1:
2: \documentclass[pre,twocolumn,endfloats,amsmath]{revtex4}
3: %\documentclass[preprint,pre,endfloats,amsmath]{revtex4}
4:
5:
6: \usepackage{graphicx}
7: \usepackage{psfrag}
8: %\usepackage{epsfig}
9: %\input psfig
10: %\input{epsf}
11: %
12: %
13: %\textwidth 140mm
14: %\oddsidemargin 15mm
15: %\evensidemargin 3mm
16:
17: \newcommand{\ket}[1]{| #1 \rangle}
18: \newcommand{\bra}[1]{\langle #1 |}
19: \newcommand{\eq}[1]{Eq.~(\ref{#1})}
20: \newcommand{\eqs}[1]{Eqs.~(\ref{#1})}
21: \newcommand{\fig}[1]{Fig.~\ref{#1}}
22: \newcommand{\app}[1]{Appendix~\ref{#1}}
23: \newcommand{\sect}[1]{Section~\ref{#1}}
24: \newcommand{\dotprod}[2]{dot(#1,#2)}
25: \newcommand{\SecRef}[1]{section~\ref{#1}}
26:
27: \begin{document}
28: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
29: \title{Diffusive Transport in Networks Built of Containers and Tubes}
30:
31: \author{L. Lizana and Z. Konkoli\footnote{Address: Chalmers University of
32: Technology, SE-412 96 G\"{o}teborg, Sweden. E-mail:
33: \texttt{zorank@fy.chalmers.se}. Tel. +46 31 772 3186; Fax: +46 31 41
34: 69 84. \\}}
35:
36: \affiliation{Department of Applied Physics, Chalmers University of
37: Technology and G{\"o}teborg University}
38:
39: \date{\today}
40:
41: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
42: \begin{abstract}
43: We developed analytical and numerical methods to study a transport of
44: non-interacting particles in large networks consisting of $M$
45: $d$-dimensional containers $C_1,\ldots,C_M$ with radii $R_i$ linked
46: together by tubes of length $l_{ij}$ and radii $a_{ij}$ where
47: $i,j=1,2,\ldots,M$. Tubes may join directly with each other forming
48: junctions. It is possible that some links are absent. Instead of
49: solving the diffusion equation for the full problem we formulated an
50: approach that is computationally more efficient. We derived a set of
51: rate equations that govern the time dependence of the number of
52: particles in each container $N_1(t),N_2(t),\ldots,N_M(t)$. In such a
53: way the complicated transport problem is reduced to a set of $M$ first
54: order integro-differential equations in time, which can be solved
55: efficiently by the algorithm presented here.
56: The workings of the method have been demonstrated on a couple of
57: examples: networks involving three, four and seven containers, and one
58: network with a three-point junction. Already simple networks with
59: relatively few containers exhibit interesting transport behavior. For
60: example, we showed that it is possible to adjust the geometry of the
61: networks so that the particle concentration varies in time in a
62: wave-like manner. Such behavior deviates from simple exponential
63: growth and decay occurring in the two container system.
64: \end{abstract}
65:
66: \maketitle
67:
68: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69: \section{Introduction} \label{Introduction}
70:
71: The goal of this work is to find a method that describes diffusive
72: transport of particles on a network built up of spherical containers
73: connected by tubes. It is possible that not all containers are
74: connected to each other and tubes may join together forming junctions
75: (without a container being present). In such a way one can generate an
76: enormous number of network topologies. One example is show in
77: \fig{MainNetwork}. The networks consists of $M$ containers
78: (reservoirs) labeled $C_1,\ldots,C_M$ of radii $R_i$ connected by
79: tubes of length $l_{ij}$ and radii $a_{ij}$ where $i,j=1,2,\ldots,M$.
80:
81: Our work is motivated by experiments discussed in
82: refs.~\cite{Orwar,KSMDKO,KDKKBKJLHVO} and ~\cite{LPHR,HR,HSR}. The
83: first set of references describes how to create and manipulate
84: microscale sized compartments (vesicles) connected by nanotubes. These
85: structures can be applied in a number of ways \cite{KDKKBKJLHVO}. For
86: example, the non-diffusive (forced) transport was studied in
87: \cite{Orwar}. In this work we study different kind of transport were
88: only passive diffusion is allowed. Also, by including the reactions in
89: the theoretical setup one could describe biochemical reactions in a
90: milieu close to their natural habitat. This idea was pursued
91: experimentally in ref. \cite{KSMDKO}.
92:
93: The second set of references deals with networks of chemical reactors
94: of macroscopic sizes connected to each other by tubes. The exchange of
95: reactants is mediated through the tubes and controlled by pumps. It
96: was shown that it is possible to use this device to carry out pattern
97: recognition tasks. In making the device smaller the external pumping
98: can be removed and the transport can be limited to pure
99: diffusion. This kind of setup is close to the situation studied here.
100:
101: For large networks, links connecting opposite sides of the network may
102: be rather long. Accordingly, one can not expect an exponential decay
103: of the number of particles in the containers and this is the situation
104: we are mostly interested in. Obviously, to describe such a situation
105: one can attempt to solve the diffusion equation numerically and obtain
106: the distribution function $\rho(\vec{r},t)$ that describes how
107: particles spread throughout the network.
108:
109: Obtaining the solution of the diffusion equation for a complicated
110: geometry for a large networks gets highly impractical. In this work
111: we develop a method of calculation that is computationally efficient
112: and can be used to study transport in large networks. Instead of
113: finding the full distribution function $\rho(\vec{r},t)$ we introduce
114: a set of slow variables that capture the most important aspect of
115: particle transport, the number of particles in each container
116: $N_1(t),N_2(t),\ldots,N_M(t)$, and derive equations that describe how
117: they change in time. This is the central result of the paper.
118:
119:
120: %The details of the dynamics within the containers is projected out. In
121: %such a way one reduces the amount of information, e.g. the
122: %distribution profile of $\rho(\vec{r},t)$ within the container can not
123: %be traced, but there is an advantage from a computational point of
124: %view since the number of dynamical variables is reduced from an
125: %infinite number to $M$. In principle, the distribution of particles in
126: %the tubes $\rho(\vec{r},t)$ has to be calculated as well but it is
127: %possible to describe this part analytically and project it out.
128:
129: A couple of related problems have been addressed previously in
130: refs. \cite{Grigoriev, Berezhkovskii, Dagdug, Dagdug2, Bezrukov}.
131: Escape of a particle through a small hole in a cavity was studied in
132: \cite{Grigoriev}. The work of \cite{Berezhkovskii} deals with the
133: problem of the hole connected to a short tube. The tube length and
134: hole radius are roughly of the same size, mimicking a cell membrane
135: having a thickness greater than zero. A couple of equilibration cases
136: have also been studied \cite{Dagdug,Dagdug2,Bezrukov}. The papers just
137: indicated treat the intra container dynamics in much more detail than
138: we do. In here, for simplicity reasons, the particle concentration is
139: assumed flat in the container and the validity of this approximation
140: is checked numerically in section \ref{IdealMixing}. In our notation,
141: the studies \cite{Grigoriev, Berezhkovskii, Dagdug, Dagdug2, Bezrukov}
142: can be classified as $M=1,2$ cases. Our main interest is in networks
143: with large $M$.
144:
145:
146: This paper is organized as follows. In section \ref{ProblemDef} the
147: problem is defined and the general results are stated. The derivation
148: of the rate \eq{GeneralEq} is explained in sections \ref{SecEmpt} and
149: \ref{SecRate} where the emptying of a single container into a tube,
150: and emptying of a container into another container through the tube is
151: studied. The single exponential asymptotics of the two container
152: system and related first order rate equations are found and discussed
153: in section \ref{SingExpSec}. Up to this point only a two-container
154: system is treated while section \ref{SecGeneralEqs} deals with an
155: arbitrary network topology. Section \ref{IdealMixing} elaborates on
156: the assumption of well stirred containers. A numerical comparison to
157: the diffusion equation is made. In section \ref{CaseStudies} numerical
158: case studies of various network structures are performed. In
159: particular a three way junction and an example of a larger network are
160: studied. The summary and outline of future work is given in section
161: \ref{Conclusions}. Technical details are found in the
162: appendices. Appendix \ref{Numerical} describes the numerical procedure
163: used for solving the rate equations. The rate equations for the cases
164: studied in \sect{CaseStudies} are explicitly derived in
165: \app{CaseRates}. It can be shown that the presence of tube junctions
166: can be eliminated altogether from the dynamical equations when the
167: time is large. This is demonstrated in \app{SectJunction}.
168:
169:
170:
171:
172: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
173: \section{Problem definition and Main Result}\label{ProblemDef}
174: Describing the particle transport in a network depicted in
175: \fig{MainNetwork} is far from trivial and in order to solve the
176: problem several assumptions are made. We assume that (i) particles
177: move solely by diffusion (the fluid in which the particles move stands
178: still) and (ii) particles do not disturb each other. With these
179: assumptions the complicated dynamical problem at hand is reduced to
180: solving the time dependent diffusion equation:
181: %
182: \begin{equation}\label{DiffEq}
183: \partial_t \rho(\vec{r},t) = \nabla \cdot \left[D(\vec{r})\nabla
184: \rho(\vec{r},t)\right].
185: \end{equation}
186: %
187: Here $\rho(\vec{r},t)$ is the concentration (particle density) and
188: $D(\vec{r})$ is the diffusion coefficient which may be position
189: dependent. The walls are particle impenetrable
190: %
191: \begin{equation}
192: \partial_n \rho(\vec{r},t) = 0
193: \end{equation}
194: %
195: where $\partial_n \equiv \hat{n}\cdot \nabla$, and $\hat{n}$ is the
196: unit vector perpendicular to the wall. The total number of particles
197: is a conserved quantity.
198:
199: Equation (\ref{DiffEq}) could in principle be solved numerically using
200: a brute force approach (e.g. the Finite Element Method or the Finite
201: Difference method). However, in sections \ref{SecEmpt} and
202: \ref{SecRate} we will show that it is possible to describe particle
203: transport in terms of a finite number of variables, the number of
204: particles in each container $N_1,...,N_M$. Also, it might be easier to
205: understand particle transport in such a setup. The dynamics of
206: $N_i(t)$ $i=1,\ldots, M$ is governed by
207: %
208: \begin{equation}\label{GeneralEq}\begin{array}{lll}
209: \dot{N}_i(t) & = & \sum_{j=1}^M {\cal C}_{ji}\frac{V_{d-1}(a_{ji})}{V_d(R_j)}
210: \int_0^t dt'{\cal N}_j(t')\sigma_{ji}(t-t')\\\\
211:
212: & & -\sum_{j=1}^M {\cal C}_{ij}\frac{V_{d-1}(a_{ij})}{V_d(R_i)}
213: \int_0^t dt'{\cal N}_i(t')\left[\Delta_{ij}(t-t')\right.\\\\
214: & & \left. +\kappa_{ij}(t-t')\right]
215: \end{array}\end{equation}
216: where
217: %
218: \begin{equation}
219: {\cal N}_i(t) \equiv\dot{N}_i(t)+N_{i0}\delta(t)
220: \end{equation}
221: %
222: and $\delta(t)$ is the Dirac delta-function, \mbox{$\int_0^\infty
223: dt\,\delta(t) = 1$}. Here and in the following the dot over symbol
224: denotes time derivative. The connectivity matrix \mbox{${\cal C}_{ij} \in
225: \{0,1\}$} describes how the nodes are linked (note that \mbox{${\cal
226: C}_{ii}=0$)}, $a_{ij}$ is the radius of the tube (link) from $i$ to
227: $j$ and $V_d(R_j)$ is the volume of a $d$ dimensional sphere
228: \mbox{$V_d(r) = [2\pi^{d/2} /d\Gamma(d/2)]r^d$} corresponding to
229: container $j$ having radius $R_j$. $N_{i0}$ denotes
230: $N_i(t=0)$. Equation (\ref{GeneralEq}) is derived under the assumption
231: of ideally mixed containers. The rate coefficients are given by
232: %
233: \begin{equation}\label{GeneralRateCoef}\begin{array}{lll}
234: \Delta_{ij}(t) &= & \displaystyle\sqrt{\frac{D_{ij}}{\pi t}}\\\\
235:
236: \kappa_{ij}(t) &= &\displaystyle2\sqrt{\frac{D_{ij}}{\pi
237: t}}\sum_{k=1}^\infty\exp\left(-\frac{(k \ell_{ij})^2}{D_{ij}t}\right)\\\\
238:
239: \sigma_{ij}(t) &= &\displaystyle2\sqrt{\frac{D_{ij}}{\pi t}}\sum_{k=0}^\infty
240: \exp\left(-\frac{((2k-1) \ell_{ij})^2}{4D_{ij}t}\right).
241: \end{array}\end{equation}
242: %
243: $\ell_{ij}$ is the link length and $D_{ij}$ is the corresponding
244: diffusion coefficient. The theory is developed for the general
245: case where the diffusion constant in each tube may be different.
246:
247: Figure \ref{MainNetwork} also shows the existence of tube
248: junctions. They are treated by \eq{GeneralEq} by letting the container
249: radius coincide with that of the tube. Since the tubes are initially
250: empty, so are junctions, $N_{i0}=0$.
251: Eqs. (\ref{GeneralEq})-(\ref{GeneralRateCoef}) are the central results
252: of this paper and their derivation is a major topic of the subsequent
253: sections.
254:
255:
256: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
257: \section{Emptying of a reservoir through an infinitely long tube}
258: \label{SecEmpt}
259:
260: To derive \eq{GeneralEq} we start with the simplest possible case and
261: consider particle escape from a container through an infinitely long
262: tube [see \fig{InfTubeFig}, panel (a)]. The main reason for this is to
263: show how to couple the dynamics of the tube and the container. Also,
264: such setup captures the short time description of the full network
265: problem when the particles escaping the containers do not yet 'feel'
266: that the system is closed [the short time dynamics is contained in
267: $\Delta(t)$, see \sect{SecRate}].
268:
269: The particle concentration $\rho(\vec{r},t)$ is governed by the
270: diffusion equation supplemented with the boundary conditions that the
271: walls are impenetrable and that $\rho(\vec{r},t)$ has to vanish for
272: $x\rightarrow \infty$. The concentration in the tube and in the
273: container are interwoven in a highly non-trivial way through what is
274: occurring at the tube opening. Given that the current density out of
275: the container $j(0,y,z,t)$ is known [see \fig{InfTubeFig} (a)] one
276: could solve the diffusion problem and find the concentration profile
277: in the container and the number of particles. Furthermore, one could
278: find a relationship
279: %
280: \begin{equation}\label{Functional}
281: \rho(0^-,y,z,t) = {\cal F} [j(0,y,z,t)]
282: \end{equation}
283: %
284: where $j(0,y,z,t) = -D\,\lim_{x\rightarrow 0^-}\hat{x}\cdot \nabla
285: \rho(x,y,z,t)$. ${\cal F}$ is a functional that we know exists but is
286: unlikely to be found in a closed analytic form.
287:
288: To find $j(0,y,z,t)$ it is assumed that the concentration in the
289: vicinity of the tube opening can be approximated
290: with
291: %
292: \begin{equation}\label{DeCoupling}
293: \rho(x,y,z,t) = f(y,z,t) c(x,t) \,\,\,\,\ x\gtrsim 0
294: \end{equation}
295: %
296: where $c(x,t)$ is a one-dimensional concentration and $f(z,y,t)$ is a
297: function that projects the value of $c(x,t)$ onto a radial direction.
298: Equation~(\ref{DeCoupling}) is valid for large $x$ when
299: $f(y,z,t)$ is constant but not in general case. Arbitrary density
300: profile at the opening will in time smear out due to radial
301: diffusion.
302:
303: By assumption, the concentration profile in the tube is governed by
304: $c(x,t)$ and it is coupled to the concentration in the container as
305: follows. Both concentration and current have to be continuous as one
306: moves from the container into the tube leading to
307: %
308: \begin{equation}\label{BC}
309: \rho(0,y,z,t) = f(y,z,t) c(0,t).
310: \end{equation}
311: %
312: and
313: %
314: \begin{equation}\label{BCFlow}
315: j(0,y,z,t) = f(y,z,t) \lim_{x\rightarrow 0^+} \frac{\partial}{\partial
316: x} c(x,t)
317: \end{equation}
318: %
319: Please note that $\lim_{x\rightarrow 0^+} \frac{\partial}{\partial
320: x}c(x,t)$ in \eq{BCFlow} is a functional of $c(0,t)$. Taking into
321: account particle conservation at the tube opening leads to
322: %
323: \begin{equation}\label{PartConservation}
324: \int_{x=0} dS \;\rho(0,y,z,t) = c(0,t)
325: \end{equation}
326: %
327: which after using \eq{BC} results in the condition $\int dS\, f(y,z,t)
328: = 1$. The problem has four unknowns $\rho(0,y,z,t)$, $j(0,y,z,t)$,
329: $c(0,t)$ and $f(y,z,t)$ and four equations (\ref{Functional}),
330: (\ref{BC})-(\ref{PartConservation}) and is fully defined. However, it
331: is not tractable in this complicated form and we proceed to simplify
332: it.
333:
334: Instead of $\rho(\vec{r},t)$ a more useful variable is the total
335: number of particles in the container
336: $N(t)=\int_{x<0}dV\rho(\vec{r},t)$ governed by
337: %
338: \begin{equation}\label{CouplingInf}
339: \dot{N}(t) = -J(t)
340: \end{equation}
341: %
342: where $J(t)$ is the flow of particles that leave the container through
343: the tube opening $J(t) = \int dS j(0,y,z,t)$. There are two special
344: cases where the current $J(t)$ can be determined analytically in terms
345: of container variables. (i) If the exit is a fully absorbing disk with
346: radius $a$, the concentration is always zero at the the interface
347: $\rho(0,y,z,t) = 0$. In \cite{BePu} it is shown that the current
348: $J_\infty$ through such disk when placed at an infinite otherwise
349: reflecting wall is $J_\infty = 4D_c a\rho_\infty$ where $\rho_\infty$
350: is the particle concentration at infinity and $D_c$ denotes the
351: diffusion constant in the container. A reasonable assumption for the
352: container, at least when the tube radius is smaller than the radius of
353: the container, is that the concentration profile far away from the
354: exit is flat and can approximately be taken to be $N(t)/V_d(R)$. Using
355: this for $\rho_\infty$ yields $J_\infty(t) = 4D_caN(t)/V_d(R)$. (ii)
356: If the opening is completely closed, the current is zero and
357: $\rho(0,y,z,t) = N(t)/V_d(R)$ [in such a case $N(t)= N(0)$]. A linear
358: interpolation between (i) and (ii) yields
359: %
360: \begin{equation}\label{LinearInterpol}
361: \rho(0,y,z,t) = \frac{N(t)}{V_d(R)}\left[1-\frac{J(t)}{J_\infty(t)}\right].
362: \end{equation}
363: %
364: Please note that in Eq.~(\ref{LinearInterpol}) $\rho(0,y,z,t)$ is
365: assumed constant across the interface. This approximation is verified
366: numerically in \fig{InletProfile} that shows $f(y,z,t)\approx
367: V_{d-1}(a)^{-1}$. Also, when $a\ll R$ one has $J(t)/J_\infty(t)\ll 1$
368: and second term in (\ref{LinearInterpol}) can be neglected. This
369: approximation is verified by numerical calculations in
370: \sect{IdealMixing}.
371:
372: At this point, the full problem has been mapped on to a very simple
373: geometry depicted in \fig{InfTubeFig} (b): a one dimensional line
374: (tube) connected to a point (container). Tube dynamics is
375: characterized by a one dimensional particle density $c(x,t)$ along the
376: line and all container dynamics, how complicated it may be, has been
377: projected on to a single variable $N(t)$.
378:
379: The only part of the problem that remains to be solved is the
380: diffusion through the tube. This part of the problem can be
381: approximated by one dimensional diffusion since $f(y,z,t)$ is
382: constant. The constant concentration profile at the tube opening
383: remains such in the tube interior (provided the tube radius does not
384: change along $x$ direction). With assumptions at hand the coupling
385: \eq{BC} becomes
386: %
387: \begin{equation}\label{BC2}
388: c(0,t) = V_{d-1}(a)\frac{N(t)}{V_d(R)}.
389: \end{equation}
390: %
391: The concentration profile along the tube [initially empty $c(x,0)=0$]
392: is given by the diffusion equation
393: %
394: \begin{equation}
395: \frac{\partial c(x,t)}{\partial t} = D \frac{\partial^2
396: c(x,t)}{\partial x^2}\;\;\;\;\;\;x\in[0,\infty)
397: \end{equation}
398: %
399: supplemented with boundary conditions according to \eq{BC2} and
400: $c(\infty,t)=0$. The solution can be found by the Laplace transform
401: method \cite{Schaum} and is given by
402: %
403: \begin{equation}\label{SolInf}
404: c(x,s)=c(0,s)\,e^{-x\sqrt{s/D}}.
405: \end{equation}
406: %
407: where $c(x,s)=\int_0^\infty dt \,c(x,t) e^{-st}$. Integrating
408: \eq{BCFlow} over the tube interface area at $x=0$ gives
409: %
410: \begin{equation}\label{FicksLaw}
411: J(t) = -D\lim_{x\rightarrow 0}\frac{\partial}{\partial x}c(x,t).
412: \end{equation}
413: %
414: Combining \eqs{CouplingInf}, (\ref{BC2}) and (\ref{FicksLaw}) leads to
415: a rate equation in the Laplace transform space
416: \begin{equation}\label{CouplingInfInS}\begin{array}{lll}
417: sN(s)-N_0 &= &\displaystyle-\lim_{x\rightarrow 0}\sqrt{sD}N(s)
418: \frac{V_{d-1}(a)}{V_d(R)}\, e^{-x\sqrt{s/D}}\\\\ &=
419: &\displaystyle-\sqrt{sD}N(s)\frac{V_{d-1}(a)}{V_d(R)}
420: \end{array}\end{equation}
421: %
422: where ${\cal L}[\dot{N}(t)]=sN(s)-N_0$. It is tempting to rewrite this
423: equation in the time domain in the form of convolution
424: %
425: \begin{equation}\label{GenRateEq}
426: \dot{N}(t) = -\int_0^t dt' k(t')N(t-t')
427: \end{equation}
428: %
429: representing a general form of a rate law, where
430: $k(t)=\frac{V_{d-1}(a)}{V_d(R)} {\cal L}^{-1}[\sqrt{Ds}]$. However,
431: this is impossible and can be seen in several ways.
432:
433: First, $\sqrt{s}$ has no well defined inverse Laplace transform and
434: $k(t)$ that would enter into the rate equation (\ref{GenRateEq}) is
435: ill-defined. Second, this problem could possibly be resolved by
436: inverting \eq{SolInf} to obtain $c(x,t)$ and inserting the result into
437: \eq{FicksLaw} which leads to
438: %
439: \begin{equation}\label{Illegal}
440: J(t) \propto -\lim_{x\rightarrow 0}\frac{\partial}{\partial x}\int_0^t
441: dt' \,\frac{x}{t'^{3/2}}\, e^{-x^2/4Dt'} N(t-t').
442: \end{equation}
443: %
444: In general $N(t)$ is unknown. To evaluate the expression above in a
445: way that would result in a rate equation involves interchanging
446: derivation and integration. This is only allowed if the integral is
447: uniformly convergent in the interval $x\in[0,\infty)$
448: \cite{LePage}. It is easy to see from \eq{Illegal} that this is not
449: the case and the interchange is illegal. Another possibility is to
450: use partial integration but this strategy does not work since one ends
451: up with non-convergent integrals as $x\rightarrow 0$. Thus,
452: \eq{GenRateEq} does not exist for a semi-infinite case. Also, it is
453: intuitively clear that one can not observe pure exponential decay
454: since the system is infinite.
455:
456: For an infinite system an asymptotic rate law of the type $\dot{N}(t)
457: \propto -N(t)$ simply does not exist. This can also be seen from the
458: exact expression for $N(t)$ which can be obtained from
459: \eq{CouplingInfInS} by solving for $N(s)$ and finding the inverse
460: Laplace transform \cite{Schaum}
461: %
462: \begin{equation}\label{SolInfInT}
463: N(t) = N_0\exp\left[ Dt \left(\frac{V_{d-1}(a)}{V_d(R)}
464: \right)^2\right] {\rm erfc} \left[\sqrt{Dt} \frac{V_{d-1}(a)}{V_d(R)}
465: \right].
466: \end{equation}
467: %
468: $N_0$ is the initial number of particles in the container. Using
469: approximation erfc$(z)\approx \frac{e^{-z^2}}{\sqrt{\pi}} \frac{1}{z}$
470: for large $z$ gives $N(t) \propto \frac{1}{\sqrt{t}}$.
471:
472: Due to the complications discussed above the rate equation has to be
473: stated in terms of $\dot{N}(t)$
474: %
475: \begin{equation}\begin{array}{lll} \label{AlmostRate}
476: \dot{N}(t) &= &-\displaystyle \frac{V_{d-1}(a)}{V_d(R)}\, {\cal L
477: }^{-1}\left[sN(s) \sqrt{\frac{D}{s}}\right]\\\\
478: &= &-\displaystyle\frac{V_{d-1}(a)}{V_d(R)} \int_0^t dt' {\cal
479: N}(t') \Delta(t-t')
480: \end{array}\end{equation}
481: %
482: where
483: %
484: \begin{equation}
485: {\cal N}(t)\equiv {\cal L }^{-1}\left[sN(s)\right] = \dot{N}(t) +
486: N_0\delta(t)
487: \end{equation}
488: %
489: and
490: %
491: \begin{equation}\label{DefOfDelta}
492: \Delta(t) \equiv {\cal L}^{-1}\left[\sqrt{\frac{D}{s}}\right] =
493: \sqrt{\frac{D}{\pi t}}.
494: \end{equation}
495: %
496: Please note that it is impossible to rewrite the right hand side of
497: \eq{AlmostRate} in such a way that it would solely involve dependence
498: on $N(t)$. When the system is closed (e.g. by adding another
499: container) the situation changes.
500:
501:
502:
503: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
504: \section{A Rate Equation for a Two Container System} \label{SecRate}
505: The system under consideration consists of a one dimensional rod
506: parameterized by $a$ and $\ell$ connected to two ideally mixed
507: containers having radii $R_1$ and $R_2$ depicted in
508: \fig{TwoContSytem}.
509:
510: The diffusion equation for the tube [initially empty, $c(x,t) = 0$]
511: connected to the two containers at $x=0$ and $x=\ell$ is given by
512: %
513: \begin{equation}
514: \frac{\partial c(x,t)}{\partial t} = D\frac{\partial^2
515: c(x,t)}{\partial x^2}\;\;\;\;\;\;x\in(0,\ell)
516: \end{equation}
517: %
518: with boundary conditions analogous to \eq{BC2}
519: %
520: \begin{equation}
521: \begin{array}{lr}
522: c(0,t) = N_1(t)\frac{V_{d-1}(a)}{V_d(R_1)}, & c(\ell,t) =
523: N_2(t)\frac{V_{d-1}(a)}{V_d(R_2)}.
524: \end{array}\end{equation}
525: %
526: The solution in Laplace transform space is given by
527: %
528: \begin{equation}\label{SolInS}
529: \displaystyle c(x,s) = \frac{V_{d-1}(a)}{V_d(R_1)} sN_1(s) \phi_1(x,s) +\frac{V_{d-1}(a)}{V_d(R_2)}
530: sN_2(s)\phi_2(x,s)
531: \end{equation}
532: %
533: where
534: %
535: \begin{equation}\begin{array}{lr}
536: \phi_1(x,s) = \frac{\sinh\left((x-\ell)\sqrt{s/D}\right)}
537: {s\sinh\left(\ell\sqrt{s/D}\right)}, &
538:
539: \phi_2(x,s) = \frac{\sinh\left(x\sqrt{s/D}\right)}
540: {s\sinh\left(\ell\sqrt{s/D}\right)}.
541: \end{array}\end{equation}
542: %
543: Matching the current at both ends as in \eq{FicksLaw}
544: %
545: \begin{equation}\begin{array}{llr}
546: \dot{N}_1(t) &=& D\,\lim_{x\rightarrow 0}\frac{\partial}{\partial
547: x}c(x,t)
548: \\\\
549: \dot{N}_2(t) &=& - D\,\lim_{x\rightarrow \ell}\frac{\partial}{\partial
550: x}c(x,t)
551:
552: \end{array}\end{equation}
553: %
554: yields a set of rate equations in Laplace transform space for the two
555: container system
556: %
557: {\small
558: \begin{subequations}
559: \label{RateInS}
560: \begin{eqnarray}
561: sN_1(s)-N_{10} & = & -sN_1(s)\frac{V_{d-1}(a)}{V_d(R_1)}
562: \sqrt{\frac{D}{s}}\frac{\cosh\left(\ell\sqrt{s/D}\right)}
563: {\sinh\left(\ell\sqrt{s/D}\right)}\nonumber\\
564: & & +sN_2(s)\frac{V_{d-1}(a)}{V_d(R_2)}\sqrt{\frac{D}{s}}
565: \frac{1}{\sinh\left(\ell\sqrt{s/D}\right)}\nonumber\\
566: & &\\
567: sN_2(s)-N_{20} & = & -sN_2(s)\frac{V_{d-1}(a)}{V_d(R_2)}
568: \sqrt{\frac{D}{s}}\frac{\cosh\left(\ell\sqrt{s/D}\right)}
569: {\sinh\left(\ell\sqrt{s/D}\right)}\nonumber\\
570: & & +sN_1(s)\frac{V_{d-1}(a)}{V_d(R_1)}\sqrt{\frac{D}{s}}
571: \frac{1}{\sinh\left(\ell\sqrt{s/D}\right)}.\nonumber\\
572: & &
573: \end{eqnarray}
574: \end{subequations}
575: }
576: %
577: The terms having minus signs are the outflow while the ones having
578: plus signs represent the inflow. A closer look at the the rate
579: equations above reveals an important link to the semi-infinite case:
580: the outflow is proportional to $\sqrt{s}$ for large $t$ (small
581: $s$). In other words, the semi-infinite case is recovered as a short
582: time expansion of \eqs{RateInS} (see \sect{SecEmpt}). The physical
583: interpretation is that initially, the particles feel as if they were
584: entering an infinitely long tube. This implies that the outflow term
585: can be divided into two parts reflecting this observation
586: %
587: \begin{equation}\begin{array}{l}\label{DeltaSigma}
588: \sqrt{\frac{D}{s}}\left[\frac{\cosh\left(\ell\sqrt{s/D}\right)}
589: {\sinh\left(\ell\sqrt{s/D}\right)}\right] \equiv \Delta(s) + \kappa(s).
590: \end{array}\end{equation}
591: %
592: $\Delta(s)$ is taken from \eq{DefOfDelta} and controls the short time
593: dynamics that resembles the one of the semi-infinite
594: system. $\kappa(s)$ describes the asymptotic long time behavior when
595: the particles are 'aware' of the existence of another side and can be
596: found from \eq{DefOfDelta} and \eq{DeltaSigma}
597: %
598: \begin{equation} \label{RateCoefInS1}
599: \kappa(s) = \sqrt{\frac{D}{s}}\frac{\exp\left(-\ell\sqrt{s/D}\right)}
600: {\sinh\left(\ell\sqrt{s/D}\right)}.
601: \end{equation}
602: %
603: The inflow rate is labeled $\sigma(s)$
604: %
605: \begin{equation}\label{RateCoefInS2}
606: \sigma(s) = \sqrt{\frac{D}{s}}\frac{1}{\sinh\left(\ell\sqrt{s/D}\right)}.
607: \end{equation}
608: %
609: The inverse Laplace transforms of $\Delta(s)$, $\kappa(s)$ and
610: $\sigma(s)$ are shown in \eq{GeneralRateCoef} and depicted in
611: \fig{RateGraph}. The long time behavior of $\kappa(t)$ and $\sigma(t)$
612: can be found from a small $s$ expansion of the \eqs{RateCoefInS1} and
613: (\ref{RateCoefInS2})
614: %
615: \begin{equation} \begin{array}{lll}\label{RateLimit2}
616: \kappa(t) &\approx & \frac{D}{\ell} - \sqrt{\frac{D}{\pi t}}\\
617: \sigma(t) &\approx & \frac{D}{\ell}
618: \left[
619: 1 -\exp\left(-\frac{6Dt}{\ell^2}\right)
620: \right].
621: \end{array}\end{equation}
622: %
623: Taking the limit $t \rightarrow \infty$ yields
624: \begin{equation}\begin{array}{lr}\label{RateLimit}
625: \;\;\sigma(\infty) = \frac{D}{\ell}\;\;
626: &\;\;\kappa(\infty) = \frac{D}{\ell}.
627: \end{array}\end{equation}
628: %
629:
630: It is more convenient to express the rate \eqs{RateInS} in time domain
631: %
632: {\small
633: \begin{subequations}
634: \label{RateEqTwoCont}
635: \begin{eqnarray}
636: \dot{N}_1(t) & = & \frac{V_{d-1}(a)}{V_d(R_2)} \int_0^t dt'
637: {\cal N}_2(t')\sigma(t-t')\nonumber\\
638: & & -\frac{V_{d-1}(a)}{V_d(R_1)}\int_0^t dt'
639: {\cal N}_1(t')\left[\Delta(t-t')
640: + \kappa(t-t')\right]\nonumber\\
641: & &\\
642: \dot{N}_2(t) & = & \frac{V_{d-1}(a)}{V_d(R_1)} \int_0^t
643: dt'{\cal N}_1(t')\sigma(t-t')\nonumber\\
644: & & -\frac{V_{d-1}(a)}{V_d(R_2)}\int_0^t dt'
645: {\cal N}_2(t')\left[\Delta(t-t')+\kappa(t-t')\right]
646: \nonumber\\
647: & &
648: \end{eqnarray}
649: \end{subequations}
650: }
651: %
652: $\kappa(t)$ and $\sigma(t)$ are not present in the rate equation for
653: the semi-infinite case (see Eq. \ref{AlmostRate}) and arise only when
654: the system is finite.
655:
656: A numerical solution to \eqs{RateEqTwoCont} is shown in
657: \fig{NumSolTwoCont} (see \app{Numerical} for a more elaborate
658: discussion regarding the numerical procedure). The number of particles
659: decays exponentially which is verified in \fig{LogTwoCont} where the
660: straight line attained after some time is the evidence of a single
661: exponential decay. The figure also shows a non-exponential regime for
662: small $t$, described by $\Delta(t)$. Terms proportional to $\Delta(t)$
663: are in the following referred to as $\Delta$-terms.
664:
665: For small $t$ particles rush into the tube with a large current. At
666: $t=0$ the current is infinite, $\lim_{t\rightarrow 0}
667: \dot{N_1}(t)=\infty$. Thus, exactly at $t=0$ it is impossible to
668: define the exit rate from container and such situation extends to any
669: other time instant. In \emph{strict} mathematical sense, it is
670: impossible to define exit rate from the container for any $t>0$ (when
671: the concentration profile at the tube opening is different from
672: zero). This can be illustrated on a simple example. Let $V$ be a
673: volume divided into two sub-volumes $V_1$ and $V_2$ such that $V_1$
674: and $V_2$ touch each other and exchange particles by diffusion. The
675: dynamical variables of interest are the total number of particles in
676: each sub-volume $N_1(t)$ and $N_2(t)$. The goal is to derive some kind
677: of rate equation for $N_1(t)$ and $N_2(t)$. We focus on the flow from
678: $V_1$ to $V_2$. In a small time interval $\epsilon$ one will have
679: $N_1(t+\epsilon)\propto N_1(t)(1-\alpha\sqrt{\epsilon})$ and
680: $N_2(t+\epsilon)\propto N_2(t)+\beta N_1(t)\sqrt{\epsilon}$, where
681: $\alpha$ and $\beta$ are numerical constants. The effective exchange
682: rate that describes the flow from $V_1$ into $V_2$ is given by
683: %
684: \begin{equation}
685: k_{21}(t)=\lim_{\epsilon\rightarrow 0}
686: \frac{N_2(t+\epsilon)-N_2(t)}{N_1(t)\epsilon}
687: \propto\lim_{\epsilon\rightarrow 0} \epsilon^{-1/2}
688: \end{equation}
689: %
690: which is infinite. At any time instant $t$ an infinite amount of
691: particles (per unit time) is rushing from $V_1$ into $V_2$ (and vise
692: versa). However, the infinite flows from $V_1$ to $V_2$ and the other
693: way around cancel each other out resulting in a finite net flow giving
694: smooth curves for $N_1(t)$ and $N_2(t)$. $t=0$ is special since there
695: is no counter flow from $V_2$ to $V_1$ which explains why
696: $\lim_{t\rightarrow 0} \dot{N}_1(t) = \infty$.
697:
698: The non-exponential regime grows with tube length $\ell$ and might
699: play a significant role in studying transport processes in networks
700: having long connections. For large times $\kappa(t)$ and $\sigma(t)$
701: start to dominate and one observes exponential decay. It will be shown
702: in \sect{SingExpSec} how to derive rate equations that describe this
703: regime.
704:
705:
706:
707:
708:
709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710: \section{Analysis of the general rate equation: Emergence of
711: a single exponential solution} \label{SingExpSec}
712: It is intuitively clear that in the case of the two-node network
713: discussed in previous section one should have an exponential decay
714: (growth) for the number of particles in the container $C_1$ ($C_2$)
715: (see Figs. \ref{NumSolTwoCont} and \ref{LogTwoCont}): asymptotically
716: the time dependence of $N_1(t)$ and $N_2(t)$ is given by
717: %
718: \begin{equation}
719: N_{1,2}(t) = N_{1,2}(\infty) + {\cal A}_{1,2} \exp(-t/\tau)
720: \end{equation}
721: %
722: where $\tau^{-1}$ is the decay exponent that governs the late time
723: asymptotics and $ {\cal A}_{1,2}$ is the amplitude of decay. This
724: fact is not easily predicted from the form of the general rate
725: equation given in (\ref{GeneralEq}). To understand the emergence of
726: such behavior a more thorough investigation of \eq{TwoContRate} is
727: needed.
728:
729: To obtain the exact value of the decay exponent one has to study the
730: structure of poles of $N_{1,2}(s)$. The poles fully determine the
731: form of $N_{1,2}(t)=\sum_{p=0}^\infty a_p e^{s_pt}$ where $a_p$ is the
732: residue of $N_{1,2}(s)$ at pole $s_p$. The values of $N_{1,2}(\infty)$
733: are determined by the $p=0$ term ($s_0=0$). The exponential decay rate
734: is determined by
735: %
736: \begin{equation}\label{tau_s1}
737: \tau^{-1} = -s_1
738: \end{equation}
739: %
740: Rewriting \eqs{RateInS} in matrix form yields
741: %
742: \begin{equation}\label{DefofM}
743: s\vec{N}(s) - \vec{N}_0 = {\cal M}\vec{N}(s)
744: \end{equation}
745: %
746: where
747: %
748: \begin{equation}\begin{array}{lr}
749: \vec{N}(s) = [N_1(s),N_2(s)]^T, & \vec{N}_0 = [N_{10},N_{20}]^T
750: \end{array}\end{equation}
751: and
752: %
753: \begin{equation}\begin{array}{l}\label{DetM}
754: {\cal M} =\frac{qD}{\ell^2}
755: \left[
756: \begin{array}{cc}
757: -\frac{V_{\rm tube}}{V_d(R_1)} \coth q
758: & \frac{V_{\rm tube}}{V_d(R_2)}\frac{1}{\sinh q}\\\\
759:
760: \frac{V_{\rm tube}}{V_d(R_1)}\frac{1}{\sinh q}
761: & -\frac{V_{\rm tube}}{V_d(R_2)}\coth q
762: \end{array}
763: \right]
764: \end{array}\end{equation}
765: %
766: with $q^2 = s\ell^2/D$ and $V_{\rm tube} = V_{d-1}(a)\ell$. The poles
767: are calculated from $\det (s-{\cal M})=0$ since \mbox{$\vec{N}(s) =
768: (s-{\cal M})^{-1}\vec{N_0}$} and $(s-{\cal M})^{-1} \propto 1/\det
769: (s-{\cal M})$. Evaluating $\det (s-{\cal M}) = 0$ gives
770: %
771: \begin{equation}\label{TransEq}
772: q^2 + q\left[\frac{V_{\rm tube}}{V_d(R_1)} +
773: \frac{V_{\rm tube}}{V_d(R_2)}\right]\coth q +
774: \frac{V_{\rm tube}^2}{V_d(R_1)V_d(R_2)} = 0
775: \end{equation}
776: %
777:
778: Equation (\ref{TransEq}) is a transcendental equation and has many
779: solutions $q_p$ that determine the value of the poles $s_p = q_p^2
780: D/\ell^2$ where $p=1,2,\ldots,\infty$ and in particular
781: %
782: \begin{equation}\label{q_s1}
783: s_1= \frac{q_1^2D}{\ell^2}
784: \end{equation}
785: %
786: that, together with \eqs{tau_s1}, gives a relationship between $q_1$
787: and the decay rate $\tau^{-1}=-q_1^2D/\ell^2$. Note that $q^2$ has
788: been factored out from \eq{TransEq} and $s_0=0$ ($q_0=0$) is the
789: additional pole that determines the values of $N_{1,2}(\infty)$. Also,
790: note that $q_1$ depends only on parameters describing geometry of the
791: network.
792:
793: Apart from determining the structure of the poles,
794: \eqs{DefofM}-(\ref{DetM}) are a good starting point for classifying
795: various schemes for obtaining approximative forms of the rate
796: \eqs{RateEqTwoCont}. Equations (\ref{RateEqTwoCont}) do not have the
797: form of the general rate law stated in \eq{GenRateEq} (due to the
798: presence of the $\Delta$-terms). Such a rate law would be easier to
799: understand intuitively. For example, the emergence of the single
800: exponential decay could be seen more easily in \eq{GenRateEq} than in
801: \eqs{RateEqTwoCont}. Also, an approximative form might be easier to
802: implement numerically, though at the cost of a lower accuracy at the
803: end. The idea is to perform a small $s$ expansion of \eq{DefofM}
804: based on a desired accuracy. In here we consider two cases.\\
805:
806: \noindent{\bf (a) Lowest order expansion:} Performing the expansion
807: %
808: \begin{equation}\begin{array}{ll}
809: q \coth q \approx 1 & \ \ \ \ \ \frac{q}{\sinh q} \approx 1
810: \end{array}\end{equation}
811: %
812: of ${\cal M}$ in \eq{DetM} leads to a matrix that is constant, and taking the
813: inverse Laplace transform of \eq{DefofM} gives the following set of
814: rate equations
815: %
816: \begin{equation}\label{DagdugRateEqn}
817: \dot{N}_1(t) = -\dot{N}_2(t) =
818: V_{\rm tube}\frac{D}{\ell^2}\left[\frac{N_2(t)}{V_d(R_2)} -
819: \frac{N_1(t)}{V_d(R_1)} \right].
820: \end{equation}
821: %
822: These equations were already stated in ref.~\cite{Dagdug}. It is
823: interesting to see that they emerge as a special case of the scheme
824: discussed here. Also, \eq{DagdugRateEqn} can be obtained by following
825: another route. Performing partial integration of \eqs{RateEqTwoCont}
826: with terms containing $\Delta(t)$ omitted leads to the exactly the
827: same form of rate equations as given in (\ref{DagdugRateEqn}). This
828: procedure is discussed below.
829:
830:
831: The $\Delta$-terms are only present when the system is infinite. Since
832: the problem is finite, terms proportional to $\Delta(t)$ will be
833: sub-leading for large $t$. This can be seen from a small $s$ (large
834: $t$) expansion of \eqs{DefOfDelta} and ({\ref{RateCoefInS1})
835: \footnote{A small $s$ expansion of \eq{DefOfDelta} and
836: (\ref{RateCoefInS1}) in Laplace transform space yields ${\cal
837: L}^{-1}[\Delta(s)] \propto t^{-1/2}$ and ${\cal L}^{-1}[\kappa(s)]
838: \propto {\rm const}$. Since $\Delta(t)$ and $\kappa(t)$ always
839: combine in a sum, $\kappa(t)$ will dominate the outflow for large
840: $t$.}. Also, partial integration of the $\Delta$-terms is impossible
841: since the derivative of $\Delta(t)$ is proportional to $t^{-3/2}$ and
842: diverges when $t \rightarrow 0$.
843:
844: Omitting $\Delta$-terms in \eqs{RateEqTwoCont}, and performing partial
845: integration leads to
846: %
847: {\small
848: \begin{subequations}
849: \label{TwoContRate}
850: \begin{eqnarray}
851: \dot{N}_1(t) &=& \frac{V_{d-1}(a)}{V_d(R_2)}
852: \int_0^t dt' N_2(t-t')
853: \dot{\sigma}(t')\nonumber\\
854: & & -\frac{V_{d-1}(a)}{V_d(R_1)}
855: \int_0^t dt'
856: N_1(t-t')\dot{\kappa}(t')\\
857: \dot{N}_2(t) &=& \frac{V_{d-1}(a)}{V_d(R_1)}\int_0^t dt'
858: N_1(t-t')\dot{\sigma}(t')\nonumber\\
859: & & -\frac{V_{d-1}(a)}{V_d(R_2)} \int_0^t dt'
860: N_2(t-t')\dot{\kappa}(t')
861: \end{eqnarray}
862: \end{subequations}
863: }
864: %
865: Since $\dot{\sigma}(t)$ is peaked for small $t$ (see \fig{RateGraph}),
866: the contribution to the integrals (convolutions) stems mainly from
867: small values of $t'$ which justifies the following approximation
868: %
869: \begin{equation}\label{intapprox}
870: \int_0^t dt' N_{1,2}(t-t')\dot{\sigma}(t') \approx N_{1,2}(t) \int_0^t dt'
871: \dot{\sigma}(t')= N_{1,2}(t)\sigma(t).
872: \end{equation}
873: %
874: where $\sigma(0)=0$ was used. The same applies for $\kappa(t)$. Using
875: \eq{intapprox} in (\ref{TwoContRate}) leads to
876: %
877: {\small
878: \begin{subequations}
879: \label{FirstAttempt}
880: \begin{eqnarray}
881: \dot{N}_1(t) & = & \frac{V_{d-1}(a)}{V_d(R_2)}
882: N_2(t)\sigma(t) - \frac{V_{d-1}(a)}{V_d(R_1)}
883: N_1(t)\kappa(t) \ \ \ \ \ \ \ \\
884: \dot{N}_2(t) & = & \frac{V_{d-1}(a)}{V_d(R_1)}
885: N_1(t)\sigma(t) - \frac{V_{d-1}(a)}{V_d(R_2)}
886: N_2(t)\kappa(t). \ \ \ \ \ \ \
887: \end{eqnarray}
888: \end{subequations}
889: }
890: %
891: Inserting $\kappa(\infty)$ and $\sigma(\infty)$ found in
892: \eq{RateLimit} into (\ref{FirstAttempt}) yields
893: (\ref{DagdugRateEqn}). This example shows how the $\Delta$-terms
894: disappears from the description when the system is finite. However,
895: contrary to the partial integration method, the expansion of
896: \eq{DefofM} gives a more systematic and controlled approach.
897:
898: Equation (\ref{DagdugRateEqn}) is simple and computationally
899: efficient. It could be easily used to describe large
900: networks. However, it has several drawbacks that can be
901: identified. The solution to \eq{DagdugRateEqn} is given by
902: %
903: \begin{equation} \label{DagdugApprox}
904: N_{1,2}(t)-N_{1,2}(\infty) =
905: \left[N_{1,2}(0)-N_{1,2}(\infty)\right]
906: \exp\left(-t/\tau_{1,a}\right).
907: \end{equation}
908: %
909: The decay rate $\tau_a^{-1}= -q_{1,a}^2D/\ell^2$ is determined by
910: %
911: \begin{equation}\label{Rate1}
912: q_{1,a}^2 = -\frac{V_{\rm tube}[V_d(R_1) + V_d(R_2)]}{V_d(R_1)V_d(R_2)}.
913: \end{equation}
914: %
915: The number of particles in each container as $t\rightarrow \infty$ is
916: %
917: \begin{equation}\label{Ninfty}
918: \frac{N_1(\infty)}{V_d(R_1)} = \frac{N_2(\infty)}{V_d(R_2)}=
919: \frac{N_{10} + N_{20}}{V_d(R_1)+V_d(R_2)}.
920: \end{equation}
921: %
922: Equation (\ref{Ninfty}) is not correct. The correct values for
923: $N_{1,2}(\infty)$ are given by
924: %
925: \begin{equation} \label{LimitDist}
926: \frac{N_1(\infty)}{V_d(R_1)} = \frac{N_2(\infty)}{V_d(R_2)}=
927: \frac{N_{10} + N_{20}}{V_d(R_1)+V_d(R_2)+V_{\rm tube}}
928: \end{equation}
929: %
930: The discrepancies between \eq{Ninfty} and (\ref{LimitDist}) become
931: increasingly important for long tubes which are likely to occur in
932: large networks. For example, in a case where the tube and reservoir
933: volumes are equal, \eq{Ninfty} predicts $N_1(\infty) =
934: N_2(\infty)=N_{tot}/2$, $N_{tot}=N_{10} + N_{20}$, while the exact
935: result from \eq{LimitDist} is $N_{tot}/3$. The particle decay exponent
936: given in \eq{Rate1} only holds when $V_{\rm tube} \rightarrow 0$. It
937: strongly deviates from the exact value when the tube is long (see
938: \fig{TransEqFig}). \\
939:
940:
941: \noindent{\bf (b) Higher order expansion:} Using the expansion
942: %
943: \begin{equation}\begin{array}{ll}\label{Ordo}
944: q \coth q \approx 1+\frac{q^2}{3} &\ \ \ \ \ \frac{q}{\sinh q} \approx 1
945: \end{array}\end{equation}
946: %
947: for ${\cal M}$, inserting in \eq{DefofM} and taking inverse Laplace
948: transform, leads to a set of rate equations (given in
949: \app{appFirstTauM2}) that are unsatisfactory due to the following
950: reasons. First, they predict a spurious jump in $N_{1,2}(t)$ as $t
951: \rightarrow 0$, and the limiting values $N_{1,2}(\infty)$ are not
952: correct. Second, the rate exponent that results from these equations
953: is not that accurate. This particular example shows that it is
954: important to have a balanced expansion for elements of ${\cal M}$. For
955: example, instead of expanding $q\coth q$ directly one has to expand
956: $\sinh q$ and $\cosh q$ separately in such a way that same powers in
957: the nominator and denominator are obtained. When this strategy is
958: followed a much better approximation is obtained as shown bellow.
959:
960: The next order expansion, gives correct limits for $N_{1,2}(t)$ when
961: $t\rightarrow 0$ and $t\rightarrow\infty$ and leads to a relatively
962: accurate value for the decay exponent. Using the expansion
963: %
964: \begin{equation}\begin{array}{ll}
965: q\coth q \approx \frac{1+q^2/2}{1+q^2/6}\ \ \ \ \
966: \frac{q}{\sinh q} \approx \frac{1}{1 +q^2/6}
967: \end{array}\end{equation}
968: %
969: in ${\cal M}$ gives the following set of equations
970: %
971: {\small
972: \begin{subequations}
973: \label{ThirdAttempt}
974: \begin{eqnarray}
975: \dot{N}_1(t) &=& -N_1(t)\frac{3D}{\ell^2}\frac{V_{\rm tube}}{V_d(R_1)}\nonumber \\
976: & & +\frac{12D^2}{\ell^4}\frac{V_{\rm tube}}{V_d(R_1)}
977: \int_0^t dt' N_1(t-t')\exp\left(-\frac{6Dt'}{\ell^2}\right) \nonumber\\
978: & & -\frac{6D^2}{\ell^4}\frac{V_{\rm tube}}{V_d(R_2)}\nonumber
979: \int_0^t dt' N_2(t-t')\exp\left(-\frac{6Dt'}{\ell^2}\right)\nonumber\\
980: & &\\
981: \dot{N}_2(t) &=& -N_2(t)\frac{3D}{\ell^2}\frac{V_{\rm tube}}{V_d(R_2)}\nonumber\\
982: & & \frac{12D^2}{\ell^4}\frac{V_{\rm tube}}{V_d(R_2)}\nonumber
983: \int_0^t dt' N_2(t-t')\exp\left(-\frac{6Dt'}{\ell^2}\right)\nonumber\\
984: & & -\frac{6D^2}{\ell^4}\frac{V_{\rm tube}}{V_d(R_1)}
985: \int_0^t dt' N_1(t-t')\exp\left(-\frac{6Dt'}{\ell^2}\right)\nonumber\\
986: & &
987: \end{eqnarray}
988: \end{subequations}
989: }
990: %
991: Solving $\det (s-{\cal M}) = 0$ to get hold of the decay exponent in
992: analytical form becomes in this case rather tedious since finding the
993: value $q_1$ amounts to finding a root of a fourth degree
994: polynomial. The calculation simplifies somewhat if equal container
995: volumes are considered, $V_d(R_1)=V_d(R_2)\equiv V_d$. In such a case
996: one has
997: %
998: \begin{equation}\begin{array}{lll}\label{Rate3}
999: q_{1,b}^2 &=&-\frac{1}{2V_d}\left[3(2V_d+V_{\rm tube})\right.\\\\
1000:
1001: & &-\left.\sqrt{3(12V_d^2-4V_dV_{\rm tube} + 3V^2_{\rm tube})}\right]
1002: \end{array}\end{equation}
1003: %
1004:
1005: The main findings of this section are summarized in
1006: Figs. \ref{NumSolTwoCont}, \ref{LogTwoCont} and
1007: \ref{TransEqFig}. Figures \ref{NumSolTwoCont} and \ref{LogTwoCont}
1008: depict a numerical solution of \eq{RateEqTwoCont} (solid line)
1009: compared with the approximations discussed in this section for a case
1010: where the tube and container volumes are equal. Figure
1011: \ref{TransEqFig} shows a detailed analysis of the decay rate.
1012:
1013: For very short times there is a difference between \eqs{ThirdAttempt}
1014: and (\ref{RateEqTwoCont}) in \fig{NumSolTwoCont} and
1015: \ref{LogTwoCont}. These arise due to the partial elimination of
1016: $\Delta$-terms. For example, \eq{ThirdAttempt} does not predict
1017: $\lim_{t\rightarrow 0} \dot{N}_1(t)=\infty$ [\fig{NumSolTwoCont}, the
1018: dashed line lies above the solid line near $t=0$ for curves depicting
1019: $N_1(t)$]. This is the reason why the curve for $N_1(t)$ obtained from
1020: \eq{ThirdAttempt} underestimates the emptying of container $C_1$. This
1021: effect is more pronounced for the $N_1(t)$ coming from
1022: \eq{DagdugApprox}. There, the $\Delta$-terms are eliminated altogether
1023: [\fig{NumSolTwoCont}, the dotted line depicting $N_1(t)$ lies above
1024: solid and dashed lines]. Also, Figure \ref{NumSolTwoCont} shows that
1025: there is a large error in $N_{1,2}(\infty)$ for curves obtained by
1026: \eqs{DagdugRateEqn}.
1027:
1028: In \fig{LogTwoCont} the natural logarithm of $[N_1(t) -
1029: N_1(\infty)]/N_{tot}$ is shown. For short times the dynamics is not
1030: exponential but after some time a straight line is attained which is
1031: the evidence of single exponential behavior. The curve corresponding
1032: to \eq{DagdugApprox} (dotted line) does not predict the correct decay
1033: exponent which is manifested in a different slope. The decay exponent
1034: predicted by \eq{ThirdAttempt} is a better estimate for the decay
1035: rate: the slopes of the dashed and solid lines more or less
1036: coincide. This fact is shown more clearly in \fig{TransEqFig}.
1037:
1038: Figure \ref{TransEqFig} depicts the dependence of $q_1^2$ as a
1039: function of the tube volume. The numerical solution to \eq{TransEq}
1040: (solid line), which gives the exact value of the decay exponent, is
1041: compared to the values of $q^2_{1,a}$ (dotted line) and $q^2_{1,b}$
1042: (dashed line). All three cases work well when $V_{tube}\rightarrow
1043: 0$. As the tube volume increases $q_{1,a}^2$ deviates more and more
1044: from $q_1^2$. The same holds for $q_{1,b}^2$, though its value lies
1045: much closer to $q_1^2$. For example, when all volumes are equal $q_1
1046: =-1.71$ and $q^2_{1,b} = -1.63$ while $q^2_{1,a} = -2$.
1047:
1048: In this section methods of finding rate equations and decay exponents
1049: for the two container problem was deduced. Figures
1050: \ref{NumSolTwoCont}, \ref{LogTwoCont} and \ref{TransEqFig} show that
1051: for increasing tube volumes the rate equations given in
1052: \eq{DagdugApprox} are not capable of describing the dynamics. The
1053: approximation given in \eq{ThirdAttempt} works better. For short time
1054: dynamics none of the developed methods are valid and the full rate
1055: \eq{GeneralEq} is the only alternative. In the subsequent section all
1056: rate equations discussed up to this point will be extended to work for
1057: any network structure.
1058:
1059:
1060: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1061: \section{The General Expression}\label{SecGeneralEqs}
1062: In this section results and methods obtained and developed for
1063: two-node network will be extended to work for any network
1064: structure. The complete dynamics for the two-node network is
1065: formulated in \eq{RateEqTwoCont}. For an arbitrary network the outflow
1066: (OUT) from container $i$ to container $j$ is proportional to
1067: $\Delta_{ij}(t)+\kappa_{ij}(t)$ and the inflow (IN) from container $j$
1068: to container $i$ is proportional to $\sigma_{ji}(t)$:
1069: %
1070: {\small
1071: \begin{subequations}
1072: \label{GenEq}
1073: \begin{eqnarray}
1074: {\rm OUT}_{i\rightarrow j} & = & \frac{V_{d-1}(a_{ij})}{V_d(R_i)} \int_0^t dt'
1075: {\cal N}_i(t')\left[\Delta_{ij}(t-t')\right.
1076: \nonumber\\
1077: && \left.+ \kappa_{ij}(t-t')\right]\\
1078: {\rm IN}_{j\rightarrow i} &= & \frac{V_{d-1}(a_{ji})}{V_d(R_j)} \int_0^t dt'
1079: {\cal N}_j(t')\sigma_{ji}(t-t').
1080: \end{eqnarray}
1081: \end{subequations}
1082: }
1083: %
1084: This implies
1085: %
1086: \begin{equation} \label{GeneralSum}
1087: \dot{N}_i(t) = \sum_{j\ne i} {\cal C}_{ij}\left[{\rm IN}_{j\rightarrow i}
1088: - {\rm OUT}_{i \rightarrow j}\right],\;\;\;\; i = 1,\ldots, M
1089: \end{equation}
1090: %
1091: where ${\cal C}_{ij}$ is the conductivity matrix discussed in
1092: \sect{ProblemDef}. The final result is stated in \eq{GeneralEq}.
1093:
1094: It was argued earlier that if the tube volume is small a very simple
1095: first order rate equation can be stated. It is found in
1096: \eq{DagdugRateEqn}. Extending this equation to the case of arbitrary
1097: topology yields
1098: %
1099: \begin{equation}\label{LargeVApprox}
1100: \dot{N}_i(t) = \sum_{j\ne i} {\cal C}_{ij}V_{d-1}(a_{ij})\frac{D_{ij}}{\ell_{ij}}
1101: \left[\frac{N_j(t)}{V_d(R_j)}-\frac{N_i(t)}{V_d(R_i)}\right].
1102: \end{equation}
1103: %
1104: Note the symmetry relations $a_{ij}= a_{ji}$, $\ell_{ij}= \ell_{ji}$
1105: and $D_{ij}= D_{ji}$.
1106:
1107: If a more sensitive solution is desired, a set of equations of the
1108: type (\ref{ThirdAttempt}) is suggested. An extension of this equation
1109: is shown below
1110: %
1111: {\small
1112: \begin{subequations}
1113: \label{SensitiveRate}
1114: \begin{eqnarray}
1115: {\rm OUT}_{i\rightarrow j} &= & N_i(t)\frac{3D_{ij}}{\ell_{ij}}
1116: \frac{V_{d-1}(a_{ij})} {V_d(R_i)}\nonumber\\
1117: & &- \frac{12D_{ij}^2} {\ell_{ij}^3}\nonumber
1118: \frac{V_{d-1}(a_{ij})}{V_d(R_i)} \int_0^t dt'
1119: \exp\left(\frac{-6Dt'}{\ell_{ij}^2}\right)
1120: N_i(t-t')\\
1121: & &\\
1122: {\rm IN}_{j\rightarrow i} & = & \frac{6D_{ji}^2} {\ell_{ji}^3}
1123: \frac{V_{d-1}(a_{ji})}{V_d(R_j)}\int_0^t dt'
1124: \exp\left(\frac{-6D_{ji}t'}{\ell_{ji}^2}
1125: \right)N_j(t-t').\nonumber\\
1126: \end{eqnarray}\end{subequations}
1127: }
1128: %
1129: Inserting \eq{SensitiveRate} in \eq{GeneralSum} yields an
1130: approximative form of rate equations for an arbitrary network.
1131:
1132: In previous section we investigated differences between \eqs{GenEq},
1133: (\ref{LargeVApprox}) and \eq{SensitiveRate} using the two-node network
1134: as a study case ($M = 2$). It is expected that the findings of
1135: previous section also apply for larger networks. This analysis is not
1136: conducted here. Having general expressions at hand more complicated
1137: network structures can be investigated. However, the assumption of
1138: well stirred containers will be discussed first since it is crucial
1139: for the derivation of the rate equations~(\ref{GeneralEq}).
1140:
1141:
1142:
1143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1144: \section{The assumption of ideally mixed containers}\label{IdealMixing}
1145: An ideally mixed container has no concentration gradients. If a
1146: particle has entered, it can be found anywhere within the compartment
1147: with equal probability. In reality, a diffusing particle examines the
1148: compartment in a random walk fashion until an opening is found. There
1149: it has a possibility to escape and change the concentration. If the
1150: tube radius is small ($a/R\ll 1$), a significantly longer time is
1151: required to find an opening and escape than to examine the majority of
1152: the compartment. The time needed to examine the majority of the
1153: compartment is called mixing time and is given by $\tau_{{\rm mix}} =
1154: \frac{R^2}{D}$. The time of finding a specific place or target having
1155: radius $a$ is given by $\tau_{{\rm target}} = \frac{R^2}{D}
1156: \frac{R}{a}$ \cite{Stange}. For cubic compartments the radius of the
1157: sphere $R$ is replaced by the edge length, in this case $2R$ (see
1158: \fig{SolDiffEq}). Thus, we argue that if $a\ll R$ then $\tau_{{\rm
1159: mix}} \ll \tau_{{\rm escape}}$ and the containers can be considered
1160: ideally mixed at all times
1161: \footnote{$\tau_{\rm target}$ is derived under the assumption of a
1162: fully absorbing target of radius $a$. This is not a correct
1163: description of a tube opening since it allows reentry. This estimate
1164: serves however as a worst case scenario.}. This is supported by
1165: numerics.
1166:
1167: Figure \ref{SolDiffEq}~(b) shows a numerical solution of the diffusion
1168: equation in two dimensions in a geometry depicted in
1169: \fig{SolDiffEq}~(a). The solution clearly shows that the assumption of
1170: well stirred containers becomes very good for $a/R \ll 1$, even for
1171: skewed initial distributions. The solution to the diffusion equation
1172: was found using a standard implicit finite difference discretization
1173: method \cite{NumRes}.
1174:
1175: In \fig{SolDiffEq}~(b) one observes a systematic discrepancy: the
1176: assumption of ideally mixed containers tends to overestimate the decay
1177: rate. This derives from the fact that the assumption of well stirred
1178: containers over estimates the number of particles at the tube inlet
1179: leading to a larger exit rate.
1180:
1181:
1182:
1183: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1184: \section{Case studies}\label{CaseStudies}
1185: Up to this point the two node network was the only example discussed.
1186: It was used as elementary building block when constructing the rate
1187: equation (\ref{GeneralEq}). Such network is rather simple and does not
1188: offer any spectacular behavior. In this section more complicated
1189: structures will be studied that are shown in Figs. \ref{panel} and
1190: \ref{LargeNetwork}: Two realizations of a three-node network are
1191: studied first, case 1 and 2, depicted in \fig{panel}~(a) and (b). They
1192: possess one more level of complexity than the two-node network shown
1193: in \fig{TwoContSytem}. Case 3, \fig{panel}~(c), is a four-node network
1194: that has a T-shape structure. Case 4, \fig{panel}~(d), has the shape
1195: of a star and involves a tube junction. It will be shown that the
1196: transport properties of this structure can be controlled in such way
1197: that it might serve as a diffusion-based transistor. These three and
1198: four node-networks, despite being rather simple, exhibit a large
1199: variety of outputs. Case 5, \fig{LargeNetwork}, demonstrates the
1200: transport properties for larger networks that are impossible to
1201: predict without using \eq{GeneralEq}. When solving \eq{GeneralEq}
1202: numerically the containers and the tubes are considered three
1203: dimensional. Moreover, the diffusion coefficients $D_{ij}$ and the
1204: tube radii $a_{ij}$ were kept the same for all tubes. All graphs are
1205: scaled with the total number of particles in the system $ N_{tot} =
1206: \sum_{i = 1}^M N_i(0)$.\\
1207:
1208:
1209: %---------------
1210: % T H E L I N E
1211: %
1212: \noindent {\bf Case 1: The Line.} Three reservoirs are lined up on a
1213: straight line as depicted in \fig{panel}~(a). The transport equations
1214: for this system can easily be written down using \eq{GeneralEq} and
1215: are listed in \app{LineRate}. A numerical solution is shown in
1216: \fig{LineGraph}. This three-node system exhibits a new characteristic
1217: that can not be found in the two node network. The curve depicting the
1218: number of particles in the middle container (solid line) has a
1219: maximum (an extremum) point.
1220:
1221: The extremum point is a manifestation of an unbalance between inflow
1222: and outflow in the middle container. This unbalance derives from an
1223: asymmetry in the structure and arise only when $\ell_{12} < \ell_{23}$
1224: or $V_1<V_3$. This provides a possibility of designing the output
1225: pattern through simple geometrical changes in the structure. Fig.
1226: \ref{LineTiming} is a simple demonstration of this design possibility
1227: where $\ell_{12}$ is varied so that the arrival time of the maximum of
1228: $N_2(t)$, depicting the number of particles in the middle compartment
1229: $C_2$, is changed. The picture also shows that an increase in
1230: $\ell_{12}$ results in both an increased arrival time and a wider
1231: peak. The height of the maximum is controlled by the value $V_2$. A
1232: decrease in $V_2$ suppresses the peak and vice versa. \\
1233:
1234:
1235:
1236: %----------------------
1237: % T H E T R I A N G L E
1238: %
1239: \noindent {\bf Case 2: The Triangle.} The connectivity of the line is
1240: changed by adding an extra link between containers $C_1$ and $C_3$ so
1241: that it forms the shape of a triangle, as shown in
1242: \fig{panel}~(b). The rate equations are derived in \app{TriangleRate}
1243: and a numerical solution is shown in \fig{Triangle2}. The initial
1244: distribution of particles, see the inset in \fig{Triangle2}, is chosen
1245: in such a way that a minimum is produced in the curve depicting the
1246: number of particles in container $C_2$.
1247:
1248: The extremum point can be enhanced or removed completely in the same
1249: way as was demonstrated for case 1. Such minimum will be absent unless
1250: the geometry and initial distribution of particles are tailored in a
1251: specific way. In general, such sensitivity of geometrical changes and
1252: changes in the location where particles are injected is observed for
1253: all cases studied in this section.
1254:
1255: The triangle structure studied here exhibits shorter equilibrium
1256: process than the linear structure considered previously [see
1257: \fig{panel}~(a)]. In the case of a triangle structure particles can
1258: spread more efficiently due to the additional routing possibility
1259: $C_1\rightarrow C_3$.\\
1260:
1261:
1262:
1263:
1264: %-------------------------
1265: % T H E T - N E T W O R K
1266: %
1267: \noindent {\bf Case 3: The T-Network.} It is interesting to see how
1268: the behavior of the cases 1 and 2 changes when a new node is added to
1269: the network. In here we study a situation where an extra node $C_4$ is
1270: connected to container $C_2$ in the structure depicted in
1271: \fig{panel}~(a). In such a way one gets a T-shaped (star) network
1272: shown in \fig{panel}~(c). This alternation of structure leads to a
1273: significant change in behavior, as shown in \fig{LineGraphTop}. When
1274: compared to the cases 1 (one maximum) and 2 (one minimum) the curve
1275: depicting $N_2(t)$ exhibits an additional extremum point: both minimum
1276: and maximum are present simultaneously. The right inset in
1277: \fig{LineGraphTop} emphasizes this fact.
1278:
1279: This scheme could be carried out further adding on more and more
1280: reservoirs and adjusting the lengths and the initial distribution so
1281: that the peaks arrive in consecutive order, possibly produce an
1282: wave-like behavior. However, since the spread of the peaks increases
1283: with increasing tube length, it might be numerically quite difficult
1284: to see when the the extremum points occur or even if they actually
1285: exist.\\
1286:
1287:
1288:
1289: %-----------------------
1290: % T H E J U N C T I O N
1291: %
1292: \noindent {\bf Case 4: The Junction.} The next interesting network to
1293: consider is the one with a junction present as depicted in
1294: \fig{panel}~(d). A particular example of a three way junction is
1295: studied. The network is built up by three reservoirs and three
1296: tubes. The ends of the tubes coincide to form a junction. To obtain
1297: the transport properties of such a network we start from the structure
1298: studied in case 3 shown in \fig{panel}~(c). The junction is obtained
1299: by reducing the radius of container $C_2$ in the middle until its
1300: radius is roughly equal to the radii of the surrounding tubes, see
1301: \fig{Junction1}. The rate equations describing the junction properties
1302: have the same form as the equations describing case 3 (listed in the
1303: appendix \ref{Trate}) with the substitution $V_d(R_3)\rightarrow
1304: V_d(a)$. The equations are not given in order to save space.
1305:
1306: A numerical solution of the rate equations for the junction is shown
1307: in \fig{Junction}. This structure allows control of the particle flow
1308: between $V_1$ and $V_3$ by adjusting the volume $V_2$ and the length
1309: $\ell_{24}$ (see \fig{Junction1}). This setup could function as a
1310: diffusion based transistor. For example, by making $\ell_{24}$ shorter
1311: than $\ell_{34}$, compartment $C_2$ will initially attract diffusing
1312: particles from $C_1$ to a greater extent than $C_3$ causing a time
1313: delay in the particle arrival into container~$C_3$.
1314:
1315: This fact is illustrated in \fig{Junction} where the maximum in the
1316: curve for the number of particles in container $C_2$ (solid line)
1317: indicates an initial accumulation of particles in $C_2$:
1318: $N_2(t)/N_{tot}$ rises from 0 to 0.5 in the interval $t=0$ to
1319: $Dt/\ell^2=5$. In this interval container $C_2$ accumulates the
1320: majority of the particles released from container $C_1$. After the
1321: peak has been reached the particles accumulated in $C_2$ are released
1322: into $C_3$: $N_2(t)/N_{tot}$ continuously drops from value of 0.5
1323: after $Dt/\ell^2=5$. By changing $\ell_{24}$ the curve for $N_2(t)$
1324: can be manipulated exactly in the same way as done in
1325: \fig{LineTiming}, but such analysis is not repeated.
1326:
1327: For large networks involving junctions it might be desirable to
1328: decrease the number of equations required to solve the transport
1329: problem. It is demonstrated in \app{SectJunction} that the presence of
1330: the junctions can be eliminated altogether when investigating dynamics
1331: in the $t \rightarrow \infty$ regime for structures having large
1332: container volumes ($a\ll R_i$, $i=1,\ldots,M$). Equations
1333: (\ref{RateEqsJunction21})-(\ref{RateEqsJunction23}) show this
1334: explicitly for the three node junction studied here.
1335:
1336: The four cases studied up to now show that the transport dynamics is
1337: very sensitive to geometrical changes and to the locations where
1338: particles are injected. Small variations in tube lengths and container
1339: volumes lead to unpredictable changes in curves depicting the time
1340: dependence of the number of particles in each container
1341: (Figs. \ref{LineGraph}, \ref{Triangle2}, \ref{LineGraphTop} and
1342: \ref{Junction}). Also, the shapes of the curves differs significantly
1343: from a single exponential decay .\\
1344:
1345:
1346: %--------------------------------
1347: % L A R G E N E T W O R K
1348: %
1349: \noindent {\bf Case 5: Large Network.} Following the methods described
1350: in this paper one could easily study diffusive transport in networks
1351: containing hundreds or thousands of containers, tubes and
1352: junctions. The computational cost scales linearly with both the number
1353: of containers and number of tubes (assuming full connectivity of the
1354: containers). We do not show explicit example with such large number of
1355: containers. To demonstrate the power of the method we study a more
1356: pedagogical example, the case of a network that is built up by seven
1357: containers, 9 tubes and a 4-way junction. In contrast to the previous
1358: cases shown in \fig{panel} it is impossible to predict the transport
1359: behavior of such a network without a numerical calculation.
1360:
1361: Figure \ref{NetworkPanel}, panels~(a)-(c), shows numerical solutions
1362: of \eq{GeneralEq} for the network depicted in
1363: \fig{LargeNetwork}. Different initial distributions are used and are
1364: introduced into inset of all figures. The darker the container appears
1365: the more particles are injected into it. The dynamics is evidently
1366: quite complex and all possible characteristics that were forced upon
1367: the other cases are present. There is an exponential growth and decay
1368: as well as curves having one or more extremum points. The different
1369: transport behavior shown in \fig{NetworkPanel} stem only from
1370: different initial conditions. If the structure no longer remained
1371: fixed even more complicated patterns could be produced, only the
1372: imagination sets the limits.
1373:
1374:
1375:
1376: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1377: \section{Concluding remarks} \label{Conclusions}
1378:
1379: We introduced a generic model for a diffusive particle transport in
1380: large networks made of containers and tubes. The diffusion equation
1381: that describes the distribution of particles $\rho(\vec{r},t)$
1382: throughout the network was taken as a starting point. Instead of
1383: calculating $\rho(\vec{r},t)$ explicitly, we followed another route
1384: and developed a theoretical technique to solve the transport problem
1385: using finite number of variables $N_1(t),\ldots,N_M(t)$ that describe
1386: the number of particles in each container. First, a set of rate
1387: equations was derived for the two-node network and second, they were
1388: generalized to work for an arbitrary network structure. In such a way
1389: we obtained the rate equations that govern dynamics of
1390: $N_1(t),\ldots,N_M(t)$. These equations are summarized in
1391: \eq{GeneralEq} and are the central result of the paper.
1392:
1393: The transport equations were found by study the exchange of chemicals
1394: between the container and the tube. It was demonstrated in
1395: \sect{SecEmpt} how to couple the dynamics in the containers and tubes,
1396: as stated in Eqs. (\ref{Functional}) and
1397: (\ref{BC})-(\ref{PartConservation}). However, the coupling is too
1398: complicated to be carried out in practice in the original
1399: form. Several approximations were made in order to make such scheme
1400: doable.
1401:
1402: The tubes were assumed to be one dimensional lines (see Eq. \ref{BC}),
1403: and the transport in the tubes was described in terms of a one
1404: dimensional diffusion problem involving the concentration profile
1405: along the line $c(x,t)$. The expression for $c(x,t)$ can be found
1406: analytically using e.g. Laplace transform technique. In such a way the
1407: tubes were eliminated from the problem.
1408:
1409: We have considered diffusive non-interacting point particles which is
1410: a plausible assumption for dilute solutions. One could consider the
1411: situation when the particles disturb each other. In that respect, the
1412: region of the tube interior is the most critical since the tubes can
1413: be very narrow and exclusion effects will be mostly pronounced in
1414: there. Such effects could be added into the theoretical description by
1415: using results obtained from studies on diffusion with exclusion in one
1416: dimensional systems~\cite{BD,JSI,EFCm}. We expect a very different
1417: transport behavior when the diameters of the tubes becomes comparable
1418: in size to that of the particles.
1419:
1420: The dynamics in the container is not tractable analytically but it was
1421: argued that when the tube radii are smaller than any other length
1422: scales in the system (e.g. tube lengths or container radii) the
1423: containers can be treated as ideally mixed at all times. This
1424: assumption was verified numerically in \sect{IdealMixing}, and
1425: simplifies all intra-container dynamics to one dynamical variable: the
1426: total number of particles in a given compartment.
1427:
1428: Evidently much of the container dynamics has been neglected but the
1429: coupling is formulated in such a way that a more detailed description
1430: can be developed should there be a need for that. For example, the
1431: container dynamics could be better treated by using the techniques
1432: presented in e.g. \cite{Dagdug}, or by further exploration of the
1433: coupling equations (\ref{Functional}) and
1434: (\ref{BC})-(\ref{PartConservation}). For example, one could keep the
1435: second term in the right hand side of \eq{LinearInterpol} and use
1436: $\rho(0,y,z,t) = N(t)/V_d(R)-J(t)/4D_ca$ instead of $\rho(0,y,z,t) =
1437: N(t)/V_d(R)$. This procedure would lead to similar rate equations as
1438: presented here with different forms for $\Delta(t)$, $\kappa(t)$ and
1439: $\sigma(t)$.
1440:
1441: Initially the particles feel as if they are escaping from the
1442: container into an infinitely long tube. In this regime the number of
1443: particles in the container decays non-exponentially. We have
1444: identified terms in the rate equations describing this behavior: all
1445: terms in \eq{GeneralEq} proportional to $\Delta(t)$ dominate when time
1446: is small. This non-exponential regime grows with increasing tube
1447: length and crossover time for this regime scales as $\ell^2/D$. Also,
1448: terms containing $\Delta(t)$ contain solely dependence on
1449: $\dot{N}_i(t)$ and can not be rewritten in the form that would involve
1450: $N_i(t)$. Accordingly, it it impossible to rewrite \eq{GeneralEq} so
1451: that it adopts a form of a general rate law (see
1452: Eq. \ref{GenRateEq}). These issues were discussed in \sect{SecEmpt}.
1453: The bottle neck lies in the definition of transport rate which, in
1454: principle, is an ill-defined quantity (see discussions at the end of
1455: sections \ref{SecEmpt} and \ref{SecRate}).
1456:
1457: Eventually, for large times the decay is exponential. In such regime
1458: term proportional to $\Delta(t)$ can be neglected in the rate
1459: \eq{GeneralEq}. Other terms proportional to $\kappa(t)$ and
1460: $\sigma(t)$ can be rewritten in the form of a general rate law,
1461: leading to \eqs{LargeVApprox} and (\ref{SensitiveRate}). We showed
1462: that \eq{LargeVApprox} is a special case of the general approximation
1463: scheme developed in \sect{SingExpSec}: starting from rate
1464: \eq{GeneralEq} in Laplace transform space we developed a series of
1465: approximations that can be used to systematically describe the
1466: asymptotic regime, resulting in \eq{SensitiveRate}. Also, we have
1467: developed a procedure that can be used to eliminate junction points
1468: for large times which reduces the number of variables further (see
1469: \app{SectJunction}).
1470:
1471: Already simple case studies that were used to illustrate the workings
1472: of the method exhibit interesting behavior. For example, one can
1473: identify three types of curves that appear in the plots depicting the
1474: time dependence of the particle number in each container
1475: (Figs. \ref{LineGraph} - \ref{LineGraphTop} and \ref{Junction}).
1476: Type~I curves occur for the two node network. These lack extremum
1477: points and the particle number either strictly rises or drops to
1478: saturate to asymptotic values. Type II curves have one maximum or
1479: minimum, and type III curves can have more and these are the most
1480: interesting. Type II and type III curves normally describe the
1481: particle number for the container in the network interior.
1482:
1483: The existence of type I curves for large networks suggests that it
1484: could be possible to understand the transport between two nodes in
1485: terms of an effective two node network where a complicated structure
1486: of links and containers in between two nodes is mapped onto an
1487: effective link connecting them. One can raise a more general
1488: question: what is the smallest network that would have the same
1489: transport properties as some sub-structure of a given large network?
1490: For example, if one is interested in only three nodes of the structure
1491: depicted in \fig{LargeNetwork}, e.g. $C_5$, $C_7$ and $C_8$, is there
1492: a star-type network, e.g. as the one depicted in \fig{panel} (c) or
1493: (d), that would have equivalent transport properties?
1494:
1495: The presence of type II and type III curves indicates the possibility
1496: that there might be curves that posses a larger number of extremum
1497: points. These are likely to occur in larger networks. We can of
1498: course manually enhance certain properties as height and width of
1499: peaks. This will however become more and more complicated as the
1500: network size increases. Problem is that the peaks that occur later
1501: have larger width and might be harder to see. For design purposes, it
1502: is therefore necessary to build a learning mechanism or a search
1503: engine, on top of our existing software, to select certain
1504: characteristics in the curves depicting time evolution of
1505: $N_1,...,N_M$. Furthermore, to exploit such effects one has to
1506: amplify them in some way. At the moment our study deals with transport
1507: only, but reactions in the containers can be included as well, and
1508: they could be tailored to amplify such effects (e.g. by choosing a
1509: reaction of enzymatic type).
1510:
1511: The techniques developed in this work can be used to study transport in
1512: various systems. In the following we give a few examples.
1513:
1514: (i) Although we exclusively study transport, our model can serve as a
1515: platform for reaction-diffusion-based bio-computing devices
1516: ~\cite{SSFA, KK, Rambidi, CZ1, CZ2, CZ3, Ji, SM,LPHR, HR,
1517: HSR,Akingbehin, AkinCon, Conrad, KirCon1, KirCon2, KamCon1, KamCon2}.
1518: In particular, our work is applicable to studies of the
1519: reaction-diffusion neuron ~\cite{Akingbehin, AkinCon, Conrad, KirCon1,
1520: KirCon2, KamCon1, KamCon2}. The reaction-diffusion neuron is an 2D
1521: array of compartments that exchange chemicals by diffusion.
1522:
1523: (ii) A large number of processes happening in the
1524: cell are governed by transport of
1525: reactants and chemical reactions. In order to avoid a need for
1526: excessive storage facilities the chemical compounds are routed in an
1527: orderly fashion between various places within and between the cells
1528: and the chemical components arrive exactly at the right place at the
1529: right time \cite{Stange,HM}. The setup in \fig{MainNetwork} captures
1530: this aspect of the cell interior.
1531:
1532: (iii) The transport on abstract mathematical networks (nodes and
1533: links) has been studied extensively~\cite{Watts,WattsBook}. Such
1534: studies are geometry free with emphasis on the topology of the network
1535: graph (the connectivity patter, the average number of neighbors
1536: etc.). The techniques developed in this work could be used to account
1537: for the fact that the links between the nodes have physical length and
1538: the transport along the links is not instantaneous.
1539:
1540: In summary, the work we have presented is a step towards understanding
1541: the transport properties of large networks where geometrical concepts
1542: such as the length of the tubes play an important role. The setup
1543: employed in this study is rather simple. In order to be able to focus
1544: on issues related to transport, the reactions are totally omitted. The
1545: concentration profile in containers is assumed flat and this is good
1546: approximation when tubes are thin. Already simple examples of
1547: networks we studied show a number of interesting properties. For
1548: example, transport properties of the networks exhibit large
1549: sensitivity to the geometrical changes in the structure. Also, one can
1550: adjust structure to obtain wave-like behavior (with one or two
1551: extremum points) in the curves that depict number of particles in
1552: containers. When the complexity of the network increases one can
1553: expect even more complicated behavior with larger number of extrema
1554: points. The setup we use is generic and it is possible to expand the
1555: model in many ways, e.g. by improving description of intra-container
1556: dynamics, incorporating reactions, or allowing particles to disturb
1557: each other. It will be an interesting problem to try to explore these
1558: questions further.
1559:
1560:
1561:
1562:
1563:
1564: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1565: \begin{acknowledgments}
1566:
1567: We would like to thank Prof. Owe Orwar and Prof. Mats Jonson for
1568: fruitful discussions. The financial support of Prof. Owe Orwar is
1569: greatly acknowledged.
1570:
1571: \end{acknowledgments}
1572:
1573:
1574:
1575:
1576: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1577: % A P P E N D I X
1578:
1579: \appendix
1580:
1581: \section{Numerical considerations}\label{Numerical}
1582: In this section the numerical solution of \eq{GeneralEq} is
1583: discussed. In \sect{SecEmpt} and \sect{SecRate} it was shown that the
1584: time derivatives $\dot{N}_i(t),\ldots,\dot{N}_M(t)$ can be infinite at
1585: $t=0$ which may cause numerical difficulties. However, the singular
1586: part of the derivative can be factored out by making the substitution
1587: %
1588: \begin{equation}\label{Substitution}
1589: \dot{N}_i(t)=-\left(\pi t\right)^{-1/2}\psi_i(t),\;\;\;\;\;i=1,\ldots,M
1590: \end{equation}
1591: %
1592: where $\psi_i(t)$ is a smooth function of time. For small $t$ the
1593: particles in the containers do not (yet) 'feel' the presence of
1594: another side and behave as if entering an infinitely long tube. An
1595: expression describing the transport behavior for for such a case was
1596: found analytically and is stated in \eq{SolInfInT}. The time
1597: derivative of \eq{SolInfInT} is proportional to $t^{-1/2}$ for small
1598: $t$. Inserting \eq{Substitution} in \eq{GeneralEq} yields an equation
1599: for $\psi_i(t)$ that was used for numerical calculations
1600: %
1601: \begin{equation}\label{GeneralEq2}\begin{array}{lll}
1602: \psi_i(t)&=& \sum_{j=1}^M {\cal C}_{ji}\frac{V_{d-1}(a_{ji})}{V_d(R_j)}
1603: \left[\sqrt{\frac{t}{\pi}} \int_0^t \frac{dt'}{\sqrt{t'}}
1604: \psi_j(t')\sigma_{ji}(t-t')\right.\\\\
1605: & & \left.+\sqrt{\pi t} \,N_{j}(0)\sigma_{ji}(t)\right]\\\\
1606:
1607: & & -\sum_{j=1}^M {\cal C}_{ij}\frac{V_{d-1}(a_{ij})}{V_d(R_i)}
1608: \left[\sqrt{\frac{t}{\pi}}\int_0^t\frac{dt'}{\sqrt{t'}}
1609: \psi_i(t')\kappa_{ij}(t-t') \right. \\\\
1610: & & +\frac{\sqrt{D_{ij}t}}{\pi}\int_0^t\frac{dt'}
1611: {\sqrt{(t-t')}}\psi_i(t')\\\\
1612: & & \left. +\sqrt{\pi t}\,N_i(0) \left(\Delta_{ij}(t) +
1613: \kappa_{ij}(t)\right)\right].
1614: \end{array}\end{equation}
1615: %
1616: Note that the expression for $\Delta_{ij}(t)$ has been inserted into
1617: the integrals. Combining \eq{SolInfInT} and (\ref{Substitution}) leads
1618: to
1619: %
1620: \begin{equation}
1621: \psi_i(t) = N_{i0}-(\pi t)^{1/2}N_i(t),\;\;\;\;\;i=1,\ldots,M.
1622: \end{equation}
1623: %
1624: and sets the initial condition $\psi_i(0)=N_{i0}$.
1625:
1626: From \eq{GeneralEq2} two types of integrals can be identified:
1627: %
1628: \begin{equation}\begin{array}{lll}
1629: I_1^{t_n}\left[\psi\right] &=
1630: &\int_0^{t_n}\frac{dt}{\sqrt{t'(t_n-t')}}\, \psi(t') \\\\
1631: I_2^{t_n}\left[\psi\right] &=
1632: &\int_0^{t_n}\frac{dt}{\sqrt{t'}}\, \psi(t')
1633: \end{array}\end{equation}
1634: %
1635: where $\psi(t)$ is non-singular in the range of integration. The
1636: quadrature formulas derived to solve $I_1^{t_n}\left[\psi\right]$ and
1637: $I_2^{t_n}\left[\psi\right]$ are based on the methods described in
1638: \cite{Mohamed}. The idea is that the singular part of the integrand,
1639: $t'^{-1/2}$ and $[t'(t_n-t')]^{-1/2}$ respectively, is treated
1640: exactly while the smooth part $\psi_i(t')$ is linearly interpolated
1641: between $t_i$ and $t_{i+1}$. The resulting quadrature formulas are of
1642: the form
1643: \begin{equation}
1644: I^{t_n}_i\left[\psi\right] \approx \sum_{j=0}^n w_{nj} \,\psi(t_j)
1645: \;\; i=1,2
1646: \end{equation}
1647: where $w_{nj}$ are weights, $n = 1,2\ldots$ and $t_n = nh$. This
1648: quadrature formula becomes exact when $\psi(t)$ is piecewise
1649: linear. Calculating weights for $I_1^n[\psi]$ yields
1650: %
1651: {\small
1652: \begin{equation}\begin{array}{lll}
1653: w_{n0}&= &\sqrt{n-1} - (n-2) \arcsin \sqrt{\frac{1}{n}}\\\\
1654: w_{ij}&= &\sqrt{(j-1)(n-j+1)}+\sqrt{(j+1)(n-j-1)}\\\\
1655: & &-2\sqrt{j(n-j)} + 2\left(j+1-\frac{n}{2}\right)\left[\arcsin\sqrt{\frac{j-1}{n}}\right.\\\\
1656: & &-2\left.\arcsin\sqrt{\frac{j}{n}} +\arcsin\sqrt{\frac{j+1}{n}}\right]\\\\
1657: w_{nn}&=&\sqrt{n-1}+\pi\left(1-\frac{n}{2}\right)+(n-2)\arctan\sqrt{n-1}.
1658: \end{array}\end{equation}
1659: }
1660: %
1661: Calculating weights for $I_2^n[\psi]$ yields
1662: %
1663: \begin{equation}\begin{array}{ll}
1664: w_{n0} &=\frac{4h^{1/2}}{3} \\\\
1665: w_{nj} &=\frac{4h^{1/2}}{3}\left[(j+1)^{3/2}-2j^{3/2}+(j-1)^{3/2}\right]\\\\
1666: w_{nn} &=\frac{2h^{1/2}}{3}\left[n^{3/2}-3(n-1)^{1/2}+2(n-1)^{3/2}\right].
1667: \end{array}\end{equation}
1668: %
1669:
1670: Finally, an expression for the total number of particles is found by
1671: integrating \eq{Substitution}
1672: %
1673: \begin{equation}
1674: N_i(t) = N_i(0) -\int_0^t \frac{dt'}{\sqrt{t'}}\, \psi_i(t'),\;\;\;\;\;i=1,\ldots,M.
1675: \end{equation}
1676: %
1677: $N_i(t),\ldots,N_M(t)$ are found by using the quadrature formula
1678: derived for $I_2^n[\psi]$.
1679:
1680:
1681:
1682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1683: \section{Finding a set of approximative rate equations from an
1684: expansion of \eq{DefofM} in the variable $s$.} \label{appFirstTauM2}
1685: In this section, details for obtaining a set of a first order rate
1686: equations from \eq{DefofM} will be outlined. The dynamical equations
1687: are found by approximating ${\cal M}$ given in \eq{DetM}. ${\cal M}$ is
1688: approximated by using the expansion stated in \eq{Ordo}. The inverse
1689: Laplace transform of \eq{DefofM} after approximation reads
1690: %
1691: {\small
1692: \begin{subequations}
1693: \label{SecondAttempt}
1694: \begin{eqnarray}
1695: \dot{N}_1(t) &= & \frac{3V_d(R_1)V_{\rm tube}}{3V_d(R_1)
1696: +V_{\rm tube}}\frac{D}{\ell^2}\left[
1697: \frac{N_2(t)}{V_d(R_2)}-\frac{N_1(t)}{V_d(R_1)}\right]\nonumber\\
1698: & & -\delta(t)\frac{V_{\rm tube}}{3V_d(R_1)
1699: +V_{\rm tube}}N_{10}\\
1700: \dot{N}_2(t) &= & \frac{3V_d(R_2)V_{\rm tube}}{3V_d(R_2)+
1701: V_{\rm tube}}\frac{D}{\ell^2}
1702: \left[\frac{N_1(t)}{V_d(R_1)}
1703: -\frac{N_2(t)}{V_d(R_2)}\right]\nonumber\\
1704: & & -\delta(t)\frac{V_{\rm tube}}{3V_d(R_2)
1705: +V_{\rm tube}}N_{20}.
1706: \end{eqnarray}
1707: \end{subequations}
1708: }
1709: %
1710: The decay rate predicted by these equations is given by $\tau_{1,b'}^{-1}=
1711: -\ell^2/q^2_{1,b'}D$ where
1712: %
1713: \begin{equation}\label{Rate2}
1714: q^2_{1,b'} = -\frac{ V_{\rm tube}[V_d(R_1) + V_d(R_2)] }
1715: { V_d(R_2)[V_d(R_1) + V_{\rm tube}/3] }
1716: \end{equation}
1717: %
1718: This decay exponent is not adequate. For example, in the case where
1719: all volumes are equal $q^2_{1,b'} = -1.5$ while the exact value is
1720: $q_{\rm exact} = -1.71$.
1721:
1722: The rate Eq.~(\ref{SecondAttempt}) can not describe the behavior of
1723: $N_{1,2}(t)$ as $t\rightarrow \infty$ and $t\rightarrow 0$: the
1724: values for $N_{1,2}(\infty)$ are given by
1725: %
1726: \begin{equation}
1727: \frac{N_1(\infty)}{V_d(R_1)} = \frac{N_2(\infty)}{V_d(R_1)} =
1728: \frac{N_{10} + N_{20}}{V_d(R_1)+V_d(R_2)
1729: + \frac{2}{3}V_{\rm tube}}
1730: \end{equation}
1731: %
1732: and for $N_{1,2}(0^+)$ there is a sudden jump
1733: %
1734: \begin{equation}
1735: N_i(0^+) = \frac{N_{i0}}{1+V_{\rm tube}/3V_d(R_i)}, \;\;i = 1,2
1736: \end{equation}
1737: %
1738: and $N_i(0^+)\neq N_i(0)$. This is not satisfactory and another type
1739: of expansion is needed if correct limits and better decay rates are to
1740: be found.
1741:
1742:
1743:
1744: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1745: \section{Rate Equations for the Case studies}\label{CaseRates}
1746: This appendix explains in a more detail how to derive the rate
1747: equations used in the cases studies in \sect{CaseStudies}. The
1748: equations are obtained from \eq{GeneralEq}. Also, \eq{LargeVApprox} is
1749: used to illustrate the impacts on dynamics from changes in the network
1750: structure in a less complicated form. The ${\cal N}_i(t)$,
1751: $\Delta_{ij}(t)$, $\kappa_{ij}(t)$ and $\sigma_{ij}(t)$ are defined in
1752: \sect{ProblemDef}. The initial distribution is set to be $N_j(0) =
1753: N_{j0}$ where $j=1,\ldots,M$ and $M$ is the total number of nodes in
1754: the system. Note the symmetry relations $a_{ij}=a_{ji}$,
1755: $\ell_{ij}=\ell_{ji}$ and $D_{ij}=D_{ji}$ which implies that
1756: $\kappa_{ij}(t) = \kappa_{ji}(t)$, $\Delta_{ij}(t) = \Delta_{ji}(t)$
1757: and $\sigma_{ij}(t)=\sigma_{ji}(t)$ (see Eq. \ref{GeneralRateCoef}).
1758:
1759:
1760:
1761: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1762: % THE LINE
1763: \subsection{The Line} \label{LineRate}
1764: The structure of the linear network studied here is shown in
1765: \fig{panel}~(a). Containers $C_1$ and $C_3$ are connected only to the
1766: middle container $C_2$. Therefore, the rate equations describing the
1767: dynamics in each container, $N_1(t)$ and $N_3(t)$ respectively, have
1768: the similar form:
1769: %
1770: \begin{equation}\begin{array}{lll}\label{LineRate1}
1771: \dot{N}_i(t) &= & \frac{V_{d-1}(a_{2i})}{V_d(R_2)}\int_0^t
1772: dt' {\cal N}_2(t')\,\sigma_{2i}(t-t')\\\\
1773:
1774: & & -\frac{V_{d-1}(a_{i2})}{V_d(R_i)}\int_0^t dt'{\cal N}_i(t')
1775: \left[\kappa_{i2}(t-t') \right.\\\\
1776:
1777: & &\left. + \Delta_{i2}(t-t')\right], \;\;\;\;\; i=1,3.
1778: \end{array}\end{equation}
1779: %
1780: The middle container $C_2$ is connected both to container $C_1$ and
1781: $C_3$. The rate equations for $N_2(t)$ reads
1782: %
1783: \begin{equation}\begin{array}{lll}\label{LineRate2}
1784: \dot{N}_2(t) & = & \frac{V_{d-1}(a_{12})}{V_d(R_1)}\int_0^t dt'
1785: {\cal N}_1(t')\,\sigma_{12}(t-t')\\\\
1786:
1787: & & +\frac{V_{d-1}(a_{32})}{V_d(R_3)}\int_0^t dt'
1788: {\cal N}_3(t')\,\sigma_{32}(t-t')\\\\
1789:
1790: & & -\frac{V_{d-1}(a_{21})}{V_d(R_2)}\int_0^t dt'
1791: {\cal N}_2(t') [\kappa_{21}(t-t')+\Delta_{21}(t-t')]\\\\
1792:
1793: & & -\frac{V_{d-1}(a_{23})}{V_d(R_2)}\int_0^t dt' {\cal N}_2(t')
1794: [\kappa_{23}(t-t')+\Delta_{23}(t-t')].
1795: \end{array}\end{equation}
1796: %
1797: Equations (\ref{LineRate1}) and (\ref{LineRate2}) have a quite
1798: complicated structure and impacts on the dynamics of $N_1(t)$,
1799: $N_2(t)$ and $N_3(t)$ due to changes in the structure (e.g tube
1800: lengths and container volumes) are not easily predicted. Simpler
1801: expressions can be obtained by using \eq{LargeVApprox} which is valid
1802: in the large time limit under the assumption that the number of
1803: particles in the tubes is negligible over time. Applying
1804: \eq{LargeVApprox} leads to
1805: %
1806: \begin{equation}\begin{array}{lll}
1807: \dot{N}_i(t) &= &\frac{V_{d-1}(a_{2i})}{V_d(R_2)}
1808: \frac{D}{\ell_{2i}}N_2(t) \\\\
1809: &&-\frac{V_{d-1}(a_{i2})}{V_d(R_i)}
1810: \frac{D}{\ell_{i2}}N_i(t),\;\;\;\;\;i = 1,3\\\\
1811:
1812: \dot{N}_2(t)&=&\frac{V_{d-1}(a_{12})}{V_d(R_2)}
1813: \frac{D}{\ell_{12}}N_1(t)+\frac{V_{d-1}(a_{32})}{V_d(R_3)}
1814: \frac{D}{\ell_{32}}N_3(t)\\\\
1815:
1816: &&\hfill-\frac{D}{V_d(R_2)}
1817: \left[\frac{V_{d-1}(a_{12})}{\ell_{12}}
1818: +\frac{V_{d-1}(a_{23})}{\ell_{23}}\right]N_2(t).
1819:
1820: %\dot{N}_3(t)&=&\frac{V_{d-1}(a_{23})}{V_d(R_2)}\frac{D}
1821: %{\ell_{23}}N_2(t)-\frac{V_{d-1}(a_{23})}{V_d(R_3)}\frac{D} {\ell_{23}}N_3(t).
1822:
1823: %\dot{N}_1(t) &= &\frac{V_{d-1}(a_{12})}{V_d(R_2)}
1824: %\frac{D}{\ell_{12}}N_2(t)
1825: %-\frac{V_{d-1}(a_{12})}{V_d(R_1)}
1826: %\frac{D}{\ell_{12}}N_1(t)\\\\\\
1827: \end{array}\end{equation}
1828:
1829:
1830: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1831: % THE TRIANGLE
1832: \subsection{The triangle}\label{TriangleRate}
1833: The triangular network studied here is depicted in
1834: \fig{panel}~(b). All the containers $C_1$, $C_2$ and $C_3$ are
1835: connected to each other and therefore the rate equations describing
1836: the time evolution of $N_1(t)$, $N_2(t)$ and $N_3(t)$ have the same
1837: form. Equation (\ref{TriangleRate1}) is the rate equation that governs
1838: the dynamics in container $C_1$. The corresponding dynamical equations
1839: for $N_2(t)$ and $N_3(t)$ are obtained in the same way but are not
1840: written down here
1841: %
1842: \begin{equation}\begin{array}{lll}\label{TriangleRate1}
1843: \dot{N}_1(t) & = & \frac{V_{d-1}(a_{21})}{V_d(R_2)}\int_0^t
1844: dt' {\cal N}_2(t')\,\sigma_{21}(t-t')\\\\
1845:
1846: & & +\frac{V_{d-1}(a_{31})}{V_d(R_3)}\int_0^t dt'
1847: {\cal N}_3(t')\,\sigma_{31}(t-t')\\\\
1848:
1849: & & -\frac{V_{d-1}(a_{12})}{V_d(R_1)}\int_0^t dt'
1850: {\cal N}_1(t')[\kappa_{12}(t-t') + \Delta_{12}(t-t')]\\\\
1851:
1852: & & -\frac{V_{d-1}(a_{13})}{V_d(R_1)}\int_0^t dt'
1853: {\cal N}_1(t')[\kappa_{13}(t-t') + \Delta_{13}(t-t')].
1854: \end{array}\end{equation}
1855: %
1856: %\begin{equation}\begin{array}{lll}\label{TriangleRate1}
1857: %\dot{N}_1(t)&=&\displaystyle\frac{V_{d-1}(a_{12})}{V_d(R_2)}\int_0^t
1858: %d\tau {\cal N}_2(\tau)\,\sigma_{21}(t-\tau)\\\\
1859: %&&\displaystyle+\frac{V_{d-1}(a_{31})}{V_d(R_3)}\int_0^t d\tau {\cal
1860: %N}_3(\tau)\,\sigma_{31}(t-\tau)\\\\
1861: %&&\displaystyle-\frac{V_{d-1}(a_{13})}{V_d(R_1)}\int_0^t d\tau {\cal
1862: %N}_1(\tau) \left[\kappa_{13}(t-\tau) + \Delta_{13}(t-\tau)\right.\\\\
1863: %&&\displaystyle\;\;+\left(\frac{a_{12}}{a_{31}}\right)^{d-1}
1864: %\left.\left(\kappa_{12}(t-\tau)+ \Delta_{12}(t-\tau)\right)\right]
1865: %\end{array}\end{equation}
1866: %
1867: %\begin{equation}\begin{array}{lll}\label{TriangleRate2}
1868: %\dot{N}_2(t)&=&\displaystyle\frac{V_{d-1}(a_{12})}{V_d(R_1)}\int_0^t
1869: %d\tau {\cal N}_1(\tau)\,\sigma_{12}(t-\tau)\\\\
1870: %&&\displaystyle+\frac{V_{d-1}(a_{23})}{V_d(R_3)}\int_0^t d\tau {\cal
1871: %N}_3(\tau)\,\sigma_{32}(t-\tau)\\\\
1872: %&&\displaystyle-\frac{V_{d-1}(a_{23})}{V_d(R_2)}\int_0^t d\tau {\cal
1873: %N}_2(\tau) \left[\kappa_{23}(t-\tau)+\Delta_{23}(t-\tau)\right.\\\\
1874: %&&\displaystyle\;\;+\left(\frac{a_{12}}{a_{23}}\right)^{d-1}\left.
1875: %\left(\kappa_{21}(t-\tau) +\Delta_{21}(t-\tau)\right)\right]
1876: %\end{array}\end{equation}
1877: %
1878: %\begin{equation}\begin{array}{lll} \label{TriangleRate3}
1879: %\dot{N}_3(t)&=&\displaystyle\frac{V_{d-1}(a_{31})}{V_d(R_1)}\int_0^t d\tau
1880: %{\cal N}_1(\tau)\,\sigma_{13}(t-\tau)\\\\
1881: %&&\displaystyle+\frac{V_{d-1}(a_{23})}{V_d(R_2)}\int_0^t d\tau
1882: %{\cal N}_3(\tau)\,\sigma_{23}(t-\tau)\\\\
1883: %&&\displaystyle-\frac{V_{d-1}(a_{23})}{V_d(R_3)}\int_0^t d\tau {\cal N}_3(\tau)
1884: %\left[\kappa_{32}(t-\tau)+\Delta_{32}(t-\tau)\right.\\\\
1885: %&&\displaystyle\;\;+\left(\frac{a_{31}}{a_{32}}\right)^{d-1}\left.\left(\kappa_{31}(t-\tau)+
1886: %\Delta_{31}(t-\tau)\right)\right].
1887: %\end{array}\end{equation}
1888: %
1889:
1890: The behavior of $N_1(t)$, $N_2(t)$ and $N_3(t)$ are very sensitive
1891: to changes in the network structure. The response from these
1892: changes are not easily predicted by \eq{TriangleRate1}. Instead
1893: \eq{LargeVApprox} can be used to state a simplified version of
1894: \eq{TriangleRate1}. Note that this equation only is valid in the
1895: large time limit if the number of particles in the tubes is small.
1896: Applying \eq{LargeVApprox} to container $C_1$ in the triangular
1897: network gives
1898: %
1899: \begin{equation}\begin{array}{lll}
1900: \dot{N}_1(t) & = \hfill
1901: & \frac{V_{d-1}(a_{21})}{V_d(R_2)}\frac{D}{\ell_{21}}N_2(t) +
1902: \frac{V_{d-1}(a_{31})}{V_d(R_3)}\frac{D}{\ell_{31}}N_3(t)\\\\
1903: && \hfill -\frac{D}{V_d(R_1)}\left[\frac{V_{d-1}(a_{12})}{\ell_{12}}+
1904: \frac{V_{d-1}(a_{13})}{\ell_{13}}\right]N_1(t).
1905: \end{array}\end{equation}
1906: %
1907: Corresponding first order rate equations for $N_2(t)$ and $N_3(t)$ are
1908: found in the same way.
1909: %
1910: %\begin{equation}\begin{array}{lll}
1911: %\dot{N}_2(t)&=&\displaystyle\frac{V_{d-1}(a_{12})}{V_d(R_1)}\frac{D}{\ell_{12}}N_1(t)+\frac{V_{d-1}(a_{23})}{V_d(R_3)}\frac{D}{\ell_{23}}N_3(t)\\\\
1912: %&&\displaystyle-\frac{D}{V_d(R_2)}\left(\frac{V_{d-1}(a_{12})}{\ell_{12}}
1913: %+\frac{V_{d-1}(a_{23})}{\ell_{23}}\right)N_2(t)
1914: %\end{array}\end{equation}
1915: %
1916: %\begin{equation}\begin{array}{lll}\label{TriangleAnalyt}
1917: %\dot{N}_3(t)&=&\displaystyle\frac{V_{d-1}(a_{31})}{V_d(R_1)}
1918: %\frac{D}{\ell_{31}}N_1(t)+ \frac{V_{d-1}(a_{23})}{V_d(R_2)}
1919: %\frac{D}{\ell_{23}}N_2(t)\\\\
1920: %&&\displaystyle-
1921: %\frac{D}{V_d(R_3)}\left(\frac{V_{d-1}(a_{31})}{\ell_{31}} +
1922: %\frac{V_{d-1}(a_{23})}{\ell_{23}}\right)N_3(t).
1923: %\end{array}\end{equation}
1924:
1925:
1926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1927: % THE T - NETWORK
1928: \subsection{The T-Network}\label{Trate}
1929: The T-shaped network under investigation in this subsection is
1930: depicted in \fig{panel}~(c). Since all containers $C_1$, $C_2$ and
1931: $C_3$ are connected to container $C_4$ and not to each other, the rate
1932: equations governing the dynamics in $C_1$, $C_2$ and $C_3$, that is
1933: $N_1(t)$, $N_2(t)$ and $N_3(t)$ respectively, all have similar form,
1934: %
1935: \begin{equation}\begin{array}{lll}\label{JunctionRate11}
1936: \dot{N}_i(t) & = & \frac{V_{d-1}(a_{4i})}{V_d(R_4)}\int_0^t dt'
1937: {\cal N}_4(t')\,\sigma_{4i}(t-t')\\\\
1938:
1939: & & -\frac{V_{d-1}(a_{i4})}{V_d(R_i)}\int_0^t dt'
1940: {\cal N}_i(t') \left[\kappa_{i4}(t-t') \right.\\\\
1941:
1942: & & \left. + \Delta_{i4}(t-t')\right], \;\;\;\;\; i=1,2,3.
1943: \end{array}\end{equation}
1944: %
1945: %\begin{equation}\begin{array}{lll}\label{JunctionRate11}
1946: %\dot{N}_1(t) &= & \frac{V_{d-1}(a)}{V_d(R_4)}\int_0^t d\tau
1947: %{\cal N}_4(\tau)\,\sigma_{14}(t-\tau)\\\\
1948: %&&-\frac{V_{d-1}(a)}{V_d(R_1)}\int_0^t d\tau {\cal N}_1(\tau)
1949: %\left[\kappa_{41}(t-\tau) + \Delta_{41}(t-\tau)\right]
1950: %\end{array}\end{equation}
1951: %
1952: %\begin{equation}\begin{array}{lll}\label{JunctionRate12}
1953: %\dot{N}_2(t) &= &\frac{V_{d-1}(a)}{V_d(R_4)}\int_0^t
1954: %d\tau {\cal N}_4(\tau)\,\sigma_{24}(t-\tau)\\\\
1955: %&&-\frac{V_{d-1}(a)}{V_d(R_2)}\int_0^t d\tau {\cal
1956: %N}_2(\tau) \left[\kappa_{42}(t-\tau) + \Delta_{42}(t-\tau)\right]
1957: %\end{array}\end{equation}
1958: %
1959: %\begin{equation}\begin{array}{lll}\label{JunctionRate13}
1960: %\dot{N}_3(t) &= &\frac{V_{d-1}(a)}{V_d(R_4)}\int_0^t d\tau
1961: %{\cal N}_4(\tau)\,\sigma_{34}(t-\tau)\\\\
1962: %&&-\frac{V_{d-1}(a)}{V_d(R_3)}\int_0^t d\tau {\cal
1963: %N}_3(\tau) \left[\kappa_{43}(t-\tau) + \Delta_{43}(t-\tau)\right].
1964: %\end{array}\end{equation}
1965: %
1966: The middle container $C_4$ has connections to all other containers
1967: $C_1$, $C_2$ and $C_3$. This leads to a rate equation for $N_4(t)$
1968: on the form
1969: %
1970: \begin{equation}\begin{array}{lll}\label{JunctionRate14}
1971: \dot{N}_4(t) &= & \sum_{j=1}^3 \frac{V_{d-1}(a_{j4})}{V_d(R_j)}
1972: \int_0^t dt' {\cal N}_j(t')\,\sigma_{j4}(t-t')\\\\
1973: & & -\frac{1}{V_d(R_4)}\int_0^t dt' {\cal N}_4(t')\\\\
1974: & & \times \sum_{j=1}^3 V_d(a_{4j})[\kappa_{4j}(t-t') +
1975: \Delta_{4j}(t-t')].
1976: \end{array}\end{equation}
1977: %
1978: The rate equations derived in this subsection are used in the study of
1979: the junction, shown in \fig{Junction1}.
1980:
1981:
1982:
1983: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1984: \section{Elimination of junctions in the asymptotic regime}\label{SectJunction}
1985: The network studied in this section is a generalization of the one
1986: depicted in \fig{Junction1}. Here, a junction having an arbitrary
1987: number of connections is studied and it will be demonstrated that the
1988: dynamical equations governing the transport in a system having
1989: junction points can be simplified in the large time limit. First, it
1990: will be shown that the current of particles in and out of a junction
1991: point eventually will balance each other out and that this occurs
1992: faster than the time it takes before an equilibrium particle
1993: distribution is attained throughout the network. This derives from the
1994: fact that the volume of the junction \mbox{(proportional to $ a^d$)}
1995: is much smaller than the volumes of the containers,
1996: \mbox{(proportional to $R_i^d$, $i=1,\ldots,M$)}, where $d$ is the
1997: dimensionality, $a$ is the tube radius and $M$ is the number of
1998: containers. The tubes connecting the junction are assumed to have the
1999: same radius. Second, it will be demonstrated how to simplify the
2000: transport equations in such a way that the dynamical variable for the
2001: junction $N(t)$ can be eliminated completely. This might be desirable
2002: when working with large networks.
2003:
2004: Consider a junction with dynamical behavior contained in $N_j(t)$ and
2005: volume $V_d(a)$ that has connections to $M$ containers with volumes
2006: $V_d(R_i)$, $i=1,\ldots,M$. The equation governing the dynamics of the
2007: junction point written in Laplace space reads
2008: %
2009: \begin{equation}\begin{array}{lll}\label{JuncRateInS}
2010: sN_j(s)-N_j(0) &= & -sN_j(s) \frac{V_{d-1}(a)}{V_d(a)}\sum_{i=1}^M\left[
2011: \kappa_{ji}(s) + \Delta_{ji}(s)\right]\\\\
2012: & & +\sum_{i=1}^M sN_i(s)\frac{V_{d-1}(a)}{V_d(R_i)}
2013: \,\sigma_{ij}(s).
2014: \end{array}\end{equation}
2015: %
2016: This equation is a generalization of \eqs{RateInS} (see
2017: \sect{SecGeneralEqs}). The junction point studied here is such that $a
2018: \ll R_i$ and since
2019: %
2020: \begin{equation}\begin{array}{lr}
2021: \frac{V_{d-1}(a)}{V_d(a)}= c_d\frac{1}{a},&
2022: \;\;\;\;\;\;\frac{V_{d-1}(a)}{V_d(R_i)} =
2023: c_d\frac{1}{a}\left(\frac{a}{R_i}\right)^d
2024: \end{array}\end{equation}
2025: %
2026: where $c_d = \frac{d}{d-1}\Gamma\left(\frac{d}{2}\right)
2027: [\Gamma\left(\frac{d-1}{2}\right)]^{-1}\,\pi^{-1/2}$ [found by using
2028: the formula for a $d$ - dimensional sphere, see \sect{ProblemDef}] the
2029: term proportional $(a/R_i)^d$ can be neglected and a closed expression
2030: for $N_j(s)$ can be obtained. Solving \eq{JuncRateInS} after eliminating
2031: the second term leads to
2032: %
2033: \begin{equation}\label{JuncSol}
2034: N_j(s) = \frac{N_j(0)}{s\left(1 + \frac{c_d}{a}\sum_{i=1}^M\left[
2035: \kappa_{ji}(s) + \Delta_{ji}(s)\right] \right)}.
2036: \end{equation}
2037: %
2038: The inverse Laplace transform of $N_j(s)$ is a sum of exponentials
2039: %
2040: \begin{equation}\label{SumOfExp}
2041: N_j(t) = \sum_{p=1}^\infty \alpha_p e^{-\beta_pt}
2042: \end{equation}
2043: %
2044: where $\alpha_p$ is the residue of $N_j(s)$ at pole $s = -\beta_p$,
2045: \mbox{$\beta_p\in \Re>0$}. The poles are given by the zeros of
2046: %
2047: \begin{equation}\label{TransEqJunction}
2048: s+c_d\sum_{i=1}^M\frac{q_{ij}D_{ij}}{a\ell_{ij}} \frac{\cosh
2049: q_{ji}} {\sinh q_{ij}}=0
2050: \end{equation}
2051: %
2052: where $q_{ij}^2=s\ell_{ij}^2/D_{ij}$. When the currents in and out of
2053: the junction point balances each other out there is no accumulation of
2054: particles and $\dot{N}_j(t)=0$. This is easily verified from
2055: \eq{SumOfExp} by evaluating the derivative in respect to $t$ at
2056: $t=\infty$
2057: %
2058: \begin{equation}
2059: \frac{d}{dt} N_j(t) = \sum_{p=1}^\infty (-\beta_p) \alpha_p
2060: e^{-\beta_pt} \rightarrow 0 \;\;{\rm as } \;\; t\rightarrow \infty.
2061: \end{equation}
2062: %
2063: Let $\tau_{\rm junction}$ be an estimate of the time it takes until
2064: $\dot{N}_j(t)=0$. It is related to the decay exponents $\beta_p$ (see
2065: Eqs. \ref{tau_s1} and \ref{q_s1}) which are zeros of
2066: \eq{TransEqJunction}. From \eq{TransEqJunction} it is clear that the
2067: zeros scales with $\frac{D}{a\ell}$ (it is assumed that $D_{ij}\sim D$
2068: and $\ell_{ij} \sim \ell$) which leads to the estimate $\tau_{\rm
2069: junction} \sim \frac{a\ell}{D}$. Let $\tau_{\rm network}$ be an
2070: estimate of the time it takes to reach an overall equilibrium particle
2071: distribution in the network. As a rough estimate \eq{Rate1} can be
2072: used which was found for a two-node network. If the containers have
2073: volumes proportional to $R^d$ then $\tau_{\rm network} \sim
2074: \frac{a\ell}{D}\left(\frac{R}{a}\right)^d$. If $R\gg a$ then
2075: $\tau_{\rm junction}\ll\tau_{\rm network}$. This is is supported by
2076: the numerical calculation shown in \fig{Junction} where the curve
2077: corresponding to $N_4(t)$ flattens out relatively fast.
2078:
2079: For networks where there are many junctions and containers involved it
2080: might be desirable to reduce the number of variables in the dynamical
2081: equations governing the particle transport. This can be done in the
2082: regime where $\dot{N}_j(t) = 0$ is valid and \eq{LargeVApprox} is
2083: applicable. Applying \eq{LargeVApprox} for a junction point $j$ having
2084: $M$ connections and using $\dot{N}_j(t)=0$ leads to
2085: %
2086: \begin{equation}
2087: 0 = \sum_{i=1}^M {\cal C}_{ij}V_{d-1}(a)\frac{D_{ij}}{\ell_{ij}}
2088: \left[\frac{N_i(t)}{V_d(R_i)}-\frac{N_j(t)}{V_d(a)}\right]
2089: \end{equation}
2090: %
2091: In this way it is possible to express $N_j(t)$ in terms of
2092: $N_1(t),\ldots,N_M(t)$. Inserting the solution to this matrix equation
2093: in the dynamical equations eliminates the explicit dependence of
2094: $N_j(t)$. \eqs{RateEqsJunction21} - (\ref{RateEqsJunction23}) shows
2095: this explicitly for the case of a three way junction depicted in
2096: \fig{Junction1}, where $N_4(t)$ has been taken away completely
2097: %
2098: \begin{equation}\begin{array}{lll}\label{RateEqsJunction21}
2099: \dot{N}_1(t) &= & \frac{\ell_{24} D}{\ell^2}
2100: \frac{V_{d-1}(a)}{V_d(R_2)}N_2(t) +\frac{\ell_{34}D}{\ell^2}
2101: \frac{V_{d-1}(a)}{V_d(R_3)}N_3(t)\\\\
2102: & & -\frac{(\ell_{24}+\ell_{34})D}{\ell^2}
2103: \frac{V_{d-1}(a)}{V_d(R_1)}N_1(t)
2104: \end{array}\end{equation}
2105: %
2106: \begin{equation}\begin{array}{lll}\label{RateEqsJunction22}
2107: \dot{N}_2(t) & = & \frac{\ell_{14}D}{\ell^2}
2108: \frac{V_{d-1}(a)}{V_d(R_1)}N_1(t)
2109: +\frac{\ell_{34}D}{\ell^2}\frac{V_{d-1}(a)}{V_d(R_3)}
2110: N_3(t)\\\\
2111: & & -\frac{(\ell_{14}+\ell_{34})D}
2112: {\ell^2}\frac{V_{d-1}(a)} {V_d(R_2)}N_2(t)\\\\\\
2113: \end{array}\end{equation}
2114: %
2115: \begin{equation}\begin{array}{lll}\label{RateEqsJunction23}
2116: \dot{N}_3(t) &= & \frac{\ell_{14}D}{\ell^2}
2117: \frac{V_{d-1}(a)}{V_d(R_2)}N_1(t)
2118: +\frac{\ell_{24}} {\ell^2}
2119: \frac{V_{d-1}(a)}{V_d(R_2)}N_2(t)\\\\
2120: & & -\frac{(\ell_{14}+\ell_{24})D}{\ell^2}
2121: \frac{V_{d-1}(a)}{V_d(R_3)}N_3(t)\\\\
2122: %
2123: \ell^2 &\equiv& \ell_{14}\ell_{24} +
2124: \ell_{24}\ell_{34} + \ell_{14}\ell_{34}.
2125: \end{array}\end{equation}
2126:
2127: \vfill
2128:
2129:
2130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2131: % R E F S
2132:
2133: %\bibliographystyle{my_prsty_title}
2134: %\bibliography{refs}
2135:
2136: \begin{thebibliography}{10}
2137:
2138: \bibitem{Orwar}
2139: A.~Karlsson, R.~Karlsson, M.~Karlsson, F.~Rytts\'en and O.~Orwar. Nature {\bf
2140: 409}, 150-153 (2001).
2141:
2142: \bibitem{KSMDKO}
2143: A.~Karlsson, K.~Sott, M.~Markström, M.~Davidson, Z.~Konkoli and O.~Orwar, J.
2144: Phys. Chem. B {\bf 109}(4), 1609-1617 (2005).
2145:
2146: \bibitem{KDKKBKJLHVO}
2147: M. Karlsson, M. Davidson, R. Karlsson, A. Karlsson, J. Bergenholtz, Z. Konkoli,
2148: A. Jesorka, T. Lobovkina, J. Hurtig, M. Voinova, and O. Orwar, Ann. Rev.
2149: Phys. Chem.~{\bf 55}, 613-649 (2004).
2150:
2151: \bibitem{LPHR}
2152: J.P.~Laplante, M.~Pemberton, A.~Hjelmfelt and J.~Ross, J. Phys. Chem. {\bf 99}
2153: (25), 10063 (1995).
2154:
2155: \bibitem{HR}
2156: A.~Hjelmfelt and J.~Ross, J. Phys. Chem. {\bf 97}, 7988 (1993).
2157:
2158: \bibitem{HSR}
2159: A.~Hjelmfelt, F.W.~Schneider and J.~Ross, Science {\bf 260}, 335 (1993).
2160:
2161: \bibitem{Grigoriev}
2162: I.V.~Grigoriev, Y.A.~Makhnovskii, A.M.~Berezhkovskii, V.Y.~Zitserman, J. Chem.
2163: Phys. {\bf 116}, 9574 (2002).
2164:
2165: \bibitem{Berezhkovskii}
2166: A.M.~Berezhkovskii and A.V.~Barzykin, Chem. Phys. Lett {\bf 383} 6-10 (2004).
2167:
2168: \bibitem{Dagdug}
2169: L.~Dagdug, A.M.~Berezhkovskii, S.Y.~Shvartsman, G.H.~Weiss, J. Chem. Phys.
2170: {\bf119}, 12473 (2003).
2171:
2172: \bibitem{Dagdug2}
2173: L.~Dagdug, A.M.~Berezhkovskii and G.H.~Weiss, Phys. Rev. E {\bf 69}, 012902
2174: (2004).
2175:
2176: \bibitem{Bezrukov}
2177: S.M.~Bezrukov, A.M.~Berezhkovskii, M.A.~Pustovit, A.~Szabo, J. Chem. Phys.
2178: {\bf113}, 8206-8211 (2000).
2179:
2180: \bibitem{BePu}
2181: H.C~Berg and E.M~Purcell, BioPhys. J. {\bf 20}, 193 (1977).
2182:
2183: \bibitem{Schaum}
2184: M.A~Spiegel, \emph{Schaum's outline of Laplace Transforms}, Cambridge
2185: University Press (1986).
2186:
2187: \bibitem{LePage}
2188: W.R.~LePage, \emph{Complex variables and the Laplace Transform for engineers},
2189: Dover Publications Inc., New York (1980).
2190:
2191: \bibitem{Stange}
2192: P.~Stange, D.~Zanette, A.~Mikhlov and B.~Hess, Biophys. Chem. {\bf 79}, 233-247
2193: (1999).
2194:
2195: \bibitem{NumRes}
2196: W.H.~Press, B.P.~Flannery, A.S.~Teukolsky and W.T.~Wetterling, \emph{Numerical
2197: Recipes}, Cambridge University Press, New York (1986).
2198:
2199: \bibitem{BD}
2200: T.~Bdineau and B.~Derrida, Phys. Rev. Lett {\bf 92}, 180601 (2004).
2201:
2202: \bibitem{JSI}
2203: R.Juh\'{a}sz, L.~Santen and F.~Igl\'{o}i, Phys. Rev. Lett. {\bf 94}, 010601
2204: (2005).
2205:
2206: \bibitem{EFCm}
2207: M.R.~Evans, D.P.~Foster, C.~God\`{e}che and D.~Mukamel, Phys. Rev. Lett. {\bf
2208: 74}, 208-211 (1995).
2209:
2210: \bibitem{SSFA}
2211: M.L.~Simpson,G.S.~Sayler, J.T.~Fleming and B. Applegate, Trends in
2212: Biotechnology {\bf 19}(8), 317 (2001).
2213:
2214: \bibitem{KK}
2215: V.~Krishnamurthy and E.V.~Krishnamurthy, BioSystems {\bf 49}, 205 (1999).
2216:
2217: \bibitem{Rambidi}
2218: N.G. Rambidi, BioSystems {\bf 44}, 1-15 (1997).
2219:
2220: \bibitem{CZ1}
2221: M.~Conrad and K.P.~Zauner, BioSystems {\bf 45}, 59 (1998).
2222:
2223: \bibitem{CZ2}
2224: K.P.~Zauner and M.~Conrad, Soft Computing {\bf 5}, 39 (2001).
2225:
2226: \bibitem{CZ3}
2227: K.P.~Zauner and M.~Conrad, Naturwissenschaften {\bf 87}, 360 (2000).
2228:
2229: \bibitem{Ji}
2230: S.~Ji, BioSystems {\bf 52}, 123 (1999).
2231:
2232: \bibitem{SM}
2233: C.~Siehs and B.~Mayer, Nanotechnology {\bf 10}, 464 (1999).
2234:
2235: \bibitem{Akingbehin}
2236: K.~Akingbehin, BioSystems {\bf 35}, 223 (1995).
2237:
2238: \bibitem{AkinCon}
2239: K.~Akingbehin and M.~Conrad, Journal of Parallel and Distributed, Computing
2240: {\bf 6}, 245 (1989).
2241:
2242: \bibitem{Conrad}
2243: M.~Conrad, European Journal of Operational Research {\bf 30}, 280 (1987).
2244:
2245: \bibitem{KirCon1}
2246: K.G.~Kirby and M.~Conrad,Physica D {\bf 22}, 205 (1986).
2247:
2248: \bibitem{KirCon2}
2249: K.G.~Kirby and M.~Conrad, Bulletin of Mathematical Biology {\bf 46} (5/6), 765
2250: (1984).
2251:
2252: \bibitem{KamCon1}
2253: R.~Kampfner and M.~Conrad, Bulletin of Mathematical Biology {\bf 45}(6), 931
2254: (1983).
2255:
2256: \bibitem{KamCon2}
2257: R.~Kampfner and M.~Conrad, Bulletin of Mathematical Biology {\bf 45}(6), 969
2258: (1983).
2259:
2260: \bibitem{HM}
2261: B.~Hess and M.Mikhailov, Science {\bf 264}, 223 (1994).
2262:
2263: \bibitem{Watts}
2264: D.J.~Watts and S.H.~Strogatz, Nature {\bf 393}, 440-442 (1998).
2265:
2266: \bibitem{WattsBook}
2267: D.J.~Watts, \emph{Small Worlds: The Dynamics of Networks between Order and
2268: Randomness}, (Princeton University, Princeton, NJ), 1999.
2269:
2270: \bibitem{Mohamed}
2271: L.M.~Delves, J.L.~Mohamed, \emph{Computational Methods for integral equations},
2272: Cambridge University Press, London (1985).
2273:
2274: \end{thebibliography}
2275:
2276:
2277:
2278: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2279: % INCLUDED FIGURES
2280: %
2281: \newpage
2282:
2283:
2284: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2285: % GENERAL NETWORK
2286: \begin{figure}
2287: \centering
2288: \psfrag{Ri}[t][t]{\footnotesize{$R_i$}}
2289: \psfrag{a}[t][t]{\footnotesize{$2a_{ij}$}}
2290: \includegraphics[width = 7cm]{MainNetwork.eps}
2291: \caption{ Schematic picture of an arbitrary network built from
2292: containers and tubes.}
2293: \label{MainNetwork}
2294: \end{figure}
2295:
2296:
2297: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2298: % STRUCTURE OF INFINITE TUBE
2299: \begin{figure}
2300: \psfrag{x = 0}[][]{$x=0$}
2301: \psfrag{R}[][]{$R$}
2302: \psfrag{2a}[][]{$2a$}
2303: \psfrag{1}[][]{$(a)$}
2304: \psfrag{2}[][]{$(b)$}
2305: \centering
2306: \includegraphics[width = 7cm]{InfTube.eps}
2307: \caption{Panel (a): A spherical compartment connected to a
2308: cylindrical infinitely long tube. If the tube radius is assumed to be
2309: small the compartment can be treated as ideally mixed at all times
2310: simplifying the dynamics in the container. The transport in the tube
2311: is reduced to a one dimensional diffusion problem. Panel (b)
2312: illustrates these simplifications.}
2313: \label{InfTubeFig}
2314: \end{figure}
2315:
2316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2317: % THE COUPLING FUNCTION f(y,z,t)
2318: % FOR A 2D CASE (I.E f(y,t))
2319: \begin{figure}
2320: \psfrag{t1}[][]{{\small $\frac{Dt_1}{\ell^2} = 0.001$}}
2321: \psfrag{t2}[][]{{\small $\frac{Dt_2}{\ell^2} = 0.01$}}
2322: \psfrag{t3}[][]{{\small $\frac{Dt_3}{\ell^2} = 1$}}
2323: \psfrag{xlabel}[][]{$y/2a$}
2324: \psfrag{ylabel}[][]{$2a f(y,t)$}
2325: \centering
2326: \includegraphics[width = 8cm]{couplingfun.eps}
2327: \caption{Behavior of $f(y,z,t)$ defined in
2328: \eq{DeCoupling} for the two dimensional system shown in
2329: \fig{SolDiffEq} [panel~(a), $a/R = 0.1$] at three different instants of
2330: time $t_1<t_2<t_3$. The graph verifies the assumption that
2331: $f(y,z,t)$ is approximately constant. In this two dimensional case
2332: $f(y,t) \rightarrow 1/2a$ as $t \rightarrow \infty$.}
2333: \label{InletProfile}
2334: \end{figure}
2335:
2336:
2337: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2338: % STRUCTURE OF TWO NODE NETWORK
2339: \begin{figure}
2340: \psfrag{x1}[][]{$x=0$}
2341: \psfrag{x2}[][]{$x=\ell$}
2342: \psfrag{a}[][]{$2a$}
2343: \psfrag{R1}[][]{$R_1$}
2344: \psfrag{R2}[][]{$R_2$}
2345: \centering
2346: \includegraphics[width = 7cm]{ReferenceNetwork.eps}
2347: \caption{Schematic picture of a two-node network.}
2348: \label{TwoContSytem}
2349: \end{figure}
2350:
2351:
2352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2353: % RATECOEFFICIENTS
2354: \begin{figure}
2355: \psfrag{xlabel}[][]{$Dt/\ell^2$}
2356: \psfrag{ylabel}[][]{$\displaystyle \frac{\ell}{D}\left[\Delta(t),
2357: \kappa(t),\sigma(t)\right]$}
2358: \centering
2359: \includegraphics[width = 8cm]{RateGraph.eps}
2360: \caption{Time dependence of the rate coefficients $\Delta(t)$,
2361: $\kappa(t)$ and $\sigma(t)$ from \eq{GeneralEq}.}
2362: \label{RateGraph}
2363: \end{figure}
2364:
2365:
2366: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2367: % NUMERICAL SOLUTION OF TWO NODE SYSTEM
2368: \begin{figure}
2369: \psfrag{xlabel}[][]{$Dt/\ell^2$}
2370: \psfrag{ylabel}[][]{$N_1(t)/N_{tot}$}
2371: \psfrag{V1}[][]{$N_1, V_1$}
2372: \psfrag{V2}[][]{$N_2, V_2$}
2373: \psfrag{LL}[][]{$\ell$}
2374: \psfrag{11}[][]{\small{$C_1$}}
2375: \psfrag{22}[][]{\small{$C_2$}}
2376: \centering
2377: \includegraphics[width = 8cm]{TwoNodes.eps}
2378: \caption{The solution of \eq{RateEqTwoCont} (solid line) is compared
2379: with a solution of \eq{ThirdAttempt} (dashed line). The solution of
2380: \eq{DagdugRateEqn}, given in (\ref{DagdugApprox}), is represented by
2381: the dotted line. The curves decaying and growing describe $N_1(t)$ and
2382: $N_2(t)$ respectively. The network structure is shown in inset. The
2383: volumes of the tube and the containers are equal and $a/\ell =
2384: 0.05$. The initial distribution of particles is $N_1(0)/N_{tot} = 1$
2385: and $N_2(0)/N_{tot}=0$, where $N_{tot}=N_1(0) + N_2(0)$, indicated by
2386: the shading in inset. This figure clearly shows that \eq{DagdugApprox}
2387: does not lead to the correct values for $N_1(t)$ and $N_2(t)$. The
2388: agreement with \eq{ThirdAttempt} is much better.}
2389: \label{NumSolTwoCont}
2390: \end{figure}
2391:
2392:
2393: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2394: % NUMERICAL SOLUTION OF TWO NODE SYSTEM (Log)
2395: \begin{figure}
2396: \psfrag{xlabel}[][]{$Dt/\ell^2$}
2397: \psfrag{ylabel}[][]{$\displaystyle \log \left[\frac{N_1(t)-N_1(\infty)}
2398: {N_{tot}}\right]$}
2399: \centering
2400: \includegraphics[width = 8cm]{LogTwoContainers.eps}
2401: \caption{The natural logarithm of $[N_1(t)-N_1(\infty)]/N_{tot}$ for
2402: the three cases illustrated in \fig{NumSolTwoCont}. The labeling of
2403: the curves is the same as in \fig{NumSolTwoCont}. The linear behavior
2404: is the evidence of the single exponential decay of the number of
2405: particles in the container $C_1$. The slope gives the value of the decay
2406: exponent. The slopes of the solid and dashed lines are close to each
2407: other showing that \eq{ThirdAttempt} is capable of estimating the
2408: decay exponent well. The slope of the dotted line differs
2409: significantly from the others which illustrates that \eq{DagdugApprox}
2410: can not describe the dynamics in an adequate way. Since the value of
2411: $N_1(\infty)$ is not the same as in all three cases [compare
2412: \eq{Ninfty} and (\ref{LimitDist})] the value of $N_1(0) - N_1(\infty)$
2413: will be different. This explains why all three curves do not coincide
2414: at $t=0$.}
2415: \label{LogTwoCont}
2416: \end{figure}
2417:
2418:
2419: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2420: % SOLUTION TO TRANSCENDENTAL EQUATION FOR
2421: % DECAY EXPONENTS
2422: \begin{figure}
2423: \psfrag{xlabel}[][]{$V_{\rm tube}/V_d$}
2424: \psfrag{ylabel}[][]{{\large $q_1^2$}}
2425: \centering
2426: \includegraphics[width = 8cm]{TransEqFig.eps}
2427: \caption{Dependence of the geometrical factor $q^2$ on the tube
2428: volume. (The volumes of the containers are equal
2429: $V_d(R_1)=V_d(R_2)\equiv V_d$.) $q_1^2$ and $\tau^{-1}$ are related
2430: through $\tau^{-1}= -q_1^2D/\ell^2$. The numerical solution of
2431: \eq{TransEq}, which gives the exact values for $q_1^2$, is represented
2432: by the solid line. It is compared with the values for $q_{1,a}^2$
2433: (Eq. \ref{Rate1}), dashed line, and $q_{1,b}^2$ (Eq. \ref{Rate3}),
2434: dotted line. The dotted line deviates significantly from the solid one
2435: as $V_{\rm tube}/V_d$ increases while the dashed line follows the
2436: exact solution better. This indicates that $q_{1,b}^2$ provides a good
2437: estimate of the decay rate, even for large tube volumes. $q_{1,a}^2$
2438: can only be used for very small values of $V_{\rm tube}/V_d$.}
2439: \label{TransEqFig}
2440: \end{figure}
2441:
2442:
2443: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2444: % SOLUTION TO DIFF.EQ. IN 2D
2445: \begin{figure}
2446: \psfrag{XX}[][]{$Dt/\ell^2$}
2447: \psfrag{YY}[][]
2448: {$\displaystyle \log \left[ \frac{N_1(t)-N_1(\infty)}{N_{tot}}\right]$}
2449: \psfrag{2a}[][]{\footnotesize{$2a$}}
2450: \psfrag{L}[][]{\footnotesize{$2R$}}
2451: \psfrag{l}[][]{\footnotesize{$\ell$}}
2452: \centering
2453: \includegraphics[width = 8cm]{2DVerification.eps}
2454: \caption{A numerical verification of the assumption of well stirred
2455: containers. For simplicity reasons a cubic geometry, as depicted in
2456: panel (a), was chosen since the explicit shape of the container looses
2457: its importance for large reservoir volumes. The edge length is set to
2458: $2R$ where $R$ is the radius of an equivalent spheric container.
2459: Panel (b) shows a numerical solution to the 2D-diffusion equation
2460: (solid line) compared to a numerical solution to \eqs{GeneralEq}
2461: (dashed line). The cases are chosen so that $a/R$ varies in three
2462: orders of magnitude showing increasing validity of the assumption of
2463: ideally mixed containers as $a/R$ decreases. The initial distribution
2464: in all three cases are skewed (a delta function in the bottom right
2465: corner of the left container) which becomes important when $a$ is
2466: large. For small $a$, $\tau_{{\rm mix}}$ is a lot shorter than
2467: $\tau_{{\rm target}}$ and the skewed initial distribution will have
2468: time to smear out before particles start exiting and the shape of the
2469: initial distribution has no effect.}
2470: \label{SolDiffEq}
2471: \end{figure}
2472:
2473:
2474:
2475: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2476: % STRUCTURE PANEL
2477: \begin{figure}
2478: \psfrag{L12}[][]{\footnotesize{$\ell_{12}$}}
2479: \psfrag{L23}[][]{\footnotesize{$\ell_{23}$}}
2480: \psfrag{L24}[][]{\footnotesize{$\ell_{24}$}}
2481: \psfrag{L31}[][]{\footnotesize{$\ell_{31}$}}
2482: \psfrag{L14}[][]{\footnotesize{$\ell_{14}$}}
2483: \psfrag{L34}[][]{\footnotesize{$\ell_{34}$}}
2484: \psfrag{N1}[][]{\small{$N_1$, $R_1$}}
2485: \psfrag{N2}[][]{\small{$N_2$,$R_2$}}
2486: \psfrag{N3}[][]{\small{$N_3$, $R_3$}}
2487: \psfrag{N4}[][]{\small{$N_4$, $R_4$}}
2488: \psfrag{1}[][]{\small{$C_1$}}
2489: \psfrag{2}[][]{\small{$C_2$}}
2490: \psfrag{3}[][]{\small{$C_3$}}
2491: \psfrag{4}[][]{\small{$C_4$}}
2492: \includegraphics[width = 8cm]{StructurePanel.eps}
2493: \caption{networks used for case studies in
2494: \sect{CaseStudies}. Panels~(a)-(d) corresponds to cases 1-4.}
2495: \label{panel}
2496: \end{figure}
2497:
2498:
2499: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2500: % NUMERICAL SOLUTION TO 'THE LINE'
2501: \begin{figure}[!ht]
2502: \centering
2503: \psfrag{XX}[][]{\footnotesize{$Dt/\ell^2$}}
2504: \psfrag{YY}[][]{\footnotesize{$N_i(t)/N_{tot}$}}
2505: \psfrag{A}[][]{$a)$}
2506: \psfrag{B}[][]{$b)$}
2507: \psfrag{N1}[][]{\small{$N_1$, $V_1$}}
2508: \psfrag{N2}[][]{\small{$N_2$, $V_2$}}
2509: \psfrag{N3}[][]{\small{$N_3$, $V_3$}}
2510: \psfrag{L12}[][]{\small{$\ell_{12}$}}
2511: \psfrag{L23}[][]{\small{$\ell_{23}$}}
2512: \psfrag{L31}[][]{\small{$\ell_{31}$}}
2513: \includegraphics[width = 8cm]{TheLine.eps}
2514: \caption{The transport properties of the structure depicted in
2515: \fig{panel}~(a), see \sect{CaseStudies} (case 1) for discussion. The
2516: curves depict a numerical solution of Eqs. (\ref{LineRate1}) and
2517: (\ref{LineRate2}) given in \app{LineRate}. The curve depicting the
2518: number of particles in the middle container $C_2$ (solid line) has a
2519: maximum. This kind of behavior does not exist for the two-node network
2520: where there is only exponential growth and decay (see
2521: \fig{NumSolTwoCont}). The network parameters were set to
2522: $\ell_{12}=\ell_{23}\equiv \ell$, $a/\ell= 1/5$, $a/R_1 = 1/10$,
2523: $a/R_1 = 1/15$, $a/R_3 = 1/20$. The initial distribution of particles
2524: is $N_1(0)/N_{tot} = 1$, $N_2(0)/N_{tot} = N_3(0)/N_{tot} = 0$, shown
2525: graphically in inset.}
2526: \label{LineGraph}
2527: \end{figure}
2528:
2529:
2530: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2531: % TIMING CONTROL FOR 'THE LINE
2532: \begin{figure}[!ht]
2533: \centering
2534: \psfrag{xlabel}[][]{\footnotesize{$Dt/\ell^2$}}
2535: \psfrag{ylabel}[][]{\footnotesize{$N_i(t)/N_{tot}$}}
2536: \includegraphics[width = 8cm]{TheLineTiming.eps}
2537: \caption{The control of the particle arrival time for three different
2538: cases shown in inset: $\ell=\ell_1$ (solid line), $\ell/\ell_2 = 1/5$
2539: (dash-dotted line) and $\ell/\ell_3 = 1/10$ (dashed line). The system
2540: is otherwise equivalent to the one studied in \fig{LineGraph}. The
2541: initial distribution of particles is shown graphically in inset.}
2542: \label{LineTiming}
2543: \end{figure}
2544:
2545:
2546:
2547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2548: % NUMERICAL SOLUTION TO 'THE TRIANGLE'
2549: \begin{figure}[!ht]
2550: \centering
2551: \psfrag{A}[][]{$a)$}
2552: \psfrag{B}[][]{$b)$}
2553: \psfrag{xlabel}[][]{\footnotesize{$Dt/\ell^2$}}
2554: \psfrag{ylabel}[][]{\footnotesize{$N_i(t)/N_{tot}$}}
2555: \includegraphics[width = 8cm]{TheTriangle.eps}
2556: \caption{The transport properties for the triangular network shown in
2557: \fig{panel} (b), see \sect{CaseStudies} (case 2) for discussion. The
2558: curves depict a numerical solution of \eq{TriangleRate1} given in
2559: \app{TriangleRate}. The curve depicting the number of particles in
2560: $C_2$ (solid line) has a minimum. This does not occur in the transport
2561: dynamics for the two-node network where there is only exponential
2562: growth and decay (see \fig{NumSolTwoCont}). The parameters were set to
2563: $\ell\equiv\ell_{21}$, $\ell/a=1$, $\ell_{23}/\ell=1/10$,
2564: $\ell_{31}/\ell=50$, $R_1/\ell = R_2/\ell = R_3/\ell=4$. The initial
2565: distribution of particles is $N_1(0)/N_{tot}=1$, $N_2(0)/N_{tot} = 0.5
2566: $, $N_3(0)/N_{tot}=0$ as indicated by shading in inset.}
2567: \label{Triangle2}
2568: \end{figure}
2569:
2570:
2571: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2572: % THE T-NETWORK
2573: \begin{figure}[!ht]
2574: \centering \psfrag{XX}[][]{$Dt/\ell^2$}
2575: \psfrag{YY}[][]{$N_i(t)/N_{tot}$}
2576: \includegraphics[width = 8cm]{TheLineTop.eps}
2577: \caption{The transport properties for the T-shaped network shown in
2578: \fig{panel} (c), see \sect{CaseStudies} (case 3) for discussion. The
2579: curves depict a numerical solution of \eqs{JunctionRate11} and
2580: (\ref{JunctionRate14}) given in \app{TriangleRate}. The figure shows
2581: two extremum points in the curve depicting the number of particles in
2582: container $C_2$ (solid line). It is not possible to have this kind of
2583: behavior for any of the cases studied involving two and three
2584: containers. The additional extremum point was produced by adding an
2585: additional container $C_4$ to the middle container $C_2$ in the linear
2586: structure shown in \fig{panel}~(a). The network parameters used were
2587: $\ell_{12}\equiv \ell$, $\ell/a = 3$, $\ell/\ell_{23}=1/2$,
2588: $\ell/\ell_{24}=10$, $R_1/\ell=2$. The initial distribution was set to
2589: $N_1(0)/N_{tot}=2/3$, $N_2(0)/N_{tot} = 0$, $N_3(0)/N_{tot} = 0$ and
2590: $N_4(0)/N_{tot} = 1/3$ according to shading in inset.}
2591: \label{LineGraphTop}
2592: \end{figure}
2593:
2594:
2595: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2596: % TOPOLOGY FOR 'THE JUNCTION'
2597: \begin{figure}
2598: \centering
2599: \psfrag{N1}[][]{\small{$N_1$, $V_1$}}
2600: \psfrag{N2}[][]{\small{$N_2$, $V_2$}}
2601: \psfrag{N3}[][]{\small{$N_3$, $V_3$}}
2602: \psfrag{N41}[][]{\small{$N_4$, $V_4$}}
2603: \psfrag{N42}[][]{\small{$N_4$, $V(a)$}}
2604: \psfrag{L14}[][]{\small{$\ell_{14}$}}
2605: \psfrag{L24}[][]{\small{$\ell_{24}$}}
2606: \psfrag{L34}[][]{\small{$\ell_{34}$}}
2607: \psfrag{limit}[][]{\small{$R_4\longrightarrow a$}}
2608: \includegraphics[width = 8cm]{junction1.eps}
2609: \caption{The transformation from a four-node network to a network
2610: involving a three way junction.}
2611: \label{Junction1}
2612: \end{figure}
2613:
2614:
2615:
2616: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2617: % NUMERICAL SOLUTION TO 'THE JUNCTION'
2618: \begin{figure}[!ht]
2619: \centering
2620: \psfrag{A}[][]{$a)$}
2621: \psfrag{B}[][]{$b)$}
2622: \psfrag{xlabel}[][]{\footnotesize{$Dt/\ell^2$}}
2623: \psfrag{ylabel}[][]{\footnotesize{$N_i(t)/N_{tot}$}}
2624: \includegraphics[width = 8cm]{TheJunction.eps}
2625: \caption{The transport properties for the network involving a three
2626: way junction shown in \fig{panel} (d), see \sect{CaseStudies} (case 4)
2627: for discussion. The curves depict a numerical solution of
2628: \eqs{JunctionRate11} and (\ref{JunctionRate14}) [with $V_d(R_4) =
2629: V_d(a)$]. The system parameters were set to $\ell_{12}/a=2$,
2630: $\ell_{13}/a=10$, $\ell_{14}/a=20$, $R_1/a = 2$ and $R_2/a = R_3/a =
2631: 4$. Initial distribution of particles $N_1(0)/N_{tot} = 1$,
2632: $N_2(0)/N_{tot} = N_3(0)/N_{tot}=N_4(0)/N_{tot}=0$ is shown in inset.}
2633: \label{Junction}
2634: \end{figure}
2635:
2636:
2637:
2638: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2639: % TOPOLOGY FOR THE 'LARGE NETWORK
2640: \begin{figure}[!ht]
2641: \centering
2642: \psfrag{1}[][]{\small{$C_1$}} \psfrag{2}[][]{\small{$C_2$}}
2643: \psfrag{3}[][]{\small{$C_3$}} \psfrag{4}[][]{\small{$C_4$}}
2644: \psfrag{5}[][]{\small{$C_5$}} \psfrag{6}[][]{\small{$C_6$}}
2645: \psfrag{7}[][]{\small{$C_7$}} \psfrag{8}[][]{\small{$C_8$}}
2646: \includegraphics[width = 7cm]{LargeNetwork.eps}
2647: \caption{Structure of the network studied in \sect{CaseStudies},
2648: case~5. The parameters describing the geometry are labeled in the
2649: same way as in \fig{panel}. The parameters used were $a/\ell=1/3$,
2650: $\ell_{23}/\ell=1$, $\ell_{45}/\ell=9$, $\ell_{43}/\ell=10$,
2651: $\ell_{35}/\ell=8$, $\ell_{36}/\ell=2$, $\ell_{28}/\ell=0.1$,
2652: $\ell_{12}/\ell=1$, $\ell_{32}/\ell=1$, $R_1/a = 4$, $R_2/a = 2$,
2653: $R_3/a = 1$, $R_4/a = 2.5$, $R_5/a = 3$, $R_6/a = 2$, $R_7/a = 2$,
2654: $R_8/a = 4$.}
2655: \label{LargeNetwork}
2656: \end{figure}
2657:
2658:
2659: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2660: % PANEL OF SOLUTIONS TO THE LARGE NETWORK
2661: \begin{figure}[!ht]
2662: \centering
2663: \psfrag{XX}[][]{\footnotesize{$Dt/\ell^2$}}
2664: \psfrag{YY}[][]{\footnotesize{$N_i(t)/N_{tot}$}}
2665: \includegraphics[width = 8cm]{PanelofNteworks.eps}
2666: \caption{The solution of \eq{GeneralEq} for the network depicted in
2667: \fig{LargeNetwork}. The panel includes graphs showing the transport
2668: dynamics for three different choices of initial distribution of
2669: particles. Panel~(a): $\tilde{N}_1(0) = \tilde{N}_2(0) =
2670: \tilde{N}_3(0) = \tilde{N}_4(0)= \tilde{N}_5(0) =0$, $\tilde{N}_6(0) =
2671: \tilde{N}_7(0) = \tilde{N}_8(0) = 1/3$; Panel~(b): $\tilde{N}_3(0) =
2672: \tilde{N}_5(0) = \tilde{N}_6(0)=0$, $\tilde{N}_2(0) =
2673: \tilde{N}_8(0)=1/8$ and $\tilde{N}_1(0) = \tilde{N}_4(0) =
2674: \tilde{N}_7(0)=1/4$; Panel~(c): $\tilde{N}_3(0) = \tilde{N}_5(0) =
2675: \tilde{N}_6(0)=0$, $\tilde{N}_1(0) = \tilde{N}_7(0) = 1/8$ and
2676: $\tilde{N}_2(0) = \tilde{N}_4(0) = \tilde{N}_8(0)=1/4$ where
2677: $\tilde{N}_i(t) = N_i(t)/N_{tot}$, $i=1,\ldots,8$. The initial
2678: conditions are also illustrated in the insets: a darker shading
2679: indicates that more particles are injected into the container.}
2680: \label{NetworkPanel}
2681: \end{figure}
2682:
2683: \end{document}
2684: