6.0038} &
994: {\tiny (1.5,0.999): 11.0707} &
995: {\tiny (\{1.5,2\},0.999): 7.3593} \\ \hline
996: {\fns Figure \ref{fig:plot3}} &
997: {\tiny (\{0.8,1.3,1.8\},0.999): 17.0328} &
998: {\tiny (\{0.8,1.3,1.8\},0.99): 17.1663} &
999: {\tiny (\{0.8,1.3,1.8\},0.9): 19.8946} \\ \hline
1000: \end{tabular}
1001: \label{tab:avg_jump}
1002: \end{table}
1003:
1004: \section{Conclusion}
1005:
1006: Qualitatively, the DODE simulations provide a richer repertoire of motion,
1007: compared to monofractal walks when $M=1$. Macroscopically, the DODE
1008: trajectories tend to cluster together more often than the monofractal
1009: walks. The clustering is even more pronounced when the motion is
1010: non-Markovian due to the memory the particle has for previously visited
1011: positions. Moreover, one does not have to hypothesize the existence of
1012: barriers to explain why a particle appears trapped in a transient
1013: confinement zone or hops large distances. The clustering of trajectories
1014: and large jumps are a natural consequence of the DODE random walk model.
1015: However, when the motion is non-Markovian, the particle has a strong
1016: propensity to jump back to the origin, a consequence of the
1017: disproportionately large weight $w_0$ associated with $u^0_j$. While
1018: jumping back to previously visited ``compartments'' is observed for
1019: experimentally observed single particle tracking data \cite{Kusumi}, one
1020: does not experimentally observe molecules jumping back from its current
1021: position to the starting point. Nonetheless, the DODE random walk models
1022: closely resemble the data from single particle tracking experiments of
1023: molecules moving on cell membranes\cite{Kusumi2,Kusumi}. This is not
1024: surprising since the motion of biomolecules on the cell surface occurs in a
1025: very heterogeneous environment.
1026:
1027: \begin{thebibliography}{999}
1028:
1029:
1030:
1031: %\bibitem{ASU} \textsc{E. Andries E., S. Steinberg S., S. Umarov}, Fractional space-time differential equations: theoretical and numerical aspects. (in preparation)
1032:
1033: \bibitem{BagleyTorvic}
1034: \textsc{R.L. Bagley, P.J. Torvic} (2000).
1035: On the existence of the order domain and
1036: the solution of distributed order equations I, II.
1037: \textit{Int. J. Appl. Math} \textbf{2}, 865-882, 965-987.
1038: %\MR{1760547}
1039:
1040: \bibitem{BouchaudGeorges90}
1041: {{Bouchaud}, J. and {Georges}, A.}, {Anomalous diffusion in disordered media:
1042: Statistical mechanisms, models and physical applications}, {Physics Reports}, 1990, 195, {127-293}.
1043:
1044:
1045: \bibitem{Caputo}
1046: \textsc{M. Caputo},
1047: \textit{Elasticit\'{a} e Dissipazione},
1048: Zanichelli, Bologna, 1969.
1049:
1050:
1051: \bibitem{Caputo1}
1052: \textsc{M. Caputo} (1967).
1053: Linear models of dissipation whose Q is almost frequency independent. II.
1054: \textit{Geophys. J. R. Astr. Soc.}, \textbf{13}, 529-539.
1055:
1056:
1057: \bibitem{Caputo2}
1058: \textsc{M. Caputo} (2001).
1059: Distributed order differential equations
1060: modeling dielectric induction and diffusion.
1061: \textit{FCAA}, \textbf{4}, 421-442.
1062:
1063:
1064: \bibitem{ChechkinGonchar}
1065: \textsc{A.V.Chechkin, V. Gonchar} (1999).
1066: A model for ordinary Levy motion. Volume 1, 14pp.
1067:
1068:
1069: \bibitem{Chechkin}
1070: \textsc{A.V. Chechkin, R. Gorenflo, I.M. Sokolov, V. Gonchar} (2003).
1071: Distributed order time fractional diffusion equation.
1072: \textit{FCAA}, \textbf{6}, 259-279.
1073:
1074: \bibitem{Cies}
1075: \textsc{M. Ciesielski and J. Leszczynski} (2003).
1076: Numerical simulations of anomalous diffusion.
1077: \textit{Computer Methods in Mechanics, June 3-6, Gliwice, Poland.}
1078:
1079: \bibitem{Diethelm}
1080: \textsc{K. Diethelm K., N.J. Ford} (2001).
1081: Numerical solution methods for distributed
1082: order differential equations.
1083: \textit{FCAA}, \textbf{4}, 531-542.
1084:
1085: \bibitem{Edidin}
1086: \textsc{M. Edidin} (1997).
1087: Lipid microdomains in cell surface membranes.
1088: \textit{Curr. Opin. Struct. Biol.}, \textbf{7}, 528-532.
1089:
1090: \bibitem{GhoshWebb}
1091: \textsc{R.N. Ghosh, W.W. Webb} (1994).
1092: Automated detection and tracking of individual and clustered cell surface low
1093: density lipoprotein receptor molecules.
1094: \textit{Biophys. J.}, \textbf{66}, 1301-1318.
1095:
1096: \bibitem{GLU}
1097: \textsc{R. Gorenflo, Yu. Luchko, S.Umarov} (2000).
1098: On the Cauchy and multipoint problems for partial
1099: pseudo-differential equations of fractional order.
1100: \textit{FCAA}, \textbf{3}(3), 249-277.
1101:
1102: %\bibitem{GorenfloMainardi1} [GorenfloMainardi1 ???]
1103:
1104: \bibitem{GM}
1105: {R. Gorenflo and F. Mainardi and D.Moretti and G. Pagnini and P.Paradisi},
1106: {Discrete random walk models for space-time fractional diffusion},
1107: {Chemical Physics},
1108: {284},
1109: {521-541},
1110: {2002}.
1111:
1112:
1113: \bibitem{GMMPP}
1114: \textsc{R. Gorenflo, F. Mainardi, D. Moretti, G. Pagnini and P. Paradisi}
1115: (2002).
1116: Discrete random walk models for space-time fractional diffusion.
1117: \textit{Chemical Physics}, \textbf{84}, 521-541.
1118:
1119: \bibitem{Kusumi2}
1120: \textsc{K. Ritchie, X.-Y. Shan, J. Kondo, K. Iwasawa,
1121: T. Fujiwara, A. Kusumi} (2005).
1122: Detection of non-Brownian diffusion in the cell membrane in single
1123: molecule tracking.
1124: \textit{Biophys. J.}, \textbf{88}, 2266-2277.
1125:
1126:
1127: \bibitem{Kusumi}
1128: \textsc{K. Suzuki, K. Ritchie, E. Kajikawa, T. Fujiwara, A. Kusumi} (2005).
1129: Rapid Hop Diffusion of a G-Protein-Coupled Receptor in the Plasma
1130: Membrane as Revealed by Single-Molecule Techniques.
1131: \textit{Biophys. J.}, \textbf{88}, 3659-3680.
1132:
1133: \bibitem{LSAT_05}
1134: \textsc{F. Liu, S. Shen, V.Anh, I. Turner} (2005).
1135: Analysis of a discrete non-Markovian random
1136: walk approximation for the time fractional diffusion equation.
1137: \textit{ANZIAM J.}, \textbf{46}, C488-C504.
1138:
1139: \bibitem{LorenzoHartley}
1140: \textsc{C.F. Lorenzo, T.T. Hartley} (2002).
1141: Variable order and distributed order fractional operators.
1142: \textit{Nonlinear Dynamics}, \textbf{29}, 57-98.
1143: %\MR{1926468}
1144:
1145: \bibitem{matlab}
1146: MATLAB. User's Guide. The MathWorks, Inc. Natick, MA 01760, 1992.
1147:
1148:
1149: \bibitem{MMM}
1150: \textsc{M. Meerschaert, P. Scheffler} (2005).
1151: Limit theorems for continuous time random walks with
1152: slowly varying waiting times,
1153: \textit{Statistics and Probability Letters}.
1154: \textbf{71}(1), 15-22.
1155:
1156: \bibitem{MS}
1157: \textsc{M. Meerschaert, P. Scheffler}.
1158: \textit{Limit Distributions for Sums of Independent Random Vectors.
1159: Heavy Tails in Theory and Practice}.
1160: John Wiley and Sons, Inc, New York, 2001.
1161: %\MR{1840531}
1162:
1163: \bibitem{MK}
1164: \textsc{R. Metzler, J. Klafter} (2000).
1165: The random walk's guide to anomalous
1166: diffusion: a fractional dynamics approach.
1167: \textit{Physics Reports} \textbf{339}, 1-77.
1168: %\MR{1809268}
1169:
1170: \bibitem{SKM}
1171: \textsc{S.G. Samko, A.A. Kilbas, O.I. Marichev},
1172: \textit{Fractional Integrals
1173: and Derivatives: Theory and Applications}.
1174: Gordon and Breach Science Publishers, New York and London, 1993.
1175:
1176: \bibitem{Saxton01}
1177: \textsc{M. Saxton} (2001).
1178: Anomalous Subdiffusion in Fluorescence Photobleaching Recovery:
1179: A Monte Carlo Study,
1180: \textit{Biophys. J.}, \textbf{81}(4), 2226-2240.
1181:
1182: \bibitem{SaxtonJacobson97}
1183: \textsc{M.J. Saxton, K. Jacobson} (1997).
1184: Single-particle tracking: applications to membrane dynamics.
1185: \textit{Ann. Rev. Biophys. Biomol. Struct.}, \textbf{26}, 373-399.
1186:
1187:
1188: \bibitem{UchaykinZolotarev}
1189: \textsc{V.V. Uchaykin, V.M. Zolotarev},
1190: Chance and Stability. Stable Distributions and their Applications,
1191: \textit{VSP}, Utrecht, 1999.
1192:
1193: \bibitem{UG05-1}
1194: \textsc{S. Umarov, R. Gorenflo} (2005).
1195: On multi-dimensional symmetric random walk models approximating
1196: fractional diffusion processes.
1197: \textit{FCAA}, \textbf{8}, 73-88.
1198:
1199: \bibitem{UG05-2}
1200: \textsc{S. Umarov, R. Gorenflo} (2005).
1201: The Cauchy and multipoint problem for distributed order fractional
1202: differential equations.
1203: \textit{ZAA}, \textbf{24}, 449-466.
1204:
1205: \bibitem{UmarovSteinberg}
1206: \textsc{S. Umarov, S. Steinberg} (2006).
1207: Random walk models associated with distributed fractional
1208: order differential equations.
1209: \textit{IMS Lecture Notes - Monograph Series}, (to appear).
1210:
1211: \bibitem{Zaslavsky}
1212: \textsc{G. Zaslavsky} (2002).
1213: Chaos, fractional kinetics, and anomalous transport.
1214: \textit{Physics Reports}, \textbf{371}, 461-580.
1215: %\MR{1937584}
1216:
1217:
1218: \end{thebibliography}
1219:
1220: \vskip 1cm
1221: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1222: \end{document}
1223: