1: \documentclass[11pt]{article}
2: \usepackage{amsfonts}
3: \usepackage{amsmath,amsbsy}
4: \usepackage{epsfig}
5: \usepackage{graphics}
6:
7: \textheight 22.1 cm
8: \textwidth 16.1 cm
9: \oddsidemargin 0 mm
10: \topmargin -3 mm
11:
12: \newtheorem{theorem}{Theorem}
13: \newtheorem{corollary}{Corollary}
14: \newtheorem{lemma}{Lemma}
15: \newtheorem{definition}{Definition}
16: \newcommand{\er}{\mathbb{R}}
17: \newcommand{\en}{\mathbb{N}}
18: \newcommand{\boldx}{\mathbf{x}}
19: \newcommand{\boldalpha}{\boldsymbol{\alpha}}
20: \newcommand{\boldv}{\mathbf{v}}
21: \newcommand{\boldy}{\mathbf{y}}
22: \newcommand{\boldf}{\mathbf{f}}
23: \newcommand{\boldF}{\mathbf{F}}
24: \newcommand{\boldI}{\mathbf{I}}
25: \newcommand{\boldC}{\mathbf{C}}
26: \newcommand{\boldS}{\mathbf{S}}
27: \newcommand{\boldR}{\mathbf{R}}
28: \newcommand{\dx}{\mbox{d}x}
29: \newcommand{\nonexploding}{{nonexploding}}
30: \newcommand{\dC}{\mbox{d}C}
31: \newcommand{\dv}{\mbox{d}v}
32: \newcommand{\dy}{\mbox{d}y}
33: \newcommand{\dbx}{\mbox{d}\boldx}
34: \newcommand{\dbv}{\mbox{d}\boldv}
35: \newcommand{\dby}{\mbox{d}\boldy}
36: \newcommand{\dt}{\mbox{d}t}
37: \newcommand{\dxi}{\mbox{d}\xi}
38: \newcommand{\dtau}{\mbox{d}\tau}
39: \newcommand{\esssup}{\mathop{\mbox{ess}\,\mbox{sup}}}
40: \newcommand{\supp}{\mathop{\mbox{supp}}}
41: \newcommand{\trace}{\mathop{\mbox{{\rm trace}}}}
42:
43: \begin{document}
44:
45: \centerline{\bf \Large GLOBAL EXISTENCE RESULTS FOR COMPLEX}
46:
47: \bigskip
48:
49: \centerline{\bf \Large HYPERBOLIC MODELS OF BACTERIAL CHEMOTAXIS}
50:
51: \medskip
52: \bigskip
53:
54: \centerline{\large Radek Erban$^*$ and Hyung Ju Hwang$^\dagger$}
55:
56: \bigskip \medskip
57:
58: {%\footnotesize
59: \centerline{$^*$University of Oxford, Mathematical Institute} %
60: \centerline{24-29 St Giles', Oxford, OX1 3LB, United Kingdom} %
61: \centerline{e-mail: \textit{erban@maths.ox.ac.uk}} \medskip %
62: \centerline{$^\dagger$Trinity College Dublin, School of Mathematics} %
63: \centerline{Dublin 2, Ireland}
64: \centerline{e-mail:
65: \textit{hjhwang@maths.tcd.ie}} }
66:
67: \bigskip \medskip
68:
69: \begin{abstract} \noindent
70: Bacteria are able to respond to environmental signals by changing their
71: rules of movement. When we take into account chemical signals in the
72: environment, this behaviour is often called chemotaxis. At the
73: individual-level, chemotaxis consists of several steps. First, the cell
74: detects the extracellular signal using receptors on its membrane. Then, the
75: cell processes the signal information through the intracellular signal
76: transduction network, and finally it responds by altering its motile
77: behaviour accordingly. At the population level, chemotaxis can lead to
78: aggregation of bacteria, travelling waves or pattern formation, and the
79: important task is to explain the population-level behaviour in terms of
80: individual-based models. It has been previously shown that the transport
81: equation framework \cite{Erban:2004:ICB,Erban:2004:STS} is suitable for
82: connecting different levels of modelling of bacterial chemotaxis. In this
83: paper, we couple the transport equation for bacteria with the
84: (parabolic/elliptic) equation for the extracellular signals. We prove global
85: existence of solutions for the general hyperbolic chemotaxis models of cells
86: which process the information about the extracellular signal through the
87: intracellular biochemical network and interact by altering the extracellular
88: signal as well. The conditions for global existence in terms of the
89: properties of the signal transduction model are given.
90: \end{abstract}
91:
92: \section{Introduction}
93:
94: \label{secintro}
95:
96: The flagellated bacteria (e.g. \textit{Escherichia coli}, \textit{%
97: Salmone\-lla typhimurium}, \textit{Bacillus subtilis}) are single-celled
98: organisms. They are usually too small to be visible by the naked eye;
99: typically, they have the size of microns (see \cite%
100: {Salyers:2001:MDD,McKane:1996:MEA} for review). The behaviour of a bacterium
101: is primarily influenced by concentrations of various chemicals inside the
102: cell. Since bacteria are small, we can assume that the concentrations of the
103: chemicals inside the cytoplasm are uniform. Therefore, we can suppose that
104: the cells are points. Moreover, to create a mathematical description of a
105: bacterium, we introduce the vector of internal state variables \cite%
106: {Erban:2004:ICB,Erban:2004:STS,Erban:2005:ICB}
107: \begin{equation}
108: \mathbf{y}=(y_{1},y_{2},\dots ,y_{m})^{T}\in \mathbb{R}^{m}, \label{rom200}
109: \end{equation}%
110: where $y_{i},$ $i=1,\dots ,m,$ are concentrations of various chemicals
111: (proteins, receptor states etc.) inside the cell involved in the processes
112: of interest. The individual behaviour of a cell primarily depends on the
113: vector $\mathbf{y}$ which is a function of time. Consequently, the state of
114: bacterium is uniquely determined by the vector $(t,\mathbf{x},\mathbf{v},%
115: \mathbf{y})$ where $\mathbf{x}\in \mathbb{R}^{N}$ is the position of a cell,
116: $\mathbf{v}\in \mathbb{R}^{N}$ is its velocity, $\mathbf{y}\in \mathbb{R}^{m}
117: $ is its internal state, $t$ is time and $N=1,2,$ or $3,$ is the dimension
118: of the physical space.
119:
120: During its life, a cell must communicate with its environment in order to
121: find nutrients, to avoid repellents, to find mates etc. For this purpose,
122: there are receptors in the cellular membrane which can detect various
123: chemicals in the environment. We describe the chemicals outside the
124: bacterium by the signaling vector (which depends on the position of the cell
125: $\mathbf{x}$ and time $t$)
126: \begin{equation}
127: \mathbf{S}(\mathbf{x},t)=(S_{1},S_{2},\dots ,S_{M})^T \in \mathbb{R}^{M}.
128: \label{rom201}
129: \end{equation}%
130: Then the evolution of the internal state vector $\mathbf{y}$ depends also on
131: the signaling vector $\mathbf{S}.$ Since we describe chemical processes, we
132: can assume that $\mathbf{y}$ evolves according to system of ordinary
133: differential equations
134: \begin{equation}
135: \frac{\mbox{d}\mathbf{y}}{\mbox{d}t}=\mathbf{F}(\mathbf{S}(\mathbf{x}),%
136: \mathbf{y}). \label{rom14}
137: \end{equation}%
138: This system formally captures all biochemistry inside the cell and
139: therefore, the concrete form of the vector function $\mathbf{F}:\mathbb{R}%
140: ^{M}\times \mathbb{R}^{m}\rightarrow \mathbb{R}^{m}$ can be very complicated
141: depending on the number of details which are included in the model.
142:
143: Bacterial movement and the signal transduction network (\ref{rom14}) will be
144: discussed in more details in Section \ref{secbiodetails}. From the
145: mathematical point of view, the movement of the flagellated bacteria can be
146: viewed as a biased random walk. The properties of this random walk depend on
147: the internal state $\mathbf{y}$ and bacterial velocity $\mathbf{v}$. The
148: classical description of the bacterial movement is the so called velocity
149: jump process \cite{Othmer:1988:MDB,Erban:2004:ICB,Erban:2004:STS}. It means
150: that the bacterium runs with some velocity and at random instants of time it
151: changes its velocity according to the Poisson process with the intensity $%
152: \lambda(\mathbf{y})$.
153:
154: Let $f(\mathbf{x},\mathbf{v},\mathbf{y},t)$ be the density function of
155: bacteria in a $(2N+m)-$dimensional phase space with coordinates $(\mathbf{x},%
156: \mathbf{v},\mathbf{y})$ where $\mathbf{x}\in \mathbb{R}^{N}$ is the position
157: of a cell, $\mathbf{v}\in V\subset \mathbb{R}^{N}$ is its velocity and $%
158: \mathbf{y}\in \mathbb{R}^{m}$ is its internal state, which evolves according
159: to (\ref{rom14}). Thus $f(\mathbf{x},\mathbf{v},\mathbf{y},t)\mbox{d}\mathbf{%
160: x}\mbox{d}\mathbf{v}\mbox{d}\mathbf{y}$ is the number of cells with position
161: between $\mathbf{x}$ and $\mathbf{x}+\mbox{d}\mathbf{x}$, velocity between $%
162: \mathbf{v}$ and $\mathbf{v}+\mbox{d}\mathbf{v}$, and internal state between $%
163: \mathbf{y}$ and $\mathbf{y}+\mbox{d}\mathbf{y}.$ Then evolution of $f$ is
164: governed by the following transport equation \cite%
165: {Erban:2004:ICB,Erban:2004:STS}
166: \begin{equation}
167: \frac{\partial f}{\partial t}+\nabla _{\mathbf{x}}\cdot \mathbf{v}f+\nabla _{%
168: \mathbf{y}}\cdot \mathbf{F}(\mathbf{S}(\mathbf{x}),\mathbf{y})f=-\lambda (%
169: \mathbf{y})f+\int_{V}\lambda (\mathbf{y})K(\mathbf{v},\mathbf{v}^{\prime },%
170: \mathbf{y})f(\mathbf{x},\mathbf{v}^{\prime },\mathbf{y},t)\mbox{d}\mathbf{v}%
171: ^{\prime } \label{vjpint}
172: \end{equation}%
173: where the kernel $K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})$ gives the
174: probability of a change in velocity from $\mathbf{v}^{\prime }$ to $\mathbf{v%
175: }$, given that a reorientation occurs. We assume that the random velocity
176: changes are the result of a Poisson process of intensity $\lambda (\mathbf{y}%
177: )$ \cite{Block:1983:AKB}. The kernel $K$ is non-negative and satisfies the
178: normalization condition
179: \begin{equation}
180: \int_{V}K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})\mbox{d}\mathbf{v}=1
181: \label{normalcond}
182: \end{equation}%
183: where
184: \begin{equation}
185: V\text{ \ is a symmetric compact set in }\mathbb{R}^{N}. \label{setvel}
186: \end{equation}%
187: Realistic examples of the kernel $K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y%
188: })$, set $V$, signal transduction network $\mathbf{F}$ and the turning
189: frequency $\lambda (\mathbf{y})$ are given in Section \ref{secbiodetails}.
190: They all satisfy the above basic assumptions.
191:
192: To write equation (\ref{vjpint}) in more compact form, we introduce the
193: kernel $T$ defined as a product of the turning frequency $\lambda $ and the
194: kernel $K$, i.e.
195: \begin{equation}
196: T(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})=\lambda \left( \mathbf{y}%
197: \right) K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y}). \label{kernelT}
198: \end{equation}%
199: Moreover, our goal is to couple equation (\ref{vjpint}) with the realistic
200: system of partial differential equations for the extracellular signal vector
201: $\mathbf{S}.$ We assume that the external signal diffuses. It can be also
202: produced or degraded by bacteria, degraded on its own or the components of $%
203: \mathbf{S}$ can react with each other in the extracellular space. Hence, the
204: general hyperbolic system of interest can be written in the following form:
205: \begin{eqnarray}
206: \frac{\partial f}{\partial t}+\nabla _{\mathbf{x}}\cdot \mathbf{v}f+\nabla _{%
207: \mathbf{y}}\cdot \mathbf{F}(\mathbf{S}(\mathbf{x}),\mathbf{y})f &=&\int_{V}T(%
208: \mathbf{v},\mathbf{v}^{\prime },\mathbf{y})\Big[f(\mathbf{v}^{\prime })-f(%
209: \mathbf{v})\Big]\mbox{d}\mathbf{v}^{\prime } \label{hyperp} \\
210: \frac{\partial \mathbf{S}}{\partial t} &=&D\triangle \mathbf{S}+\mathbf{R}(%
211: \mathbf{S},n) \label{parabS}
212: \end{eqnarray}%
213: where $n\equiv n(\mathbf{x},t)$ is the macroscopic density of individuals at
214: point $\mathbf{x}\in \mathbb{R}^{N}$ and time $t$ given as
215: \begin{equation}
216: n(\mathbf{x},t)=\int_{\mathbb{R}^{m}}\int_{V}f(\mathbf{x},\mathbf{v},\mathbf{%
217: y},t)\mbox{d}\mathbf{v}\mbox{d}\mathbf{y}, \label{equationforn}
218: \end{equation}%
219: D is a diagonal $M\times M$ matrix which diagonal elements are diffusion
220: constants of different chemicals in the extracellular signal vector $\mathbf{%
221: S}$ and the term $\mathbf{R}:\mathbb{R}^{M}\times \mathbb{R}\rightarrow
222: \mathbb{R}^{M}$ describes the creation, reaction and degradation of the
223: signals.
224:
225: The goal of this paper is to prove global existence results for the system (%
226: \ref{hyperp}) -- (\ref{parabS}). We will focus on one-dimensional case in
227: what follows. In Section \ref{secmotiv}, we will start with a simple model
228: of the signal transduction (\ref{simmodel}) which was used previously \cite%
229: {Erban:2004:ICB,Erban:2004:STS}. The simple model (\ref{simmodel}) has the
230: essential properties of the realistic models of the signal transduction, but
231: it is more tractable from the mathematical point of view than more complex
232: models of bacterial chemotaxis. We prove the global existence of solutions
233: of the one-dimensional version of (\ref{hyperp}) -- (\ref{parabS}) with the
234: simplified model of the signal transduction.
235:
236: In order to study the general case, we first review the relevant biology in
237: Section \ref{secbiodetails}. This will help us to specify the realistic
238: conditions on signal transduction model $(\ref{rom14})$, turning frequency $%
239: \lambda(\mathbf{y})$, turning kernel $K(\mathbf{v},\mathbf{v}^{\prime },%
240: \mathbf{y})$, set $V$, diffusion matrix $D$ and the reaction term $\mathbf{R}%
241: (\mathbf{S},n)$ in equations (\ref{hyperp}) --(\ref{parabS}). In Section \ref%
242: {secglobalexistence}, we study the global existence for the system (\ref%
243: {hyperp}) -- (\ref{parabS}) for general models of bacterial signal
244: transduction which are introduced in Section \ref{secbiodetails}. We also
245: consider that equation (\ref{parabS}) is at quasi-equilibrium, i.e. we
246: consider the elliptic equation for the signal in Section \ref%
247: {secglobalexistence}.
248:
249: Hence, this paper consists of two main mathematical results. First, we prove
250: the global existence of solutions to the problem (\ref{hyperp}) -- (\ref%
251: {parabS}) for the simplified model of signal transduction (\ref{simmodel})
252: and for the system of parabolic equations (\ref{parabS}) for the
253: extracellular signal (see Section \ref{secmotiv}, Theorem \ref{theorem1}).
254: Then, we prove the global existence of solutions for the general model of
255: signal transduction (\ref{rom14}) coupled with the system of elliptic
256: equations for the extracellular signal (see Section \ref{secglobalexistence}%
257: , Theorem \ref{theorem2}). The necessary growth assumptions on turning
258: frequency $T(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})$ are given in terms
259: of the signal derivative along the cell trajectory. It means that the growth
260: estimates on $T$ include the temporal derivative as well as the spatial
261: derivative of the extracellular signal. Finally, we provide discussion and
262: comparison with relevant results from the literature in Section \ref%
263: {secdiscussion}.
264:
265: \section{Global existence for a simplified model of signal transduction}
266:
267: \label{secmotiv}
268:
269: A simplified model of excitation-adaptation dynamics was studied in \cite%
270: {Erban:2004:ICB,Erban:2004:STS,Othmer:1998:OCS,Dolak:2005:KMC}
271: where $\mathbf{y}%
272: =(y_{1},y_{2})^{T}\in \mathbb{R}^{2}$ and the right hand side of equation (%
273: \ref{rom14}) was given as
274: \begin{equation}
275: \mathbf{F}\equiv \left(
276: \begin{array}{c}
277: F_{1} \\
278: F_{2}%
279: \end{array}%
280: \right) =\left(
281: \begin{array}{c}
282: \displaystyle\frac{g(\mathbf{S}(\mathbf{x},t))-(y_{1}+y_{2})}{t_{e}}%
283: \raisebox{-5.6mm}{\rule{0pt}{12.2mm}} \\
284: \displaystyle\frac{g(\mathbf{S}(\mathbf{x},t))-y_{2}}{t_{a}}%
285: \end{array}%
286: \right) \label{simmodel}
287: \end{equation}%
288: where $t_{e}$ and $t_{a}$ are positive constants and $g:\mathbb{R}%
289: ^{M}\rightarrow \lbrack 0,\infty )$. We will see in Section \ref%
290: {secbiodetails} that the simplified model (\ref{simmodel}) has the essential
291: properties of realistic signal transduction models. Hence, the model (\ref%
292: {simmodel}) is a natural starting point of this paper. For simplicity, we
293: work in a one-dimensional physical space, i.e. $N=1,$ and the goal of this
294: section is to prove Theorem \ref{theorem1} about the system (\ref{hyperp})
295: -- (\ref{parabS}). In what follows, we denote $L^{p}(\Omega )$, $1\leq p\leq
296: \infty $, $\Omega \subset \mathbb{R}^{d}$, the Banach space of measurable
297: functions with the finite norms
298: \begin{equation*}
299: \Vert h\Vert \hbox{\raise -0.5mm \hbox{$_{L^p}$}}=\left( \int_{\Omega }|h(%
300: \mathbf{x})|^{p}\mbox{d}\mathbf{x}\right) ^{1/p},\;\;\mbox{for}\;1\leq
301: p<\infty ,\qquad \mbox{and}\qquad \Vert h\Vert
302: \hbox{\raise -0.5mm
303: \hbox{$_{L^\infty}$}}=\mathop{\mbox{ess}\,\mbox{sup}}_{\Omega }|h(\mathbf{x}%
304: )|.
305: \end{equation*}%
306: We denote $W^{k,p}(\Omega )$, $1\leq p\leq \infty $, $\Omega \subset \mathbb{%
307: R}^{d}$, the usual Sobolev space
308: \begin{equation*}
309: W^{k,p}(\Omega )=\left\{ h\in L^{p}(\Omega )\,|\,\forall \boldsymbol{\alpha}%
310: \in \mathbb{N}_{0}^{d},|\boldsymbol{\alpha}|\leq k\Rightarrow \frac{\partial
311: ^{|\boldsymbol{\alpha}|}h}{\partial x_{1}^{\alpha _{1}}x_{2}^{\alpha
312: _{2}}\dots x_{d}^{\alpha _{d}}}\in L^{p}(\Omega )\right\}
313: \end{equation*}%
314: where $\boldsymbol{\alpha}=(\alpha _{1},\alpha _{2},\dots ,\alpha _{d})\in
315: \mathbb{N}_{0}^{d}$ is a vector of nonnegative integers and $|%
316: \boldsymbol{\alpha}|=\alpha _{1}+\alpha _{2}+\dots +\alpha _{d}.$ The norm
317: in $W^{k,p}(\Omega )$ is defined as
318: \begin{equation*}
319: \Vert h\Vert \hbox{\raise -0.5mm \hbox{$_{W^{k,p}}$}}=\sum_{%
320: \boldsymbol{\alpha}\in \mathbb{N}_{0}^{d},\,|\boldsymbol{\alpha}|\leq
321: k}\left\Vert \frac{\partial^{|\boldsymbol{\alpha}|}h}{\partial
322: x_{1}^{\alpha _{1}}x_{2}^{\alpha _{2}}\dots x_{d}^{\alpha_{d}}}\right\Vert
323: _{L^{p}}
324: \end{equation*}%
325: To simplify mathematical formulas, we will make use of the following
326: notation. For any function $h:\mathbb{R}^{d}\rightarrow \mathbb{R}$, $\nabla
327: h$ denotes the gradient of $h$ with respect to all variables and $\nabla
328: _{\!x_{1}x_{2}}h$ is the 2-dimensional gradient vector with respect to the
329: variables $x_{1}$ and $x_{2}$ only, i.e.
330: \begin{equation}
331: \nabla h=\left( \frac{\partial h}{\partial x_{1}},\frac{\partial h}{\partial
332: x_{2}},\dots ,\frac{\partial h}{\partial x_{d}}\right) ,\qquad \mbox{and}%
333: \qquad \nabla _{\!x_{1}x_{2}}h=\left( \frac{\partial h}{\partial x_{1}},%
334: \frac{\partial h}{\partial x_{2}}\right) . \label{gradnotation}
335: \end{equation}%
336: We already made use of this notation in equation (\ref{hyperp}) where the
337: gradients of the function $f$ were taken only with respect to the selected
338: parts of the state vector. In this section, we study the movement of cells
339: in one dimension, i.e. $N=1$. Moreover, we assume that the external signal
340: diffuses and it is produced by bacteria and degraded on its own. Hence, the
341: system of equations (\ref{hyperp}) -- (\ref{parabS}) reads as follows
342: \begin{equation}
343: \frac{\partial f}{\partial t}+\nabla _{\!x}\cdot vf+\nabla _{\!\mathbf{y}%
344: }\cdot \mathbf{F}(\mathbf{S}(\mathbf{x}),\mathbf{y})f=\int_{V}T(v,v^{\prime
345: },\mathbf{y})\Big[f(v^{\prime })-f(v)\Big]\mbox{d}v^{\prime }
346: \label{hyperpsim}
347: \end{equation}%
348: \begin{equation}
349: \frac{\partial S_{i}}{\partial t}=d_{i}\frac{\partial ^{2}S_{i}}{\partial
350: x^{2}}+k_{i}n-k_{i}^{0}S_{i},\qquad \;i=1,\dots ,M, \label{parabSsim}
351: \end{equation}%
352: where $d_{i}$, $k_{i}$ and $k_{i}^{0}$ are positive constants and $n\equiv
353: n(x,t)$ is the macroscopic density of individuals at point $x\in \mathbb{R}$
354: and time $t$ given by (\ref{equationforn}). Position $x$ and velocity $v$
355: are scalars for $N=1$, so we do not use bold letters for position and
356: velocity in equation (\ref{hyperpsim}). Otherwise, equation (\ref{hyperpsim}%
357: ) is the same as equation (\ref{hyperp}). Following notation (\ref%
358: {gradnotation}), symbol $\nabla _{\!x}f$ denotes the partial derivative of
359: distribution function $f$ with respect to $x.$ Let us note that (depending
360: on the form of function $g$ in (\ref{simmodel})) some extracellular signals
361: might be attractants and some extracellular signals might be repellents. If
362: we have sufficient growth estimates on function $g$ and kernel $T$, we can
363: guarantee the global existence of solutions of system (\ref{hyperpsim}) -- (%
364: \ref{parabSsim}) as it is shown in the following theorem.
365:
366: \begin{theorem}
367: Consider that the function $\mathbf{F}$ is given by $(\ref{simmodel})$.
368: Assume that there exist non-decreasing positive continuous functions $\Phi
369: ,\Psi \in C\left( \mathbb{R}\right) $ satisfying
370: \begin{equation}
371: \left\vert g(\mathbf{z})\right\vert +\left\vert \nabla g(\mathbf{z}%
372: )\right\vert \leq \Phi \left( \left\vert \mathbf{z}\right\vert \right) \quad
373: \text{and}\quad \left\vert T(v,v^{\prime },\mathbf{y})\right\vert
374: +\left\vert \nabla T(v,v^{\prime },\mathbf{y})\right\vert \leq \Psi \left(
375: \left\vert \mathbf{y}\right\vert \right) . \label{growthassumptions}
376: \end{equation}%
377: Assume that $f_{0}\in W^{1,1}({}\mathbb{R}\times V\times \mathbb{R}^{2})\cap
378: W^{1,\infty }({}\mathbb{R}\times V\times \mathbb{R}^{2})$ with compact
379: support and $\mathbf{S}_{0}\in \lbrack W^{1,\infty }({}\mathbb{R})]^{M}$
380: with compact support. Then there exist global solutions of the system $(\ref%
381: {hyperpsim})$ -- $(\ref{parabSsim})$ satisfying
382: \begin{equation}
383: f(\cdot ,\cdot ,\cdot ,t)\in W^{1,1}({}\mathbb{R}\times V\times \mathbb{R}%
384: ^{2})\cap W^{1,\infty }({}\mathbb{R}\times V\times \mathbb{R}^{2}),
385: \label{spacef}
386: \end{equation}%
387: \begin{equation}
388: \mathbf{S}(\cdot ,t)\in \left[ W^{1,\infty }(\mathbb{R})\right] ^{M}
389: \label{spaceS}
390: \end{equation}%
391: and initial conditions $f(\cdot ,\cdot ,\cdot ,0)=f_{0}(\cdot ,\cdot ,\cdot )
392: $ and $\mathbf{S}(\cdot ,0)=\mathbf{S}_{0}(\cdot )$. \label{theorem1}
393: \end{theorem}
394:
395: \bigskip
396:
397: \noindent First, the characteristics of the hyperbolic equation (\ref%
398: {hyperpsim}) are given for $N=1$ as
399: \begin{equation}
400: \frac{\mbox{d}X}{\mbox{d}s}=V,\quad \frac{\mbox{d}V}{\mbox{d}s}=0,\quad
401: \frac{\mbox{d}\mathbf{Y}}{\mbox{d}s}=\mathbf{F}(\mathbf{S}\left( X\left(
402: s\right) ,s\right) ,\mathbf{Y}(s)). \label{chareqn}
403: \end{equation}%
404: Then along back-time characteristics starting at $\left( x,v,\mathbf{y}%
405: ,t\right) $, we have for $0\leq s\leq t,$%
406: \begin{eqnarray}
407: X\left( s;x,v\mathbf{,y},t\right) &=&x-v\left( t-s\right) ,\ \ \label{X} \\
408: \mathbf{Y}\left( s;x,v\mathbf{,y},t\right) &=&\mathbf{y}-\int_{s}^{t}\mathbf{%
409: F}\left( \mathbf{S}\left( X(\tau ),\tau \right) ,\mathbf{Y}(\tau )\right) %
410: \mbox{d}\tau . \label{Y}
411: \end{eqnarray}%
412: Next, we will prove several auxiliary lemmas.
413:
414: \begin{lemma}
415: Derivation of the characteristics $(\ref{X})$ and $(\ref{Y})$ with respect
416: to the initial conditions gives, for $0\leq s\leq t,$
417: \begin{equation}
418: \frac{\partial X}{\partial x}=1,\quad \frac{\partial \mathbf{Y}}{\partial
419: \mathbf{y}}=\exp \left[ \frac{\partial \mathbf{F}}{\partial \mathbf{y}}(s-t)%
420: \right] \qquad \mbox{where}\quad \frac{\partial \mathbf{F}}{\partial \mathbf{%
421: y}}=\left(
422: \begin{array}{cc}
423: -\displaystyle\frac{1}{t_{e}}\raisebox{-5.6mm}{\rule{0pt}{12.2mm}} & -%
424: \displaystyle\frac{1}{t_{e}} \\
425: 0 & -\displaystyle\frac{1}{t_{a}}%
426: \end{array}%
427: \right) . \label{derchar}
428: \end{equation}%
429: Moreover,
430: \begin{equation}
431: \det \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}=\exp \left[ \left(
432: \frac{1}{t_{e}}+\frac{1}{t_{a}}\right) \left( t-s\right) \right] \geq 1.
433: \label{determyy}
434: \end{equation}%
435: \label{lemma3}
436: \end{lemma}
437:
438: \noindent \textbf{Proof.} We differentiate (\ref{Y}) with respect to $%
439: \mathbf{y}$ to get%
440: \begin{equation}
441: \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}=\mathbf{I}_{2}+\int_{t}^{s}%
442: \frac{\partial \mathbf{F}}{\partial \mathbf{y}}\frac{\partial \mathbf{Y}}{%
443: \partial \mathbf{y}}\left( \tau \right) d\tau . \label{derY}
444: \end{equation}%
445: where $\mathbf{I}_{2}$ is the $2\times 2$ identity matrix. Let
446: \begin{equation*}
447: \mathbf{G}\left( s\right) =\int_{t}^{s}\frac{\partial \mathbf{F}}{\partial
448: \mathbf{y}}\frac{\partial \mathbf{Y}}{\partial \mathbf{y}}\left( \tau
449: \right) d\tau ,
450: \end{equation*}%
451: then we have%
452: \begin{equation*}
453: \mathbf{G}^{\prime }(s)=\frac{\partial \mathbf{F}}{\partial \mathbf{y}}\frac{%
454: \partial \mathbf{Y}}{\partial \mathbf{y}}(s).
455: \end{equation*}%
456: Using (\ref{derY}), we obtain
457: \begin{equation*}
458: \mathbf{G}^{\prime }\left( s\right) -\frac{\partial \mathbf{F}}{\partial
459: \mathbf{y}}\mathbf{G}\left( s\right) =\frac{\partial \mathbf{F}}{\partial
460: \mathbf{y}}.
461: \end{equation*}%
462: Integrating the last equation, we have
463: \begin{equation*}
464: \mathbf{G}\left( s\right) =\exp \left[ \frac{\partial \mathbf{F}}{\partial
465: \mathbf{y}}(s-t)\right] -\mathbf{I}_{2},
466: \end{equation*}%
467: which deduce (\ref{derchar}). Computing the determinant of (\ref{derchar}),
468: we derive (\ref{determyy}).
469:
470: \rightline{Q.E.D.}
471:
472: \bigskip
473:
474: \begin{lemma}
475: Let us assume $(\ref{simmodel})$ and $(\ref{growthassumptions})$. Then the
476: solution of $(\ref{chareqn})$ satisfies
477: \begin{equation}
478: |\mathbf{Y}(\tau)|\leq C\left\{ 1+\Phi \left( \sup_{0\leq s\leq \tau
479: }\left\vert \mathbf{S}\left( X\left( s\right) \!,s\right) \right\vert
480: \right) \right\} \label{grad Y-S}
481: \end{equation}%
482: where $C$ depends on the $y$-support of $f_{0}$ and $\mathbf{S}_{0},$ $%
483: t_{a}, $ and $t_{e}.$ \label{lemma4}
484: \end{lemma}
485:
486: \noindent \textbf{Proof.} Using the assumption (\ref{growthassumptions}) and
487: applying the Gronwall inequality to the ordinary differential equation (\ref%
488: {chareqn}) yields%
489: \begin{eqnarray*}
490: Y_{2}\left( \tau \right) &=&Y_{2}\left( 0\right) \exp \left( -\tau
491: /t_{a}\right) +\frac{1}{t_{a}}\int_{0}^{\tau }g\left( \mathbf{S}(X\left(
492: s\right) \!,s)\right) \exp [(s-\tau )/t_{a}]ds \\
493: &\leq &\left\vert Y_{2}\left( 0\right) \right\vert +\frac{1}{t_{a}^{2}}\Phi
494: \left( \sup_{0\leq s\leq \tau }\left\vert \mathbf{S}\left( X\left( s\right)
495: \!,s\right) \right\vert \right) .
496: \end{eqnarray*}%
497: In a similar way, we get%
498: \begin{eqnarray*}
499: Y_{1}\left( \tau \right) &=&Y_{1}\left( 0\right) \exp \left( -\tau
500: /t_{e}\right) +\frac{1}{t_{e}}\int_{0}^{\tau }\{g\left( \mathbf{S}(X\left(
501: s\right) \!,s)\right) -Y_{2}\left( s\right) \}\exp [(s-\tau )/t_{e}]ds \\
502: &\leq &\left\vert Y_{1}\left( 0\right) \right\vert +\frac{1}{t_{e}^{2}}%
503: \sup_{0\leq s\leq \tau }\left\vert Y_{2}\left( s\right) \right\vert +\frac{1%
504: }{t_{e}^{2}}\Phi \left( \sup_{0\leq s\leq \tau }\left\vert \mathbf{S}\left(
505: X\left( s\right) \!,s\right) \right\vert \right) .
506: \end{eqnarray*}%
507: Thus we deduce (\ref{grad Y-S}).
508:
509: \rightline{Q.E.D.}
510:
511: \begin{lemma}
512: \label{S-gradS} If $n\in L^{\infty }([0,\infty ):L^{1}\left( \mathbb{R}%
513: \right) \cap L^{2}\left( \mathbb{R}\right) ),$ then the solution $\mathbf{S}$
514: of the system of equations $(\ref{parabSsim})$ satisfies
515: \begin{equation*}
516: \left\Vert \mathbf{S}\left( t\right) \right\Vert _{L^{\infty }}\leq
517: C\sup_{0\leq \tau \leq t}\left\Vert n\left( t\right) \right\Vert
518: _{L^{1}}=C\left\Vert n\left( 0\right) \right\Vert _{L^{1}},
519: \end{equation*}%
520: \begin{equation*}
521: \left\Vert \frac{\partial \mathbf{S}}{\partial x}\left( t\right) \right\Vert
522: _{L^{\infty }}\leq C\left[ 1+\left\Vert n\left( 0\right) \right\Vert
523: _{L^{1}}\left( 1+\left( \ln t\right) _{+}+\left\vert \ln \left( \sup_{0\leq
524: \tau \leq t}\left\Vert n\left( \tau \right) \right\Vert _{L^{2}}\right)
525: \right\vert \right) \right]
526: \end{equation*}%
527: where $\left( \cdot \right) _{+}$ means the positive part and the constant $C
528: $ depends only on $k_{i}$, $k_{i}^{0}$ and $d_{i}.$ \label{lemma2}
529: \end{lemma}
530:
531: \noindent\textbf{Proof.} See \cite[Lemma 4]{Hwang:2005:GEC}.
532:
533: \rightline{Q.E.D.}
534:
535: \bigskip
536:
537: \noindent \textbf{Proof of Theorem \ref{theorem1}.} Integrating (\ref%
538: {hyperpsim}) along the characteristic (\ref{X}) -- (\ref{Y}) from $0$ to $t$
539: and using (\ref{growthassumptions}), we get
540: \begin{equation*}
541: f\left( x,v,\mathbf{y},t\right) \leq f_{0}\left( X\left( 0\right) ,v,\mathbf{%
542: Y}\left( 0\right) \right) +
543: \end{equation*}%
544: \begin{equation*}
545: +\;C\int_{0}^{t}\Psi \left( \left\vert \mathbf{Y}(\tau )\right\vert \right)
546: \times \left[ f\left( X(\tau ),v,\mathbf{Y}(\tau ),\tau \right)
547: +\int_{V}f\left( X(\tau ),v^{\prime },\mathbf{Y}(\tau ),\tau \right)
548: dv^{\prime }\right] \mbox{d}\tau +
549: \end{equation*}%
550: \begin{equation*}
551: +\int_{0}^{t}\left\vert \nabla _{\mathbf{y}}\cdot \mathbf{F}\left( \mathbf{S(%
552: }X\left( \tau \right) ),\mathbf{Y}\left( \tau \right) \right) \right\vert
553: f\left( X\left( \tau \right) ,v,\mathbf{Y}\left( \tau \right) ,\tau \right) %
554: \mbox{d}\tau .
555: \end{equation*}%
556: Since $\nabla _{y}\cdot \mathbf{F=-}\frac{1}{t_{e}}-\frac{1}{t_{a}}$, we get
557: (using Lemma \ref{lemma4})
558: \begin{equation}
559: f\left( x,v,\mathbf{y},t\right) \;\leq \;f_{0}\left( X\left( 0\right) ,v,%
560: \mathbf{Y}\left( 0\right) \right) +C\int_{0}^{t}f\left( X\left( \tau \right)
561: ,v,\mathbf{Y}\left( \tau \right) ,\tau \right) \mbox{d}\tau +
562: \label{ineqpom}
563: \end{equation}%
564: \begin{equation*}
565: +C\int_{0}^{t}\Psi \left( C\left[ 1+\Phi \left( \sup_{0\leq s\leq \tau
566: }\left\vert \mathbf{S}\left( X\left( s\right) ,s\right) \right\vert \right) %
567: \right] \right) \times
568: \end{equation*}%
569: \begin{equation*}
570: \times \left[ |V|f\left( X(\tau ),v,\mathbf{Y}(\tau ),\tau \right)
571: +\int_{V}f\left( X(\tau ),v^{\prime },\mathbf{Y}(\tau ),\tau \right)
572: dv^{\prime }\right] \mbox{d}\tau
573: \end{equation*}%
574: where $C$ is a constant depending only on support of $f_{0}$, $S_{0}$, $t_{e}
575: $ and $t_{a}.$ Using Lemma \ref{lemma3}, we have
576: \begin{equation}
577: \left( \det \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}\right) ^{-1}\leq
578: 1,\qquad \left( \det \frac{\partial X}{\partial x}\right) ^{-1}=1
579: \label{det}
580: \end{equation}%
581: and so%
582: \begin{equation*}
583: \int_{\mathbb{R}\times V\times \mathbb{R}^{M}}\int_{V}f^{p}\left( X\left(
584: \tau \right) ,v^{\prime },\mathbf{Y}\left( \tau \right) ,\tau \right)
585: dv^{\prime }\mbox{d}x\mbox{d}v\mbox{d}\mathbf{y}=
586: \end{equation*}%
587: \begin{equation*}
588: =\left\vert V\right\vert \int f^{p}\left( X\left( \tau \right) ,v^{\prime },%
589: \mathbf{Y}\left( \tau \right) ,\tau \right) \left( \det \frac{\partial
590: \mathbf{Y}}{\partial \mathbf{y}}\right) ^{-1}\left( \det \frac{\partial X}{%
591: \partial x}\right) ^{-1}dv^{\prime }\mbox{d}X\mbox{d}\mathbf{Y}\leq
592: \end{equation*}%
593: \begin{equation*}
594: \leq \left\vert V\right\vert \int f^{p}\left( X\left( \tau \right)
595: ,v^{\prime },\mathbf{Y}\left( \tau \right) ,\tau \right) \mbox{d}v^{\prime }%
596: \mbox{d}X\mbox{d}\mathbf{Y}.
597: \end{equation*}%
598: Taking the $p$-th power of (\ref{ineqpom}) and integrating over $x,$ $v,$
599: and $\mathbf{y}$ yields
600: \begin{equation}
601: \Vert f(t)\Vert \hbox{\raise -1mm \hbox{$_{L^{p}}$}}\leq \label{pom1}
602: \end{equation}%
603: \begin{equation*}
604: \leq \Vert f_{0}\Vert \hbox{\raise -1mm \hbox{$_{L^{p}}$}}+C\left\{ 1+\Psi
605: \left( C\left[ 1+\Phi \left( \sup_{0\leq s\leq t}\big\vert\mathbf{S}\left(
606: X\left( s\right) ,s\right) \big\vert\right) \right] \right) \right\} \times
607: \!\int_{0}^{t}\Vert f\left( \tau \right) \Vert
608: \hbox{\raise -1mm
609: \hbox{$_{L^{p}}$}}\mbox{d}\tau .
610: \end{equation*}%
611: Lemma \ref{S-gradS} implies
612: \begin{equation}
613: \sup_{0\leq \tau \leq t}\Vert \mathbf{S}\left( \mathbf{\cdot },\tau \right)
614: \Vert \hbox{\raise -1mm \hbox{$_{L^{\infty}}$}}\leq C\sup_{0\leq \tau \leq
615: t}\Vert n\left( t\right) \Vert \hbox{\raise -1mm
616: \hbox{$_{L^1}$}}=C\Vert n(0)\Vert \hbox{\raise
617: -1mm \hbox{$_{L^1}$}}\leq C\Vert f_{0}\Vert
618: \hbox{\raise -1mm
619: \hbox{$_{L^1}$}}. \label{pom2}
620: \end{equation}%
621: Consequently, using (\ref{pom1}) and (\ref{pom2}), we obtain
622: \begin{equation}
623: \Vert f(t)\Vert \hbox{\raise -1mm \hbox{$_{L^p}$}}\leq \Vert f_{0}\Vert %
624: \hbox{\raise -1mm \hbox{$_{L^{p}}$}}+C\Big\{1+\Psi \big(C[1+\Phi (C\Vert
625: f_{0}\Vert \hbox{\raise -1mm \hbox{$_{L^1}$}})]\big)\Big\}\times
626: \int_{0}^{t}\Vert f(\tau )\Vert \hbox{\raise -1mm \hbox{$_{L^p}$}}\mbox{d}%
627: \tau . \label{pom3}
628: \end{equation}%
629: Applying the Gronwall inequality, we obtain, for all $1\leq p\leq \infty ,$
630: \begin{equation}
631: \Vert f(t)\Vert \hbox{\raise -1mm \hbox{$_{L^p}$}}\leq C\left(
632: k_{i},k_{i}^{0},d_{i},t_{e},t_{a},\Vert f_{0}\Vert
633: \hbox{\raise -1mm
634: \hbox{$_{L^{p}}$}},\mathop{\mbox{supp}}f_{0},\Psi ,\Phi ,|V|\right) <\infty .
635: \label{pom4}
636: \end{equation}%
637: We now compute a priori estimates on derivatives of $f$. We differentiate (%
638: \ref{hyperpsim}) with respect of $x$, integrate along the characteristic (%
639: \ref{X}) -- (\ref{Y}) from $0$ to $t$ and use (\ref{growthassumptions}) to
640: get%
641: \begin{equation*}
642: \left\vert \frac{\partial f}{\partial x}\left( x,v,\mathbf{y},t\right)
643: \right\vert \;\;\leq \;\left\vert \frac{\partial f_{0}}{\partial x}\left(
644: X\left( 0\right) ,v,\mathbf{Y}\left( 0\right) \right) \right\vert
645: +\;C\int_{0}^{t}\Psi \left( \left\vert \mathbf{Y}(\tau )\right\vert \right)
646: \times
647: \end{equation*}%
648: \begin{equation*}
649: \times \left[ |V|\left\vert \frac{\partial f}{\partial x}\left( X(\tau ),v,%
650: \mathbf{Y}(\tau ),\tau \right) \right\vert +\int_{V}\left\vert \frac{%
651: \partial f}{\partial x}\left( X(\tau ),v^{\prime },\mathbf{Y}(\tau ),\tau
652: \right) \right\vert \mbox{d}v^{\prime }\right] \mbox{d}\tau +
653: \end{equation*}%
654: \begin{equation*}
655: +\int_{0}^{t}\big\vert\nabla _{\mathbf{y}}\cdot \mathbf{F}\left( \mathbf{S(}%
656: X\left( \tau \right) ),\mathbf{Y}\left( \tau \right) \right) \big\vert%
657: \left\vert \frac{\partial f}{\partial x}\left( X\left( \tau \right) ,v,%
658: \mathbf{Y}\left( \tau \right) ,\tau \right) \right\vert \mbox{d}\tau +
659: \end{equation*}%
660: \begin{equation*}
661: +\int_{0}^{t}\left\vert \frac{\partial g}{\partial \mathbf{z}}\left( \mathbf{%
662: S}(X\left( \tau \right) )\right) \right\vert \left\vert \frac{\partial
663: \mathbf{S}}{\partial x}\left( X\left( \tau \right) ,\tau \right) \right\vert
664: \left\vert \nabla _{\mathbf{y}}f\left( X\left( \tau \right) ,v,\mathbf{Y}%
665: \left( \tau \right) ,\tau \right) \right\vert \mbox{d}\tau .
666: \end{equation*}%
667: Similarly, differentiating (\ref{hyperpsim}) with respect of $y_{1}$ or $%
668: y_{2}$, integrating along the characteristic (\ref{X}) -- (\ref{Y}) from $0$
669: to $t$ and using (\ref{growthassumptions}), we obtain
670: \begin{equation*}
671: \left\vert \nabla_{\mathbf{y}}f\left( x,v,\mathbf{y},t\right) \right\vert
672: \;\;\leq \;\left\vert \nabla _{\mathbf{y}}f_{0}\left( X(0),v,\mathbf{Y}%
673: (0)\right) \right\vert +C\int_{0}^{t}\Psi \left( \left\vert \mathbf{Y(}%
674: X\left( \tau \right) ,\tau )\right\vert \right) \times \;
675: \end{equation*}%
676: \begin{equation*}
677: \times \Big\{|V|\big[\left\vert f\right\vert +\left\vert \nabla _{\mathbf{y}%
678: }f\right\vert \big]\left( X(\tau ),v,\mathbf{Y}(\tau ),\tau \right) +\int_{V}%
679: \big[\left\vert f\right\vert +\left\vert \nabla _{\mathbf{y}}f\right\vert %
680: \big]\left( X(\tau ),v^{\prime },\mathbf{Y}(\tau ),\tau \right) dv^{\prime }%
681: \Big\}\mbox{d}\tau +
682: \end{equation*}%
683: \begin{equation*}
684: +C\int_{0}^{t}\big\vert\nabla _{\mathbf{y}}\cdot \mathbf{F}\left( \mathbf{S(}%
685: X\left( \tau \right) ),\mathbf{Y}\left( \tau \right) \right) \big\vert%
686: \big\vert\nabla _{\mathbf{y}}f\left( X\left( \tau \right) ,v,\mathbf{Y}%
687: \left( \tau \right) ,\tau \right) \big\vert\mbox{d}\tau .
688: \end{equation*}%
689: If the interior of set $V$ is nonempty, we can also define the derivatives
690: of $f$ with respect of $v$ for any point in the interior of set $V.$
691: Differentiating (\ref{hyperpsim}) with respect of $v$ and integrating along
692: the characteristic (\ref{X}) -- (\ref{Y}) from $0$ to $t$, it implies
693: \begin{equation*}
694: \left\vert \frac{\partial f}{\partial v}\left( x,v,\mathbf{y},t\right)
695: \right\vert \;\;\leq \;\left\vert \frac{\partial f_{0}}{\partial v}\left(
696: X\left( 0\right) ,v,\mathbf{Y}\left( 0\right) \right) \right\vert
697: +\int_{0}^{t}\left\vert \frac{\partial f}{\partial x}\left( X\left( \tau
698: \right) ,v,\mathbf{Y}\left( \tau \right) ,\tau \right) \right\vert \mbox{d}%
699: \tau +
700: \end{equation*}%
701: \begin{equation*}
702: +\;C\int_{0}^{t}\Psi \left( \left\vert \mathbf{Y}(\tau )\right\vert \right)
703: \times
704: \bigg\{
705: |V|
706: \left\vert
707: f\left( X(\tau ),v,\mathbf{Y}(\tau),\tau \right)
708: \right\vert
709: +
710: \end{equation*}%
711: \begin{equation*}
712: +
713: |V|
714: \left\vert
715: \frac{\partial f}{\partial v}
716: \left( X(\tau ),v,\mathbf{Y}(\tau ),\tau \right)
717: \right\vert
718: +
719: \int_{V}
720: \left\vert
721: f
722: \left( X(\tau ),v^{\prime },\mathbf{Y}(\tau),\tau \right)
723: \right\vert
724: dv^{\prime }
725: \bigg\}\mbox{d}\tau
726: +
727: \end{equation*}%
728: \begin{equation*}
729: +\int_{0}^{t}\big\vert\nabla _{\mathbf{y}}\cdot \mathbf{F}\left( \mathbf{S(}%
730: X\left( \tau \right) ),\mathbf{Y}\left( \tau \right) \right) \big\vert%
731: \left\vert \frac{\partial f}{\partial v}\left( X\left( \tau \right) ,v,%
732: \mathbf{Y}\left( \tau \right) ,\tau \right) \right\vert \mbox{d}\tau .
733: \end{equation*}%
734: Using (\ref{pom4}), (\ref{det}), Lemma \ref{S-gradS} and Gronwall
735: inequality, we deduce%
736: \begin{equation}
737: \left\Vert \frac{\partial f}{\partial x}(t)\right\Vert _{L^{p}}+\left\Vert
738: \frac{\partial f}{\partial v}(t)\right\Vert _{L^{p}}+
739: \left\Vert
740: \nabla_{\mathbf{y}}f(t)
741: \right\Vert _{L^{p}}\leq
742: \label{derestimates}
743: \end{equation}%
744: \begin{equation*}
745: \leq C\left( k_{i},k_{i}^{0},d_{i},t_{e},t_{a},\Vert f_{0}\Vert
746: \hbox{\raise
747: -1mm \hbox{$_{W^{1,p}}$}},\Vert \mathbf{S}\left( 0\right) \Vert
748: \hbox{\raise
749: -1mm \hbox{$_{W^{1,p}}$}},\mathop{\mbox{supp}}f_{0},\mathop{\mbox{supp}}%
750: \mathbf{S}_{0},\Psi ,\Phi ,|V|\right) <\infty .
751: \end{equation*}%
752: Combining (\ref{pom4}) and (\ref{derestimates}), we obtain (\ref{spacef}).
753: Using Lemma \ref{S-gradS}, we get the estimate (\ref{spaceS}).
754:
755: \rightline{Q.E.D}
756:
757: \medskip
758:
759: \noindent \textbf{Remark.} Using Sobolev embedding theorems, we get global
760: existence of classical solutions provided that initial data are smooth.
761:
762: \section{Biological background}
763:
764: \label{secbiodetails}
765:
766: In order to study the general system (\ref{hyperp}) -- (\ref{parabS}), we
767: have to first specify realistic assumptions on the parameters of the model.
768: To this end, we summarize the relevant biological processes in Section \ref%
769: {secbache} and we extract the mathematical assumptions in Section \ref%
770: {secmajorassumptions}. These assumptions will be later used to prove the
771: global existence results in Section \ref{secglobalexistence}.
772:
773: \subsection{Bacterial chemotaxis}
774:
775: \label{secbache}
776:
777: As discussed before, the bacterial movement can be viewed as a biased random
778: walk. Bacterial motility is commonly provided by flagella, which are long,
779: spiral-shaped protein rods that stick out from the surface of the cell \cite%
780: {Salyers:2001:MDD}. The example of flagellated bacterium is the enteric
781: bacterium \textit{E.coli} which has 6-8 flagella. It has two modes of
782: behaviour based on counterclockwise and clockwise flagellar rotation. When
783: the flagella rotate counterclockwise (CCW), they all point in one direction
784: and consequently the cell moves forward in a straight \textquotedblleft
785: run". The speed of running is $s=10-20\mu \hbox{m/sec}.$ Clockwise (CW)
786: rotation of the flagella causes the flagella to point in different
787: directions, and the cell tumbles in place. Tumbling reorients the cell, so
788: that it can move in new direction when running starts again.
789:
790: For \textit{E.coli}, the duration of both runs and tumbles are exponentially
791: distributed with means of 1 sec and $10^{-1}$ sec respectively if an
792: extracellular chemical signal is not present \cite{Block:1983:AKB}. Under
793: the influence of an attractant, the cell increases its time in running in a
794: favourable direction -- see Figure \ref{figrw}.
795: \begin{figure}[tbp]
796: \centerline{\epsfxsize=4.5in\epsfbox{noincrattractant.eps}}
797: \caption{(a) \textit{A typical bacterial trajectory when no attractant is
798: present.} (b) \textit{Under the influence of an attractant, the cell
799: increases its time in running in a favourable direction.}}
800: \label{figrw}
801: \end{figure}
802: As the mean time for tumbling is ten times smaller than the mean time of
803: running, we can often neglect the time spent tumbling and we can model the
804: movement of the bacterium as a velocity jump process \cite%
805: {Othmer:1988:MDB,Erban:2004:ICB,Erban:2004:STS} as we already did in Section %
806: \ref{secintro}. It means that the bacterium runs in some direction and at
807: random instants of time it changes its direction with mean turning rate $%
808: \lambda(\mathbf{y}).$
809:
810: Since the bacteria move with more or less constant speed, the set $V$ of all
811: available velocities might be considered equal to $V=s \mathcal{S}^{N-1}$
812: where $\mathcal{S}^{N-1}$ is a unit sphere in $\mathbb{R}^{N}$ and $s$ is
813: the speed of the bacterium. Let us note that set $V=s \mathcal{S}^{N-1}$
814: satisfies the general condition (\ref{setvel}) (the presented theory works
815: for any set $V$ which satisfy (\ref{setvel})).
816:
817: The kernel $K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})$ gives the
818: probability of a change in velocity from $\mathbf{v}^{\prime}$ to $\mathbf{v}
819: $, given that a reorientation occurs. The simples possibility is to assume
820: that kernel is constant, i.e.
821: \begin{equation}
822: K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})=\frac{1}{|V|}.
823: \label{romundistr1}
824: \end{equation}%
825: This formula satisfies the normalization condition (\ref{normalcond}). The
826: underlying assumption behind (\ref{romundistr1}) is that (during the tumble)
827: bacterium simply choose a new direction randomly which is relatively a good
828: approximation for the bacterial chemotaxis, although there is also some bias
829: in the direction of the preceding run \cite{Berg:1972:CEC,Berg:1975:HBS}.
830: More realistically, one can assume that the turning kernel is a function of
831: the angle between new and old velocity, i.e.
832: \begin{equation}
833: K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})= k(\theta), \qquad \mbox{where}
834: \; \; \cos(\theta) = \frac{\mathbf{v} \cdot \mathbf{v^{\prime}}}{|\mathbf{v}%
835: | \, |\mathbf{v^{\prime}}|}. \label{romundistr2}
836: \end{equation}%
837: Whatever the choice of $K(\mathbf{v},\mathbf{v}^{\prime },\mathbf{y})$ is,
838: we may assume that it is bounded from above by a constant, i.e.
839: \begin{equation}
840: K(\mathbf{v},\mathbf{v}^{\prime},\mathbf{y})\le C \label{romundistr3}
841: \end{equation}%
842: where $C$ is independent of $\mathbf{v},$ $\mathbf{v}^{\prime}$ and $\mathbf{%
843: y}$. Next, we have to specify the choice of (\ref{rom14}) and the properties
844: of the turning frequency $\lambda(\mathbf{y}).$
845:
846: Chemotaxis is the process by which a cell alters its movement in response to
847: an extracellular chemical signal. From the microscopic (cell) point of view,
848: bacterial chemotaxis consists of several steps. First, the cell detects the
849: signal using its receptors. Then the signal information propagates through
850: the signal transduction biochemical network described by (\ref{rom14}). The
851: output of this network is a phosphorylated form of the protein CheY (denoted
852: CheY-P) which alters the motor behaviour of the flagellar motors and
853: consequently, the movement of the cell. CCW is the default state in the
854: absence of CheY-P, which binds to motor proteins and increases CW rotation.
855: Attractant binding to a receptor reduces the phosphorylation rate of CheY
856: and thereby increases the time spent in running state which constitutes the
857: fast response to a signal called \textit{excitation} of signal transduction
858: network. Another important aspect of signal transduction network is \textit{%
859: adaptation} which means that the response (probability per unit time of
860: CCW/CW rotation of flagella) returns to baseline levels on a time scale that
861: is slow compared to excitation, provided that there is no further change in
862: attractant concentration around the cell.
863:
864: A schematic of the signal transduction pathway is shown in Figure \ref%
865: {pathways} and it can be described as follows \cite%
866: {Spiro:1997:MEA,Stock:1996:C,Erban:2004:STS}.
867: \begin{figure}[th]
868: \centerline{\epsfxsize=5in\epsfbox{sigtranbac.eps}}
869: \caption{\textit{Excitation and adaptation in signal transduction pathway of
870: E. coli (from \protect\cite{Erban:2005:ICB}, with permission).}}
871: \label{pathways}
872: \end{figure}
873: Aspartate, the attractant most commonly-used in experiments (denoted S in
874: Figure \ref{pathways}), binds directly to the periplasmic domain of its
875: receptor, Tar. The cytoplasmic domain of Tar forms a stable complex with the
876: signaling proteins CheA and CheW (denoted A and W, respectively, in Figure %
877: \ref{pathways}), and the stability of this complex is not affected by ligand
878: binding \cite{Gegner:1992:AMR}. The signaling currency is in the form of
879: phosphoryl groups (-P), made available to the CheY (denoted Y in Figure \ref%
880: {pathways}) and CheB (not shown in Figure \ref{pathways}) through
881: autophosphorylation of CheA. Receptor complexes have two alternative
882: signaling states. In the attractant-bound form, the receptor inhibits CheA
883: autokinase activity; in the unliganded form, the receptor stimulates CheA
884: activity. Consequently, the response of the signal transduction network to a
885: step increase of the attractant concentration is as follows. First, the
886: attractant binding to a receptor reduces the autophosphorylation rate of
887: CheA. The level of phosphorylated CheA is thus lowered, causing less
888: phosphate to be transferred to CheY, yielding a lowered level of CheY-P. As
889: a result, tumbling is suppressed, and the cell's run length increases. This
890: constitutes the \textit{excitation} response of the system. Next slow
891: methylation and demethylation reactions begin to influence the response.
892: Ligand-bound receptors are more readily methylated than unliganded
893: receptors, and the lowered level of CheA-P causes a decrease in the level of
894: CheB-P, thereby reducing its demethylation activity. As a result, the
895: equilibrium of the system shifts in the direction of the higher methylation
896: states. The autophosphorylation rate of CheA is faster when the associated
897: Tar-CheA-CheW complex is in a higher methylation state, and so there is
898: finally a shift back toward the receptor states containing CheA-P. As a
899: result, CheY-P returns to its prestimulus level, and thus so does the CW
900: bias of the cell. This constitutes the \textit{adaptation} response. These
901: key steps, excitation via reduction in CheY-P, when a receptor is occupied,
902: and adaptation via methylation of the receptors, have been already
903: incorporated in the mathematical models of the bacterial signal transduction
904: \cite{Spiro:1997:MEA,Barkai:1997:RSB,Morton-Firth:1999:FEB}.
905:
906: Since the turning rate of bacterium is altered by CheY \cite{Cluzel:2000:UBM}%
907: , we can write $\lambda (\mathbf{y})\equiv \lambda (y_{1})$ where $y_{1}$
908: denotes the concentration of the phosphorylated form of CheY. Hence, the
909: individual-based model for bacterial chemotaxis is fully specified by the
910: equation (\ref{rom14}) which is integrated along the trajectory of each
911: cell, and by the $y_{1}$ component of the solution together with $\lambda
912: =\lambda (y_{1})$. The essential aspects of the dynamics which must be
913: captured by model (\ref{rom14}) are (i) it must exhibit excitation, which
914: here means a change in the turning frequency $\lambda (y_{1})$ in response
915: to a stimulus, (ii) the bias must return to baseline levels (i.e., the
916: response must adapt) on a time scale that is slow compared to excitation,
917: and (iii) the signal transduction network should amplify signals
918: appropriately \cite{Bourret:1991:STP,Segall:1986:TCB}. The mathematical
919: assumptions on (\ref{rom14}) and $\lambda (y_{1})$ are given in Section \ref%
920: {secmajorassumptions}.
921:
922: \subsection{Mathematical assumptions on the signal transduction network}
923:
924: \label{secmajorassumptions}
925:
926: The mathematical model of the signal transduction network (\ref{rom14})
927: can be rewritten in the following form
928: \begin{equation}
929: \frac{\mbox{d}\mathbf{y}}{\mbox{d}t}=\mathbf{F}(\mathbf{C}(t),\mathbf{y}%
930: )\qquad \mbox{where}\qquad \mathbf{C}(t)=\mathbf{S}(\mathbf{x}(t),t).
931: \label{rom14new}
932: \end{equation}%
933: The vector function $\mathbf{C}(t)$ gives signal values which are seen by a
934: cell along its trajectory. Time evolution of $\mathbf{y}$ in equation (\ref%
935: {rom14new}) is controlled by the input time dependent vector $\mathbf{C}(t).$
936: Therefore, it is natural to describe the behaviour of $\mathbf{F}$ in terms
937: of the input function $\mathbf{C}(t).$
938:
939: The mathematical formulation of the \textit{adaptation} property of the
940: signal transduction network (\ref{rom14}) can be written in the following
941: form. There exists a universal constant $\overline{y}_{1}$ such that for any
942: constant signal along the trajectory $\mathbf{C}_{0}$, i.e. $\mathbf{C}%
943: (t)\equiv \mathbf{C}_{0}=\mbox{{\rm const}}$ and for any initial condition $%
944: \mathbf{y}(0)=\mathbf{y}_{0}$, the solution of the system $(\ref{rom14})$
945: satisfies
946: \begin{equation}
947: \lim_{t\rightarrow \infty }y_{1}(t)=\overline{y}_{1}. \label{adaptproperty}
948: \end{equation}%
949: Formula (\ref{adaptproperty}) describes the perfect adaptation. From the
950: application point of view, it is desirable that the signal transduction
951: model satisfies (at least approximately) the adaptation property for a
952: reasonably large set of signals. However, the existence theorems presented
953: in Section \ref{secglobalexistence} do not require perfect adaptation and we
954: will prove the existence of solutions even for models which do not satisfy (%
955: \ref{adaptproperty}). It is worthwhile to note that the simplified model of
956: excitation-adaptation dynamics (\ref{simmodel}) from Section \ref{secmotiv}
957: satisfied adaptation property (\ref{adaptproperty}). In fact, $%
958: y_{1}(t)\rightarrow 0$ as $t\rightarrow \infty $ for any constant signal,
959: i.e. $y_{1}$ adapts perfectly to any constant stimulus. Moreover, model (\ref%
960: {simmodel}) describes the excitation-adaptation dynamics as discussed in
961: Section \ref{secbache} provided that we choose $t_{e}<t_{a}$. Here, the time
962: constants $t_{e}$ and $t_{a}$ are labeled in anticipation of using $y_{1}$
963: for the internal response, and $y_{2}$ as the adaptation variable, and
964: therefore we call $t_{e}$ and $t_{a}$ the excitation and adaptation time
965: constant, respectively \cite{Erban:2004:ICB}.
966:
967: In order to model the random walk of the individual bacterium, we must have
968: a good understanding of the dependence of the (output) turning rate $\lambda
969: (y_{1})$ on the (input) signal function $\mathbf{C}(t).$ If the input signal
970: function is constant then the behaviour of $\lambda (y_{1})$ follows the
971: adaptation property. On the other hand, time dependent input $\mathbf{C}(t)$
972: can introduce large variations in $\lambda (y_{1}).$ The time derivative of $%
973: \mathbf{C}(t),$ i.e. the time derivative of the signal seen by a cell, is
974: equal to
975: \begin{equation}
976: \frac{\mbox{d}\mathbf{C}}{\mbox{d}t}=\mathbf{v}\cdot \frac{\partial \mathbf{S%
977: }}{\partial \mathbf{x}}+\frac{\partial \mathbf{S}}{\partial t}.
978: \label{timdercell}
979: \end{equation}%
980: To see what type of conditions on the turning rate $\lambda$
981: are reasonable, let us consider the time
982: independent signal (attractant) with a maximum at the point $x_{m}$ as it is
983: schematically shown in one dimension in Figure \ref{figexpblow} (panel in
984: the middle).
985: \begin{figure}[th]
986: \centerline{\epsfxsize=4.8in\epsfbox{explainblowup.eps}}
987: \caption{\textit{Schematic of behaviour of hypothetical cells which
988: \textquotedblleft perfectly avoid going in wrong directions" (panel on the
989: left) and hypothetical cells which \textquotedblleft perfectly follow good
990: directions" (panel on the right). Details are explained in the text. }}
991: \label{figexpblow}
992: \end{figure}
993: We consider that bacteria move with the fixed speed either to the right or
994: left and we discuss the following two simple cases of dependence of output $%
995: \lambda (y_{1})$ on input $\mathbf{C}(t)$.
996: \begin{equation*}
997: \mbox{(a)}\;\;\lambda (y_{1})=\left\{
998: \begin{array}{ll}
999: 1\quad & \mbox{for}\;\mbox{d}C/\mbox{d}t\geq 0; \\
1000: \infty \quad & \mbox{for}\;\mbox{d}C/\mbox{d}t<0;%
1001: \end{array}%
1002: \right. \qquad \mbox{(b)}\;\;\lambda (y_{1})=\left\{
1003: \begin{array}{ll}
1004: 0\quad & \mbox{for}\;\mbox{d}C/\mbox{d}t\geq 0; \\
1005: 1\quad & \mbox{for}\;\mbox{d}C/\mbox{d}t<0.%
1006: \end{array}%
1007: \right.
1008: \end{equation*}%
1009: Let us note that cases (a) and (b) are considered as definitions of the
1010: input-output behaviour in two extreme cases (these definitions are not
1011: connected with any underlying differential equation in this example).
1012:
1013: First, suppose that a bacterium is at the position $x<x_{m}$. If we use
1014: input-output behaviour (a), then the cell goes to the right. It sometimes
1015: \textquotedblleft turns" to the left but it instantly turns back. So, the
1016: cell spends all the time going to the right, and case (a) is an example of
1017: the individual-based model where cells perfectly avoid going in wrong
1018: directions. If we use input-output behaviour (b), then the right going cells
1019: never turn (for $x<x_{m}$). Hence, case (b) is an example of the
1020: individual-based model
1021: where cells perfectly follow good directions. Both cases (a) and (b)
1022: describe the simple transport of bacteria for $x<x_{m}$. The difference of
1023: these models is when cells reach the maximum of the signal $x_{m}$. In case
1024: (a), cells instantly turn back. It means that the final positions of all
1025: bacteria are equal to $x_{m}$ and a Dirac-like distribution is created in
1026: finite time (see Figure \ref{figexpblow}, panel on the left). In case (b),
1027: cells continue movement to the region $x>x_{m}$ and the final distribution
1028: profile is smooth, as shown schematically in Figure \ref{figexpblow} (panel
1029: on the right).
1030:
1031: The previous simple example shows that singularities might develop if the
1032: turning rate is too large (without a reasonable control by the signal
1033: change), as in case (a) where cells perfectly avoid going in wrong
1034: directions. This observation suggests for growth conditions on $\lambda
1035: (y_{1})$ from above which prevent formation of singularities. The necessary
1036: conditions on the turning frequency $\lambda (y_{1})$ is $\lambda
1037: (y_{1})\geq 0$ and our heuristic conclusions can be incorporated to the
1038: following growth estimate
1039: \begin{equation}
1040: \lambda (y_{1})\leq C\left( 1+\Lambda \left( \left\vert \mathbf{C}%
1041: \right\vert \right) +\left\vert \frac{\mbox{d}\mathbf{C}}{\mbox{d}t}%
1042: \right\vert \right) , \label{growthlambda}
1043: \end{equation}%
1044: where $\Lambda \left( \cdot \right) \in C\left( \mathbb{R}\right) $ is a
1045: non-negative, nondecreasing continuous function. The verification of growth
1046: estimate (\ref{growthlambda}) depends on the particular form of $\mathbf{F}%
1047: (\cdot ,\cdot )$ and $\lambda (\cdot ).$ For example, if (\ref{rom14}) and $%
1048: \lambda (\cdot )$ satisfy
1049: \begin{equation}
1050: |y_{1}|\leq C_{1}\left( 1+\left\vert \frac{\mbox{d}\mathbf{C}}{\mbox{d}t}%
1051: \right\vert ^{\omega }\right) ,\qquad \lambda (y_{1})\leq
1052: C_{2}(1+|y_{1}|^{\sigma }),\qquad \omega \sigma \leq 1, \label{growthylam}
1053: \end{equation}%
1054: then (\ref{growthlambda}) follows. There are several other conditions on $%
1055: \mathbf{F}(\cdot ,\cdot )$ and $\lambda (\cdot )$ which also guarantee
1056: growth estimate (\ref{growthlambda}). Hence, we do not formulate our growth
1057: estimates in terms of $\mathbf{F}(\cdot ,\cdot )$ and $\lambda (\cdot )$,
1058: but we simply assume (\ref{growthlambda}) directly in our existence
1059: theorems. Using formula (\ref{kernelT}), we can formulate the estimate (\ref%
1060: {growthlambda}) also in terms of the kernel $T(\mathbf{v},\mathbf{v}^{\prime
1061: },\mathbf{y}).$
1062:
1063: Using estimate (\ref{romundistr3}) and definition (\ref{kernelT}), we can
1064: write the growth assumption on $T$ in the following form
1065: \begin{equation}
1066: T(v,v^{\prime },\mathbf{y}) \leq C |\lambda (y_{1})|. \label{growthT}
1067: \end{equation}%
1068: We also have to assume a growth assumption of ${\nabla }_{\mathbf{y}}{\cdot }%
1069: \mathbf{F}$. In Theorem \ref{theorem2}, we assume that there exists a
1070: non-negative, nondecreasing continuous function $\Pi \left( \cdot \right)
1071: \in C\left( \mathbb{R}\right) $ satisfying%
1072: \begin{equation}
1073: \left\vert {\nabla }_{\mathbf{y}}{\cdot }\mathbf{F}\left( \mathbf{z,y}%
1074: \right) \right\vert \leq C\left( 1+\Pi \left( \left\vert \mathbf{z}%
1075: \right\vert \right) \right). \label{divF2}
1076: \end{equation}
1077: Notice that our simple model (\ref{simmodel}) satisfies (\ref{divF2}). A
1078: different condition on ${\nabla }_{\mathbf{y}}{\cdot }\mathbf{F}$ is studied
1079: also in Corollary \ref{cor1}.
1080:
1081: \subsection{Mathematical assumptions on the dynamics of the extracellular
1082: signals}
1083:
1084: \label{secmajorassumptionssignal}
1085:
1086: Various forms of $\mathbf{R}(\mathbf{S},n)$ can be considered. The simplest
1087: case from the mathematical point of view is when the extracellular signals
1088: are nutrients which are consumed by cells, i.e.
1089: \begin{equation}
1090: \mathbf{R}(\mathbf{S},n)= - \mathbf{K} \mathbf{S} n \label{eatonly}
1091: \end{equation}
1092: where $\mathbf{K}$ is a diagonal nonnegative $M \times M$ matrix (with rate
1093: constants on the diagonal). One can also assume that the cells produce
1094: signals which are degraded at some rate, i.e.
1095: \begin{equation}
1096: \mathbf{R}(\mathbf{S},n)= n [k_1,k_{2},\dots ,k_{M}]^{T} - \mathbf{K}
1097: \mathbf{S} \label{proddegr}
1098: \end{equation}
1099: where $k_1,$ \dots, $k_M$ are rates of production of the different
1100: components of the signal and $\mathbf{K}$ is a diagonal nonnegative $M
1101: \times M$ matrix. If we allow the nondiagonal terms in matrix $\mathbf{K}$,
1102: then the extracellular coupling of the signals (e.g. reactions between signals)
1103: is added to the model. One can also consider that some signals can be
1104: produced by cells and some signals can be degraded by cells, i.e.
1105: effectively combining (\ref{eatonly}) and (\ref{proddegr}). Moreover, we can
1106: also assume that some signals can be attractants while other signals can be
1107: repellents etc.
1108:
1109: Depending on the model system, there are many possibilities to specify the
1110: dynamics of the extracellular signal. In what follows, we use (\ref{proddegr}%
1111: ). However, it is possible to modify and prove the following existence
1112: theorems using different evolution equations for the extracellular signal
1113: too. The only requirement is that the evolution equation for the
1114: extracellular signal must satisfy suitable growth estimates similar to the
1115: estimates which are proven in Lemma \ref{gradS} for (\ref{proddegr}).
1116:
1117: \section{Global existence for the general signal transduction models}
1118:
1119: \label{secglobalexistence}
1120:
1121: In this section, we prove global existence results using the framework of
1122: Sections \ref{secmajorassumptions} and \ref{secmajorassumptionssignal}. We
1123: will work in one-dimensional physical space, i.e. $N=1$ and we first assume
1124: the case of elliptic equations for the extracellular signals. Hence, system
1125: of equations (\ref{hyperp}) -- (\ref{parabS}) reads as follows
1126: \begin{equation}
1127: \frac{\partial f}{\partial t}+\nabla_{\!x}\cdot \mathbf{v} f + \nabla _{\!%
1128: \mathbf{y}} \cdot \mathbf{F}(\mathbf{S}(\mathbf{x}),\mathbf{y}) f = \int_{V}
1129: T(v,v^{\prime},\mathbf{y}) \Big[ f(v^{\prime}) - f(v) \Big] \mbox{d}%
1130: v^{\prime}, \label{hyperpell}
1131: \end{equation}
1132: \begin{equation}
1133: d_{i}\frac{\partial ^{2}S_{i}}{\partial x^{2}}+k_{i}n-k_{i}^{0}S_{i} = 0,
1134: \qquad \;i=1,\dots ,M, \label{ellipticS}
1135: \end{equation}
1136: where $d_{i}$, $k_{i}$ and $k_{i}^{0}$ are positive constants and $n\equiv
1137: n(x,t)$ is the macroscopic density of individuals at point $x \in \mathbb{R}$
1138: and time $t$ given by (\ref{equationforn}). Now, we can formulate the
1139: existence theorem.
1140:
1141: \begin{theorem}
1142: \label{theorem2} Let us assume $(\ref{growthlambda})$, $(\ref{growthT})$ and
1143: $(\ref{divF2})$. Assume that $f_{0}\in L^{1}\cap L^{\infty }({}\mathbb{R}%
1144: \times V\times \mathbb{R}^{m})$ and let initial condition $S_{0}\in \lbrack
1145: W^{2,p}({}\mathbb{R})]^{M}$ satisfies $(\ref{ellipticS})$. Then there exists
1146: a global solution of system $(\ref{hyperpell})$ -- $(\ref{ellipticS})$
1147: satisfying, for all $t\geq 0$
1148: \begin{equation}
1149: f(\cdot ,\cdot ,\cdot ,t)\in L^{1}\cap L^{\infty }({}\mathbb{R}\times
1150: V\times \mathbb{R}^{m}), \label{fpestimate}
1151: \end{equation}%
1152: \begin{equation}
1153: \mathbf{S}(\cdot ,t)\in \lbrack W^{2,p}(\mathbb{R})]^{M},\ \ \ \ \ \text{for
1154: all }1\leq p<+\infty, \label{Spestimate}
1155: \end{equation}%
1156: and initial conditions $f(\cdot ,\cdot ,\cdot ,0)=f_{0}(\cdot ,\cdot ,\cdot
1157: ) $ and $\mathbf{S}(\cdot ,0)=\mathbf{S}_{0}(\cdot )$.
1158: \end{theorem}
1159:
1160: \medskip
1161:
1162: \noindent \textbf{Remark.} To avoid technicalities, we focus in Theorem \ref%
1163: {theorem2} only on $L^{p}$ estimates of $f$. The results could be extended
1164: to $W^{k,p}$ estimates under suitable growth assumptions on derivatives of $%
1165: T\left( v,v^{\prime },y\right)$ and $\mathbf{F}$.
1166:
1167: \medskip
1168:
1169: \noindent In order to prove Theorem \ref{theorem2}, we formulate some
1170: auxiliary lemmas. We start with the generalization of the Gronwall
1171: inequality.
1172:
1173: \begin{lemma}
1174: \label{gronwall}Let $a\left( s\right) $ and $b\left( s\right) $ be positive
1175: integrable functions on $[0,t]$. Let $w\left( t\right) $ be positive and
1176: differentiable in $t,$ and satisfy
1177: \begin{equation*}
1178: w^{\prime }\leq a\left( t\right) w\ln w+b\left( t\right) w.
1179: \end{equation*}%
1180: Then
1181: \begin{equation*}
1182: w\left( t\right) \leq \left[ w\left( 0\right) \exp \left(
1183: \int_{0}^{t}b\left( s\right) e^{-\int_{0}^{s}a\left( \tau \right) d\tau
1184: }ds\right) \right] ^{\exp \left( \int_{0}^{t}a\left( s\right) ds\right) }.
1185: \end{equation*}
1186: \end{lemma}
1187:
1188: \noindent \textbf{Proof. } See \cite[Lemma 4]{Hwang:2005:GEC}.
1189:
1190: \rightline{Q.E.D.}
1191:
1192: \noindent The characteristics of the hyperbolic equation (\ref{hyperpell})
1193: are given for $N=1$ as the solution of (\ref{chareqn}). The back-in-time
1194: characteristics starting at $\left( x,v,\mathbf{y},t\right) $ are given as
1195: \begin{eqnarray}
1196: X\left( s;x,v\mathbf{,y},t\right) &=&x-v\left( t-s\right) ,\ \ \label{X2}
1197: \\
1198: \mathbf{Y}\left( s;x,v\mathbf{,y},t\right) &=&\mathbf{y}-\int_{s}^{t}%
1199: \mathbf{F}\left( \mathbf{S}\left( X(\tau ),\tau \right) ,\mathbf{Y}(\tau
1200: )\right) \mbox{d}\tau . \label{Y2}
1201: \end{eqnarray}%
1202: The generalization of Lemma \ref{lemma3} is given as the following Lemma.
1203:
1204: \begin{lemma}
1205: \label{detGeneral} Derivation of the characteristics $(\ref{X2})$ and $(\ref%
1206: {Y2})$ with respect to the initial conditions gives
1207: \begin{equation}
1208: \frac{\partial X}{\partial x}=1 \qquad \text{and} \qquad \frac{\partial
1209: \mathbf{Y}}{\partial \mathbf{y}} = \exp \left[ -\int_{s}^{t}\frac{\partial
1210: \mathbf{F}}{\partial \mathbf{y}} \left( \mathbf{S}\left( X(\tau ),\tau
1211: \right) ,\mathbf{Y}(\tau)\right) \mbox{d}\tau \right].
1212: \label{dercharGeneral}
1213: \end{equation}%
1214: Moreover,
1215: \begin{equation}
1216: \det \frac{\partial \mathbf{Y}}{\partial \mathbf{y}} = \exp \left[
1217: -\int_{s}^{t}\nabla _{\mathbf{y}}\cdot \mathbf{F} \left( \mathbf{S}\left(
1218: X(\tau ),\tau \right) ,\mathbf{Y}(\tau)\right) \mbox{d}\tau \right].
1219: \label{determyyGeneral}
1220: \end{equation}
1221: \end{lemma}
1222:
1223: \noindent \textbf{Proof.} We differentiate (\ref{Y2}) with respect to $%
1224: \mathbf{y}$ to get%
1225: \begin{equation}
1226: \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}=\mathbf{I}_{m}+\int_{t}^{s}%
1227: \frac{\partial \mathbf{F}}{\partial \mathbf{y}}\left( \mathbf{S}\left(
1228: X(\tau ),\tau \right) ,\mathbf{Y}(\tau )\right) \frac{\partial \mathbf{Y}}{%
1229: \partial \mathbf{y}}\left( \tau \right) \mbox{d}\tau . \label{derY2}
1230: \end{equation}%
1231: where $\mathbf{I}_{m}$ is the $m\times m$ identity matrix. Let
1232: \begin{equation*}
1233: \mathbf{G}\left( s\right) =\int_{t}^{s}\frac{\partial \mathbf{F}}{\partial
1234: \mathbf{y}}\left( \mathbf{S}\left( X(\tau ),\tau \right) ,\mathbf{Y}(\tau
1235: )\right) \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}\left( \tau \right) %
1236: \mbox{d}\tau ,
1237: \end{equation*}%
1238: then we have%
1239: \begin{equation*}
1240: \mathbf{G}^{\prime }\left( s\right) -\mathbf{G}\left( s\right) \frac{%
1241: \partial \mathbf{F}}{\partial \mathbf{y}}\left( \mathbf{S}\left( X(\tau
1242: ),\tau \right) ,\mathbf{Y}(\tau )\right) =\frac{\partial \mathbf{F}}{%
1243: \partial \mathbf{y}}\left( \mathbf{S}\left( X(\tau ),\tau \right) ,\mathbf{Y}%
1244: (\tau )\right) .
1245: \end{equation*}%
1246: Integrating the last equation, we obtain (\ref{dercharGeneral}). Since the
1247: determinant of the exponential of the matrix is the exponential of the trace
1248: of the matrix, we have
1249: \begin{equation*}
1250: \det \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}=\exp \left[ %
1251: \mathop{\mbox{{\rm trace}}}\left( -\int_{s}^{t}\frac{\partial \mathbf{F}}{%
1252: \partial \mathbf{y}}\left( \mathbf{S}\left( X(\tau ),\tau \right) ,\mathbf{Y}%
1253: (\tau )\right) \mbox{d}\tau \right) \right] =
1254: \end{equation*}%
1255: \begin{equation*}
1256: =\exp \left[ -\int_{s}^{t}\nabla _{\mathbf{y}}\cdot \mathbf{F}\left( \mathbf{%
1257: S}\left( X(\tau ),\tau \right) ,\mathbf{Y}(\tau )\right) \mbox{d}\tau \right]
1258: .
1259: \end{equation*}%
1260: Hence, we have proved (\ref{determyyGeneral}).
1261:
1262: \rightline{Q.E.D.}
1263:
1264: \medskip
1265:
1266: \noindent Next, we present the growth estimates on the extracellular signal $%
1267: \mathbf{S}$ and on its derivatives. The time and space derivatives of the
1268: signal vector $\mathbf{S}$ are controlled by logarithm of the $L^{2}$-norm
1269: of the cell density. Note that the analogous result was also shown in \cite[%
1270: Lemma 4]{Hwang:2005:GEC} for the parabolic equation for the extracellular
1271: signal. The difference between \cite[Lemma 4]{Hwang:2005:GEC} and Lemma \ref%
1272: {gradS} is that we prove also the estimate on the time derivative as well as
1273: the estimate on the space derivative of the signal.
1274:
1275: \begin{lemma}
1276: \label{gradS}If $n\in L^{\infty }([0,\infty ):L^{1}\left( \mathbb{R}\right)
1277: \cap L^{2}\left( \mathbb{R}\right) ),$ then the solution $\mathbf{S}$ in $(%
1278: \ref{ellipticS})$ satisfies
1279: \begin{equation*}
1280: \left\Vert \mathbf{S}\left( t\right) \right\Vert _{L^{\infty }}\leq
1281: C\left\Vert n\left( t\right) \right\Vert _{L^{1}}=C\left\Vert n\left(
1282: 0\right) \right\Vert _{L^{1}},
1283: \end{equation*}%
1284: \begin{equation}
1285: \left\Vert \frac{\partial \mathbf{S}}{\partial x}\left( t\right) \right\Vert
1286: _{L^{\infty }}\leq C\left[ 1+\left\Vert n\left( 0\right) \right\Vert
1287: _{L^{1}}\left\{ 1+\ln \left( \left\Vert n(t) \right\Vert _{L^{2}}+1\right)
1288: \right\} \right] , \label{sxderestim}
1289: \end{equation}%
1290: \begin{equation}
1291: \left\Vert \frac{\partial \mathbf{S}}{\partial t}\left( t\right) \right\Vert
1292: _{L^{\infty }}\leq C\left[ 1+\left\Vert n\left( 0\right) \right\Vert
1293: _{L^{1}}\left\{ 1+\ln \left( \left\Vert n(t) \right\Vert _{L^{2}}+1\right)
1294: \right\} \right] . \label{stderestim}
1295: \end{equation}%
1296: where the constant $C$ depends only on $k_{i},$ $k_{i}^{0}$, $d_{i}$ and $V$.
1297: \end{lemma}
1298:
1299: \noindent \textbf{Proof.} Let $1\leq i\leq M.$ Taking the Fourier transform
1300: of $(\ref{ellipticS})$ in the $x-$variable, we obtain
1301: \begin{equation*}
1302: \hat{S}_{i}\left( \xi ,t\right) =\frac{k_{i}}{d_{i}}\frac{\hat{n}\left( \xi
1303: ,t\right) }{\xi ^{2}+k_{i}^{0}/d_{i}}.
1304: \end{equation*}%
1305: Thus we have%
1306: \begin{eqnarray*}
1307: \left\Vert S_{i}(t)\right\Vert _{L^{\infty }} &\leq &\left\Vert \hat{S}%
1308: _{i}(t)\right\Vert _{L^{1}}\leq \frac{k_{i}}{d_{i}}\left\Vert \hat{n}%
1309: (t)\right\Vert _{L^{\infty }}\int_{-\infty }^{\infty }\frac{1}{\xi
1310: ^{2}+k_{i}^{0}/d_{i}}\mbox{d}\xi \\
1311: &\leq &C\left( \frac{k_{i}}{d_{i}},\frac{k_{i}^{0}}{d_{i}}\right) \left\Vert
1312: n(t)\right\Vert _{L^{1}}=C\left( \frac{k_{i}}{d_{i}},\frac{k_{i}^{0}}{d_{i}}%
1313: \right) \left\Vert n(0)\right\Vert _{L^{1}}.
1314: \end{eqnarray*}%
1315: Next we estimate the $x-$derivative of the signal as follows.%
1316: \begin{equation*}
1317: \left\Vert \frac{\partial S_{i}}{\partial x}\left( t\right) \right\Vert
1318: _{L^{\infty }}\leq \left\Vert \xi \hat{S}_{i}(t)\right\Vert _{L^{1}}\leq
1319: \frac{k_{i}}{d_{i}}\int_{-\infty }^{\infty }\frac{\left\vert \xi \right\vert
1320: \left\vert \hat{n}\left( \xi ,t\right) \right\vert }{\xi ^{2}+k_{i}^{0}/d_{i}%
1321: }\mbox{d}\xi =\frac{k_{i}}{d_{i}}\left\{ I_{1}+I_{2}\right\} ,
1322: \end{equation*}%
1323: \begin{equation*}
1324: \mbox{where}\quad I_{1}=\int_{\left\vert \xi \right\vert \leq \Vert
1325: n(t)\Vert _{L^{2}}^{2}}\frac{\left\vert \xi \right\vert \left\vert \hat{n}%
1326: \left( \xi ,t\right) \right\vert }{\xi ^{2}+k_{i}^{0}/d_{i}}\mbox{d}\xi
1327: \quad \mbox{and}\quad I_{2}=\int_{\left\vert \xi \right\vert \geq \Vert
1328: n(t)\Vert _{L^{2}}^{2}}\frac{\left\vert \xi \right\vert \left\vert \hat{n}%
1329: \left( \xi ,t\right) \right\vert }{\xi ^{2}+k_{i}^{0}/d_{i}}\mbox{d}\xi .
1330: \end{equation*}%
1331: First, we estimate the integral $I_{1}$. We obtain%
1332: \begin{equation*}
1333: I_{1}\leq \left\Vert \hat{n}(t)\right\Vert _{L^{\infty }}\int_{\left\vert
1334: \xi \right\vert \leq \Vert n(t)\Vert _{L^{2}}^{2}}\frac{\left\vert \xi
1335: \right\vert }{\xi ^{2}+k_{i}^{0}/d_{i}}\mbox{d}\xi =
1336: \end{equation*}%
1337: \begin{equation*}
1338: =\left\Vert \hat{n}(t)\right\Vert _{L^{\infty }}\ln \left( \frac{\Vert
1339: n(t)\Vert _{L^{2}}^{4}}{k_{i}^{0}/d_{i}}+1\right) \leq \left\Vert
1340: n(t)\right\Vert _{L^{1}}\ln \left( \frac{\Vert n(t)\Vert _{L^{2}}^{4}}{%
1341: k_{i}^{0}/d_{i}}+1\right)
1342: \end{equation*}%
1343: We use H\"{o}lder's inequality with $p=q=2$ to estimate $I_{2}$ as%
1344: \begin{eqnarray*}
1345: I_{2} &\leq &\left\Vert \hat{n}(t)\right\Vert _{L^{2}}\left(
1346: \int_{\left\vert \xi \right\vert \geq \Vert n(t)\Vert _{L^{2}}^{2}}\left(
1347: \frac{\xi }{\xi ^{2}+k_{i}^{0}/d_{i}}\right) ^{2}\mbox{d}\xi \right) ^{1/2}
1348: \\
1349: &\leq &\left\Vert n(t)\right\Vert _{L^{2}}\left( \int_{\left\vert \xi
1350: \right\vert \geq \Vert n(t)\Vert _{L^{2}}^{2}}\xi ^{-2}\mbox{d}\xi \right)
1351: ^{1/2}\leq \sqrt{2}.
1352: \end{eqnarray*}%
1353: By combining the estimates for $I_{1}$ and $I_{2}$, we obtain (\ref%
1354: {sxderestim}). In order to estimate the time derivative of the extracellular
1355: signal, we take the time derivative of $(\ref{ellipticS})$ and apply the
1356: Fourier transform in the $x-$variable to get%
1357: \begin{equation*}
1358: \frac{\partial \hat{S}_{i}}{\partial t}\left( \xi ,t\right) =\frac{k_{i}}{%
1359: d_{i}}\frac{\partial \hat{n}}{\partial t}\left( \xi ,t\right) \frac{1}{\xi
1360: ^{2}+k_{i}^{0}/d_{i}}.
1361: \end{equation*}%
1362: By integrating $(\ref{hyperpell})$ over $v$ and $y,$ we get%
1363: \begin{equation*}
1364: \frac{\partial n}{\partial t}=-\frac{\partial j}{\partial x}\quad %
1365: \mbox{where}\quad j\left( x,t\right) =\iint_{V\mathbb{\times R}^{m}}vf\left(
1366: x,v,\mathbf{y},t\right) \mbox{d}v\mbox{d}\mathbf{y}.
1367: \end{equation*}%
1368: Thus we have%
1369: \begin{equation*}
1370: \frac{\partial \hat{S}_{i}}{\partial t}\left( \xi ,t\right) =\frac{k_{i}}{%
1371: d_{i}}\frac{-i\xi \hat{\jmath}\left( \xi ,t\right) }{\xi ^{2}+k_{i}^{0}/d_{i}%
1372: }.
1373: \end{equation*}%
1374: Then we have%
1375: \begin{equation*}
1376: \left\Vert \frac{\partial S_{i}}{\partial t}(t)\right\Vert _{L^{\infty
1377: }}\leq \left\Vert \frac{\partial \hat{S}_{i}}{\partial t}(t)\right\Vert
1378: _{L^{1}}\leq \frac{k_{i}}{d_{i}}\int_{-\infty }^{\infty }\frac{\left\vert
1379: \xi \right\vert \left\vert \hat{\jmath}\left( \xi \right) \right\vert }{\xi
1380: ^{2}+k_{i}^{0}/d_{i}}\mbox{d}\xi .
1381: \end{equation*}%
1382: Notice that%
1383: \begin{equation*}
1384: \left\Vert \hat{\jmath}(t)\right\Vert _{L^{\infty }}\leq \left\Vert
1385: j(t)\right\Vert _{L^{1}}\leq \iiint_{\mathbb{R\times }V\mathbb{\times R}%
1386: ^{m}}\left\vert v\right\vert f\left( x,v,\mathbf{y},t\right) \mbox{d}x%
1387: \mbox{d}v\mbox{d}\mathbf{y}\leq
1388: \end{equation*}%
1389: \begin{equation*}
1390: \leq C\left( V\right) \left\Vert n\left( t\right) \right\Vert
1391: _{L^{1}}=C\left( V\right) \left\Vert n(0)\right\Vert _{L^{1}},
1392: \end{equation*}%
1393: \begin{equation*}
1394: \left\Vert \hat{\jmath}(t)\right\Vert _{L^{2}}=\left\Vert j(t)\right\Vert
1395: _{L^{2}}\leq \left( \int_{\mathbb{R}}\left( \iint_{V\mathbb{\times R}%
1396: ^{m}}\left\vert v\right\vert f\left( x,v,\mathbf{y},t\right) \mbox{d}v%
1397: \mbox{d}\mathbf{y}\right) ^{2}\mbox{d}x\right) ^{1/2}\leq
1398: \end{equation*}%
1399: \begin{equation*}
1400: \leq C\left( V\right) \left\Vert n\left( t\right) \right\Vert _{L^{2}}
1401: \end{equation*}%
1402: where we have used that $V$ is compact. Using similar ideas as in the proof
1403: of estimate (\ref{sxderestim}), we prove (\ref{stderestim}).
1404:
1405: \rightline{Q.E.D.}
1406:
1407: \medskip
1408:
1409: \begin{lemma}
1410: \label{lem7} Let $\mathbf{F}$ satisfy $(\ref{divF2})$. Then the
1411: characteristics $(\ref{X2})$ -- $(\ref{Y2})$ satisfy for all $0\leq s\leq t,$
1412: \begin{equation}
1413: \left[ \det \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}\left( s\right) %
1414: \right] ^{-1}\leq \exp \big[Ct\,\big\{1+\Pi \left( C\left\Vert
1415: f_{0}\right\Vert _{L^{1}}\right) \big\}\big]. \label{deter}
1416: \end{equation}
1417: \end{lemma}
1418:
1419: \noindent \textbf{Proof.}
1420: Using Lemma \ref{detGeneral} and (\ref{divF2}), we obtain
1421: \begin{equation*}
1422: \left[ \det \frac{\partial \mathbf{Y}}{\partial \mathbf{y}}(s)\right]
1423: ^{-1}=\exp \left[ \int_{s}^{t}\nabla _{\mathbf{y}}\cdot \mathbf{F}\left(
1424: \mathbf{S}\left( X(\tau ),\tau \right) ,\mathbf{Y}(\tau )\right) \mbox{d}%
1425: \tau \right] \leq
1426: \end{equation*}%
1427: \begin{equation*}
1428: \leq \exp \left[ C\int_{s}^{t}1+\Pi \left( \left\vert \mathbf{S}\left(
1429: X(\tau ),\tau \right) \right\vert \right) \mbox{d}\tau \right] \leq
1430: \exp \left[
1431: C t \left\{
1432: 1+\Pi
1433: \left(
1434: \sup_{0\leq \tau \leq t}\left\Vert \mathbf{S}\left( \tau \right)
1435: \right\Vert _{L^{\infty }}
1436: \right)
1437: \right\}
1438: \right].
1439: \end{equation*}%
1440: Using Lemma \ref{gradS}, we deduce (\ref{deter}).
1441:
1442: \vskip -2mm
1443:
1444: \rightline{Q.E.D.}
1445:
1446: \medskip
1447:
1448: \noindent \textbf{Proof of Theorem 2.} Using $(\ref{growthlambda})$ and $(%
1449: \ref{growthT})$, we obtain
1450: \begin{equation}
1451: T(v,v^{\prime },\mathbf{y})\leq C\left( 1+\Lambda \left( \left\vert \mathbf{C%
1452: }\right\vert \right) +\left\vert \frac{\mbox{d}\mathbf{C}}{\mbox{d}t}%
1453: \right\vert \right) . \label{estimT}
1454: \end{equation}%
1455: Integrating (\ref{hyperpell}) along the characteristic (\ref{X2}) -- (\ref%
1456: {Y2}) from $0$ to $t$ and using (\ref{estimT}), we obtain
1457: \begin{equation*}
1458: f\left( x,v,\mathbf{y},t\right) \leq f_{0}\left( X\left( 0\right) ,v,\mathbf{%
1459: Y}\left( 0\right) \right) +
1460: \end{equation*}%
1461: \begin{equation*}
1462: +\;C\left( V\right) \int_{0}^{t}
1463: \bigg\{\left( 1+\left[ \Lambda \left(
1464: \left\vert \mathbf{S}\right\vert \right) +\left\vert \frac{\partial \mathbf{S%
1465: }}{\partial t}\right\vert +\left\vert \frac{\partial \mathbf{S}}{\partial x}%
1466: \right\vert \right] \left( X\left( \tau \right) ,\tau \right) \right) \times
1467: \end{equation*}%
1468: \begin{equation*}
1469: \times \int_{V}f\left( X(\tau ),v^{\prime },\mathbf{Y}(\tau ),\tau \right) %
1470: \mbox{d}v^{\prime }\bigg\}\mbox{d}\tau +
1471: \end{equation*}%
1472: \begin{equation*}
1473: +\int_{0}^{t}\big\vert\nabla _{\mathbf{y}}\cdot \mathbf{F}\left( \mathbf{S(}%
1474: X\left( \tau \right) ),\mathbf{Y}\left( \tau \right) \right) \big\vert %
1475: f\left( X\left( \tau \right) ,v,\mathbf{Y}\left( \tau \right) ,\tau \right) %
1476: \mbox{d}\tau ,
1477: \end{equation*}%
1478: where we used that $V$ is compact. By virtue of assumption (\ref{divF2}), $%
1479: \left\vert \nabla _{\mathbf{y}}\cdot \mathbf{F}\right\vert $ is bounded by
1480: \hfill \break
1481: $%
1482: C\left( 1+\Pi \left( \left\vert \mathbf{S}\left( X(\tau ),\tau \right)
1483: \right\vert \right) \right) .$ Thus we have
1484: \begin{equation}
1485: f\left( x,v,\mathbf{y},t\right) \leq f_{0}\left( X\left( 0\right) ,v,\mathbf{%
1486: Y}\left( 0\right) \right) + \label{fint}
1487: \end{equation}%
1488: \begin{equation*}
1489: +\;C\left( V\right) \int_{0}^{t}
1490: \bigg\{\left( 1+\left[ \Lambda \left(
1491: \left\vert \mathbf{S}\right\vert \right) +\left\vert \frac{\partial \mathbf{S%
1492: }}{\partial t}\right\vert +\left\vert \frac{\partial \mathbf{S}}{\partial x}%
1493: \right\vert \right] \left( X\left( \tau \right) ,\tau \right) \right) \times
1494: \end{equation*}%
1495: \begin{equation*}
1496: \times \int_{V}f\left( X(\tau ),v^{\prime },\mathbf{Y}(\tau ),\tau \right) %
1497: \mbox{d}v^{\prime }\bigg\}\mbox{d}\tau +
1498: \end{equation*}%
1499: \begin{equation*}
1500: +C(V)\int_{0}^{t}\left( 1+\Pi \left( \left\vert \mathbf{S}\left( X(\tau
1501: ),\tau \right) \right\vert \right) \right) f\left( X\left( \tau \right) ,v,%
1502: \mathbf{Y}\left( \tau \right) ,\tau \right) \mbox{d}\tau .
1503: \end{equation*}%
1504: Using Lemma \ref{lem7}, we obtain for $t\geq 0,$
1505: \begin{equation*}
1506: \int_{\mathbb{R}\times V\times \mathbb{R}^{M}}\int_{V}f^{p}\left( X\left(
1507: \tau \right) ,v^{\prime },\mathbf{Y}\left( \tau \right) ,\tau \right)
1508: dv^{\prime }\mbox{d}x\mbox{d}v\mbox{d}\mathbf{y}=
1509: \end{equation*}%
1510: \begin{equation*}
1511: =\left\vert V\right\vert \int f^{p}\left( X\left( \tau \right) ,v^{\prime },%
1512: \mathbf{Y}\left( \tau \right) ,\tau \right) \left( \det \frac{\partial
1513: \mathbf{Y}}{\partial \mathbf{y}}\right) ^{-1}\left( \det \frac{\partial X}{%
1514: \partial x}\right) ^{-1}dv^{\prime }\mbox{d}X\mbox{d}\mathbf{Y}\leq
1515: \end{equation*}%
1516: \begin{equation*}
1517: \leq \left\vert V\right\vert e^{Ct}\int f^{p}\left( X\left( \tau \right)
1518: ,v^{\prime },\mathbf{Y}\left( \tau \right) ,\tau \right) \mbox{d}v^{\prime }%
1519: \mbox{d}X\mbox{d}\mathbf{Y}.
1520: \end{equation*}%
1521: We take the $p$-th power of (\ref{fint}) and integrate over $x,$ $v,$ and $%
1522: \mathbf{y}$ to get for $t\geq 0,$
1523: \begin{equation}
1524: \Vert f(t)\Vert \hbox{\raise -0.5mm \hbox{$_{L^{p}}$}}\leq \Vert f_{0}\Vert %
1525: \hbox{\raise -0.5mm \hbox{$_{L^{p}}$}}+C\left( V\right) \!e^{Ct}\int_{0}^{t}%
1526: \bigg\{(1+\Lambda \left( \left\Vert \mathbf{S}\left( \tau \right) \right\Vert
1527: _{L^{\infty }}\right) +\Pi \left( \left\Vert \mathbf{S}\left( \tau \right)
1528: \right\Vert _{L^{\infty }}\right) + \label{fLp}
1529: \end{equation}%
1530: \begin{equation*}
1531: +\left\Vert \frac{\partial \mathbf{S}}{\partial t}\left( \tau \right)
1532: \right\Vert _{L^{\infty }}+\left\Vert \frac{\partial \mathbf{S}}{\partial x}%
1533: \left( \tau \right) \right\Vert _{L^{\infty }}\bigg\}\times \Vert f\left(
1534: \tau \right) \Vert \hbox{\raise -0.5mm
1535: \hbox{$_{L^{p}}$}}\mbox{d}\tau .
1536: \end{equation*}%
1537: Using Lemma \ref{gradS}, we get for all $0\leq t\leq T$ and with $p=2$
1538: \begin{equation}
1539: \Vert f(t)\Vert \hbox{\raise -0.5mm \hbox{$_{L^2}$}}\leq \Vert f_{0}\Vert %
1540: \hbox{\raise -0.5mm \hbox{$_{L^{2}}$}}+Ce^{Ct}\!\!\!\int_{0}^{t}\left[
1541: 1+\left\Vert n\left( 0\right) \right\Vert _{L^{1}}\left\{ 1+\ln \left(
1542: \left\Vert n(\tau )\right\Vert _{L^{2}}+1\right) \right\} \right] \times
1543: \Vert f(\tau )\Vert \hbox{\raise -0.5mm
1544: \hbox{$_{L^2}$}}\}\mbox{d}\tau \leq \label{f-L2!}
1545: \end{equation}%
1546: \begin{equation*}
1547: \leq \Vert f_{0}\Vert \hbox{\raise -0.5mm \hbox{$_{L^{2}}$}}%
1548: +Ce^{Ct}\!\!\!\int_{0}^{t}\left[ 1+\left\Vert f\left( 0\right) \right\Vert
1549: _{L^{1}}\left\{ 1+\ln \left( \left\Vert f(\tau )\right\Vert
1550: _{L^{2}}+1\right) \right\} \right] \times \Vert f(\tau )\Vert
1551: \hbox{\raise -0.5mm
1552: \hbox{$_{L^2}$}}\}\mbox{d}\tau .
1553: \end{equation*}%
1554: By applying the Gronwall Lemma \ref{gronwall} to (\ref{f-L2!}), we obtain
1555: for $t\geq 0,$
1556: \begin{equation*}
1557: \Vert f(t)\Vert \hbox{\raise -0.5mm \hbox{$_{L^2}$}}\leq C\left(
1558: k_{i},k_{i}^{0},d_{i},\Lambda ,\Pi ,V,\Vert f_{0}\Vert
1559: \hbox{\raise -0.5mm
1560: \hbox{$_{L^{1}}$}},\Vert f_{0}\Vert \hbox{\raise -0.5mm \hbox{$_{L^{2}}$}}%
1561: \right) <\infty ,
1562: \end{equation*}%
1563: \begin{equation}
1564: \left\Vert \frac{\partial \mathbf{S}}{\partial x}\left( t\right) \right\Vert
1565: _{L^{\infty }}\leq C\left( k_{i},k_{i}^{0},d_{i},\Lambda ,\Pi ,V,\Vert
1566: f_{0}\Vert \hbox{\raise -0.5mm \hbox{$_{L^{1}}$}},\Vert f_{0}\Vert %
1567: \hbox{\raise -0.5mm \hbox{$_{L^{2}}$}}\right) <\infty, \label{SLinf}
1568: \end{equation}%
1569: where we used Lemma \ref{gradS} to get estimate (\ref{SLinf}).
1570: We now apply (\ref{SLinf}) to (\ref{fLp}) and we get (for $t\geq 0$ and for
1571: all $1\leq p\leq \infty $)%
1572: \begin{equation}
1573: \Vert f(t)\Vert \hbox{\raise -0.5mm \hbox{$_{L^p}$}}\leq C\left(
1574: k_{i},k_{i}^{0},d_{i},\Lambda ,\Pi ,V,\Vert f_{0}\Vert
1575: \hbox{\raise -0.5mm
1576: \hbox{$_{L^{1}}$}},\Vert f_{0}\Vert \hbox{\raise -0.5mm \hbox{$_{L^{2}}$}}%
1577: ,\Vert f_{0}\Vert \hbox{\raise -0.5mm \hbox{$_{L^{2}}$}}\right) <\infty ,
1578: \label{pestim}
1579: \end{equation}%
1580: i.e. we have obtained (\ref{fpestimate}). Using the elliptic equation (\ref%
1581: {ellipticS}), the second derivative of the extracellular signal can be
1582: expressed as
1583: \begin{equation}
1584: \frac{\partial ^{2}S_{i}}{\partial x^{2}}=-\frac{k_{i}}{d_{i}}n+\frac{%
1585: k_{i}^{0}}{d_{i}}S_{i}. \label{boot}
1586: \end{equation}%
1587: Using (\ref{pestim}) and the elliptic theory, we deduce (\ref{Spestimate}).
1588: Thus we complete the proof of Theorem \ref{theorem2}.
1589:
1590: \rightline{Q.E.D.}
1591:
1592: \medskip
1593:
1594: \noindent We conclude this section with two corollaries. They provide
1595: other conditions for the global existence of solutions. The proofs are
1596: omitted because they are similar to proofs of Theorem \ref{theorem1} and
1597: Theorem \ref{theorem2}.
1598:
1599: \begin{corollary}
1600: Assume $(\ref{growthylam})$ and $(\ref{growthT})$. Suppose there exists a
1601: non-negative, nondecreasing continuous function $\Pi \left( \cdot \right)
1602: \in C\left( \mathbb{R}\right) $ and $\gamma >0$ with $\omega \gamma \leq 1$
1603: satisfying
1604: \begin{equation}
1605: {\nabla }_{\mathbf{y}}{\cdot }\mathbf{F}\leq 0\text{ \ \ \ and \ \ }{\nabla }%
1606: _{\mathbf{y}}{\cdot }\mathbf{F}\left( \mathbf{z},\mathbf{y}\right) \leq
1607: C\left( 1+\Pi \left( \left\vert \mathbf{z}\right\vert \right) +\left\vert
1608: \mathbf{y}\right\vert ^{\gamma }\right) \label{divF1}
1609: \end{equation}%
1610: Assume that $f_{0}\in L^{1}\cap L^{\infty }({}\mathbb{R}\times V\times
1611: \mathbb{R}^{m})$ and let the initial condition $S_{0}\in \lbrack W^{2,p}({}%
1612: \mathbb{R})]^{M}$ satisfy $(\ref{ellipticS})$. Then there exists a global
1613: solution of the system $(\ref{hyperpell})$ -- $(\ref{ellipticS})$
1614: satisfying, for all $t\geq 0$
1615: \begin{equation*}
1616: f(\cdot ,\cdot ,\cdot ,t)\in L^{1}\cap L^{\infty }({}\mathbb{R}\times
1617: V\times \mathbb{R}^{m}),
1618: \end{equation*}%
1619: \begin{equation*}
1620: \mathbf{S}(\cdot ,t)\in \lbrack W^{2,p}(\mathbb{R})]^{M},\text{ \ \ \ \ for
1621: all }1\leq p<+\infty
1622: \end{equation*}%
1623: and initial conditions $f(\cdot ,\cdot ,\cdot ,0)=f_{0}(\cdot ,\cdot ,\cdot )
1624: $ and $\mathbf{S}(\cdot ,0)=\mathbf{S}_{0}(\cdot )$. \label{cor1}
1625: \end{corollary}
1626:
1627: \begin{corollary}
1628: Assume
1629: \begin{equation}
1630: \lambda (y_{1})\leq C,\qquad T(v,v^{\prime },\mathbf{y})\leq C(1+|\lambda
1631: (y_{1})|). \label{growthTlam}
1632: \end{equation}%
1633: We further assume that $\mathbf{F}$ satisfies either $(\ref{divF2})$ or
1634: $(\ref{divF1})$.
1635: Assume that $f_{0}\in L^{1}\cap L^{\infty }({}\mathbb{R}\times
1636: V\times \mathbb{R}^{m})$ and $\mathbf{S}_{0}\in \lbrack W^{1,\infty }({}%
1637: \mathbb{R})]^{M}$ with compact support. Then there exists a global solution of
1638: system of equations
1639: $(\ref{hyperpell})$ and $(\ref{parabSsim})$ satisfying
1640: \begin{equation*}
1641: f(\cdot ,\cdot ,\cdot ,t)\in L^{1}\cap L^{\infty }({}\mathbb{R}\times
1642: V\times \mathbb{R}^{m}),
1643: \end{equation*}%
1644: \begin{equation*}
1645: \mathbf{S}(\cdot ,t)\in \left[ W^{1,\infty }(\mathbb{R})\right] ^{M}
1646: \end{equation*}%
1647: and initial conditions $f(\cdot ,\cdot ,\cdot ,0)=f_{0}(\cdot ,\cdot ,\cdot )
1648: $ and $\mathbf{S}(\cdot ,0)=\mathbf{S}_{0}(\cdot )$.
1649: \end{corollary}
1650:
1651: \section{Discussion}
1652:
1653: \label{secdiscussion}
1654:
1655: The simplified model of the bacterial signal transduction was studied
1656: in \cite{Erban:2004:ICB,Erban:2004:STS} where equation (\ref{rom14})
1657: was given as (\ref{simmodel}). Using model (\ref{simmodel}) for the steady
1658: extracellular signal, one can derive the closed macroscopic (Keller-Segel,
1659: chemotaxis) equation
1660: for some parameter regimes. See \cite{Erban:2004:ICB} in 1D and \cite%
1661: {Erban:2004:STS} in 2D/3D. Hence, the transport equation framework can be
1662: used to study the macroscopic behaviour in terms of microscopic
1663: parameters for the steady extracellular signals and simplified models
1664: of the signal transduction.
1665:
1666: Here, we focused on more complex models where we coupled the complex
1667: transport equation (\ref{hyperp}) with the parabolic or elliptic equation
1668: for the signal (\ref{parabS}). The starting point of the analysis of such
1669: complex models is the existence theory. In this paper, we provided several
1670: sets of sufficient conditions for the global existence of solutions
1671: of system (\ref{hyperp}) -- (\ref{parabS}). There
1672: are many open questions remaining, e.g. the existence theory
1673: in $N$-dimensional physical space. It is also not clear whether one can
1674: derive the closed evolution equation for the density of cells $n(x,t)$ as
1675: we did for the simple case of noninteracting particles \cite%
1676: {Erban:2004:ICB,Erban:2004:STS}. If we are not able to derive
1677: the macroscopic equations then suitable computational approaches
1678: have to be used to study the macroscopic behaviour of bacteria
1679: \cite{Erban:2005:CAE}.
1680:
1681: There are several related results on kinetic models of the
1682: cellular movement. They often do not take the intracellular
1683: dynamics into account. Kinetic models were derived in \cite%
1684: {Alt:1980:BRW,Othmer:1988:MDB} using stochastic models of
1685: the movement of cells like bacteria or leukocytes.
1686: Reference \cite{Othmer:2002:DLT} addresses the formal
1687: diffusion limit of kinetic models to the classical Keller-Segel model.
1688: The discussion on issues of aggregation,
1689: blow-up, and collapse for certain class of random walks can be
1690: found in \cite%
1691: {Othmer:1997:ABC}. A Boltzmann-type kinetic model for chemotaxis
1692: without the internal dynamics coupled with an elliptic equation for
1693: the extracellular signal is studied in \cite{Chalub:2004:KMC} where
1694: global existence and rigorous diffusion limit to the Keller-Segel model
1695: were proven. In \cite{Hwang:2005:DDL,Hwang:2005:GSN}, a more general
1696: kinetic model was treated in
1697: two and three dimensions. A one-dimensional hyperbolic model was
1698: studied in
1699: \cite{Hwang:2005:GEC}. The papers
1700: \cite{Hwang:2005:DDL,Hwang:2005:GEC,Hwang:2005:GSN}
1701: took into account the effect of the gradient
1702: and the temporal derivative of the chemical signal and showed the global
1703: existence of smooth solutions with smooth initial data as well as the
1704: rigorous diffusive limit to the classical Keller-Segel model. However, all
1705: the rigorous global existence results so far have not included the temporal
1706: derivative of the signal in the growth condition of the turning frequency
1707: as we did in this paper. See
1708: also \cite{Perthame:2004:PMC} for more related works.
1709:
1710: \section*{Acknowledgements}
1711:
1712: This work was partially supported by the Max Planck Institute for
1713: Mathematics in Sciences, Biotechnology and Biological Sciences Research
1714: Council, University of Oxford, Trinity College Dublin and Linacre College,
1715: Oxford.
1716:
1717: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1718:
1719: \begin{thebibliography}{10}
1720:
1721: \bibitem{Alt:1980:BRW}
1722: W~Alt, \emph{Biased random walk models for chemotaxis and related diffusion
1723: approximations}, Journal of Mathematical Biology \textbf{9} (1980), 147--177.
1724:
1725: \bibitem{Barkai:1997:RSB}
1726: N.~Barkai and S.~Leibler, \emph{Robustness in simple biochemical networks},
1727: Nature \textbf{387} (1997), 913--917.
1728:
1729: \bibitem{Berg:1975:HBS}
1730: H.~Berg, \emph{How bacteria swim}, Scientific American \textbf{233} (1975),
1731: 36--44.
1732:
1733: \bibitem{Berg:1972:CEC}
1734: H.~Berg and D.~Brown, \emph{Chemotaxis in {E}sterichia coli analysed by
1735: three-dimensional tracking}, Nature \textbf{239} (1972), 500--504.
1736:
1737: \bibitem{Block:1983:AKB}
1738: S.~Block, J.~Segall, and H.~Berg, \emph{Adaptation kinetics in bacterial
1739: chemotactics}, Journal of Bacteriology \textbf{154} (1983), no.~1, 312--323.
1740:
1741: \bibitem{Bourret:1991:STP}
1742: R.~Bourret, K.~Borkovich, and M.~Simon, \emph{Signal transduction pathways
1743: involving protein phosphorylation in prokaryotes}, Annual Review of
1744: Biochemistry \textbf{60} (1991), 401--441.
1745:
1746: \bibitem{Chalub:2004:KMC}
1747: F.~Chalub, P.~Markowich, B.~Perthame, and C.~Schmeiser, \emph{Kinetic models
1748: for chemotaxis and their drift-diffusion limits}, Monatshefte f\"ur
1749: Mathematik \textbf{142} (2004), no.~1-2, 123--141.
1750:
1751: \bibitem{Cluzel:2000:UBM}
1752: P.~Cluzel, M.~Surette, and S.~Leibler, \emph{An ultrasensitive bacterial motor
1753: revealed by monitoring signaling proteins in single cells}, Science
1754: \textbf{287} (2000), 1652--1655.
1755:
1756: \bibitem{Dolak:2005:KMC}
1757: Y.~Dolak and C.~Schmeiser, \emph{Kinetic models for chemotaxis: Hydrodynamic
1758: limits and spatio-temporal mechanisms}, Journal of Mathematical Biology
1759: \textbf{51} (2005), no.~6, 595 -- 615.
1760:
1761: \bibitem{Erban:2005:ICB}
1762: R.~Erban, \emph{From individual to collective behaviour in biological systems},
1763: Ph.D. thesis, University of Minnesota, 2005.
1764:
1765: \bibitem{Erban:2005:CAE}
1766: R.~Erban, I.~Kevrekidis, and H.~Othmer, \emph{An equation-free computational
1767: approach for extracting population-level behavior from individual-based
1768: models of biological dispersal}, submitted to Physica D, 30 pages, 2005,
1769: available as http://arxiv.org/physics/0505179.
1770:
1771: \bibitem{Erban:2004:ICB}
1772: R.~Erban and H.~Othmer, \emph{From individual to collective behaviour in
1773: bacterial chemotaxis}, SIAM Journal on Applied Mathematics \textbf{65}
1774: (2004), no.~2, 361--391.
1775:
1776: \bibitem{Erban:2004:STS}
1777: \bysame, \emph{From signal transduction to spatial pattern formation in {{\em
1778: E. coli}}: A paradigm for multi-scale modeling in biology}, Multiscale
1779: Modeling and Simulation \textbf{3} (2005), no.~2, 362--394.
1780:
1781: \bibitem{Gegner:1992:AMR}
1782: J.~Gegner, D.~Graham, A.~Roth, and F.~Dahlquist, \emph{Assembly of an {MCP}
1783: receptor, {C}he{W} and kinase {C}he{A} complex in the bacterial chemotaxis
1784: signal transduction pathway}, Cell \textbf{70} (1992), 975--982.
1785:
1786: \bibitem{Hwang:2005:DDL}
1787: H.~Hwang, K.~Kang, and A.~Stevens, \emph{Drift-diffusion limits of kinetic
1788: models for chemotaxis: a generalization}, Discrete and Continuous Dynamical
1789: Systems B \textbf{5} (2005), no.~2, 319--334.
1790:
1791: \bibitem{Hwang:2005:GEC}
1792: \bysame, \emph{Global existence of classical solutions for a hyperbolic
1793: chemotaxis model and its parabolic limit}, to appear in Indiana University
1794: Mathematics Journal, 2005.
1795:
1796: \bibitem{Hwang:2005:GSN}
1797: \bysame, \emph{Global solutions of nonlinear transport equations for
1798: chemosensitive movements}, SIAM Journal on Mathematical Analysis \textbf{36}
1799: (2005), no.~4, 1177--1199.
1800:
1801: \bibitem{McKane:1996:MEA}
1802: L.~McKane and J.~Kandel, \emph{{M}icrobiology, {E}ssentials and
1803: {A}pplications}, McGraw-Hill, 1996.
1804:
1805: \bibitem{Morton-Firth:1999:FEB}
1806: C.~Morton-Firth, T.~Shimizu, and D.~Bray, \emph{A free-energy-based stochastic
1807: simulation of the {T}ar receptor complex}, Journal of Molecular Biology
1808: \textbf{286} (1999), 1059--1074.
1809:
1810: \bibitem{Othmer:1988:MDB}
1811: H.~Othmer, S.~Dunbar, and W~Alt, \emph{Models of dispersal in biological
1812: systems}, Journal of Mathematical Biology \textbf{26} (1988), 263--298.
1813:
1814: \bibitem{Othmer:2002:DLT}
1815: H.~Othmer and T.~Hillen, \emph{The diffusion limit of transport equations 2:
1816: Chemotaxis equations}, SIAM Journal on Applied Mathematics \textbf{62}
1817: (2002), 1222--1250.
1818:
1819: \bibitem{Othmer:1998:OCS}
1820: H.~Othmer and P.~Schaap, \emph{Oscillatory c{AMP} signaling in the development
1821: of {D}ictyostelium discoideum}, Comments on Theoretical Biology \textbf{5}
1822: (1998), 175--282.
1823:
1824: \bibitem{Othmer:1997:ABC}
1825: H.~Othmer and A.~Stevens, \emph{Aggregation, blow up and collapse: The {ABC}'s
1826: of taxis in reinforced random walks}, SIAM Journal on Applied Mathematics
1827: \textbf{57} (1997), no.~4, 1044--1081.
1828:
1829: \bibitem{Perthame:2004:PMC}
1830: B.~Perthame, \emph{{PDE} models for chemotactic movements: Parabolic,
1831: hyperbolic and kinetic}, Applications of Mathematics \textbf{49} (2004),
1832: no.~6, 539--564.
1833:
1834: \bibitem{Salyers:2001:MDD}
1835: A.~Salyers and D.~Whitt, \emph{{M}icrobiology, {D}iversity, {D}isease and the
1836: {E}nviroment}, Fitzgerald Science Press, 2001.
1837:
1838: \bibitem{Segall:1986:TCB}
1839: J.~Segall, S.~Block, and H.~Berg, \emph{Temporal comparisons in bacterial
1840: chemotaxis}, Proceedings of the National Academy of Sciences USA \textbf{83}
1841: (1986), 8987--8991.
1842:
1843: \bibitem{Spiro:1997:MEA}
1844: P.~Spiro, J~Parkinson, and H.~Othmer, \emph{A model of excitation and
1845: adaptation in bacterial chemotaxis}, Proceedings of the National Academy of
1846: Sciences USA \textbf{94} (1997), 7263--7268.
1847:
1848: \bibitem{Stock:1996:C}
1849: J.~Stock and M.~Surette, \emph{Chemotaxis}, Escherichia coli and salmonella:
1850: cellular and molecular biology (F.~Neidhardt, ed.), ASM Press, Washington,
1851: D.C., 1996, pp.~1103--1129.
1852:
1853: \end{thebibliography}
1854:
1855: %\bibliographystyle{amsplain}
1856: %\bibliography{bibrad}
1857:
1858: \end{document}
1859: