1: \documentclass[10pt,twocolumn]{article}
2: \usepackage{graphicx}
3: \usepackage{amsmath,amsthm,amssymb}
4: \usepackage{ccaption}
5: \usepackage{bigdelim,multirow}
6:
7:
8:
9: \setlength{\textwidth}{16cm}
10: \setlength{\textheight}{23cm}
11: \setlength{\oddsidemargin}{0cm}
12: \setlength{\evensidemargin}{0in}
13: \setlength{\topmargin}{0cm}
14:
15: \newtheorem{lem}{Lemma}
16: \newtheorem{ex}{Example}
17: \newtheorem{step}{Step}
18:
19:
20: \title{Stochastic approach to molecular interactions and
21: computational theory of metabolic and genetic regulations}
22: \author{H. Kimura, H. Okano and R.J. Tanaka\footnote{Corresponding author.
23: Tel.: +81(52)736-5861, Fax: +81(52)736-5862, Email: reiko@bmc.riken.jp}}
24: \date{Bio-Mimetic Control Research Center, RIKEN,\\
25: Shimo-shidami, Moriyama-ku, Nagoya 463-0003, Japan}
26: %\renewcommand{\baselinestretch}{2}
27:
28:
29: \begin{document}
30: \maketitle
31:
32:
33: %\newpage
34: \begin{abstract}
35: Binding and unbinding of ligands to specific sites of a macromolecule are
36: one of the most elementary molecular interactions inside the cell that embody
37: the computational processes of biological regulations.
38: The interaction between transcription factors and the operators of genes
39: and that between ligands and binding sites of allosteric enzymes are
40: typical examples of such molecular interactions.
41: In order to obtain the general mathematical framework of biological
42: regulations, we formulate these interactions as finite Markov processes
43: and establish a computational theory of regulatory activities of macromolecules
44: based mainly on graphical analysis of their state transition diagrams.
45: The contribution is summarized as follows:
46:
47: \noindent
48: (1) Stochastic interpretation of Michaelis-Menten equation is given. \\
49: (2) Notion of \textit{probability flow} is introduced in relation to
50: detailed balance.\\
51: (3) A stochastic analogy of \textit{Wegscheider condition} is given
52: in relation to loops in the state transition diagram. \\
53: (4) A simple graphical method of computing the regulatory activity in terms of
54: ligands' concentrations is obtained for Wegscheider cases. \\
55: (5) A general comutational scheme of the stationary probability distribution
56: is obtained in terms of probability flows.
57: \end{abstract}
58:
59:
60:
61: \noindent
62: \textit{keywords}: Biological regulation, molecular interaction, Markov process,
63: stationary probability distribution, probability flow
64:
65:
66: %\newpage
67: \section{Introduction}
68:
69: Control is a ubiquitous built-in mechanism supporting a variety of functions of living organisms. The advent of cybernetics in the late 1940's established a close link between control in living organisms and that of man-made artifact. Notions like feedback, feedforward, robustness, stability, and noise, which are fundamental in control engineering turned out to be relevant to the control mechanisms of living organisms. Rapid progress in molecular biology in the last fifty years opened up a new world of intracellular control of biochemical processes including those of metabolism and gene expression. Studying control mechanisms became an important issue of biology. Monod correctly called this field \textit{microscopic cybernetics} \cite{monod}, while others called it \textit{regulatory biology} \cite{yanofsky}. The continuing endeavor to clarify the material basis of regulations reveals vast complexity of regulatory networks associated with huge variety of material links inside a cell, which in turn motivated considerable interest in the quantitative analysis of regulation mechanisms through mathematical modeling and simulation.
70: The modeling/simulation paradigm has been used to reveal the underlying structural properties of complex regulatory networks.
71: Existence of cellular switches \cite{ackers}\cite{cherry}\cite{gardner}\cite{killer}\cite{mochizuki}\cite{ozbudak}\cite{ptashne}, evaluation of molecular fluctuations \cite{gilles}\cite{kaern}\cite{mcadams}\cite{ordenaa}, analysis of biological oscilations \cite{bliss}\cite{chen}\cite{griffith}, consideration of network robustness \cite{kitano}\cite{yi}, theoretical demonstration of cellular behaviors \cite{endy}\cite{mca}\cite{savageau}, are examples of its contributions. Excellent reviews of these achievements are found in \cite{smolen}\cite{thomas}\cite{tyson}.
72: Recently, we proposed the notion of compound control \cite{tanaka} which captures a salient characteristic feature of biological control. Compound control is a biological way of realizing proper input/output relationship of regulatory dynamics to cope with diverse, complicated and unpredictable environmental changes.
73:
74: It is paradoxical that the theoretical and/or physical basis of the modeling/simulation paradigm has not been well exploited, in spite of the remarkable success it has attained. For instance, in many papers on intracellular regulations, the traditional Michaelis-Menten-Hill equation in enzymology is extensively used for describing molecular interactions, e.g., the action of repressors in transcription regulation, or the effect of feedback \cite{bliss}\cite{gardner}\cite{mackey} \cite{mochizuki}\cite{ozbudak}\cite{wong}, without sound justification.
75:
76: The transcription rate is regulated through binding of transcription
77: factors in some domains of DNA,
78: the so-called \textit{cis}-regulatory elements \cite{bolouri}, that are able to affect transcription initiation. The affinity of \textit{cis}-regulatory elements to the transcription factors determines the overall transcription rate of the gene. They are correlated with each other and collectively exhibit synergistic effects through which a complex computation is performed to achieve adequate transcription control \cite{lloyd}.
79: These factors are represented in the context of a thermodynamical equilibrium energy distribution suggesting to plausible models of transcription regulation \cite{berg}\cite{bintu}\cite{panh}\cite{wolfd}.
80: This paper aims to establish a unified framework with solid mathematical grounding and physical plausibility that is able to capture various bio-physical attributes of intracellular regulations.
81:
82: A significant portion of genes encode enzymes which are involved in metabolic control, another important form of intracellular regulation. Metabolic control is a relatively old subject of bio-chemistry, where allosteric enzymes are the major agents of control. Allosteric enzyme is a macro-molecule that has several binding sites which can bind substrate and ligands. The binding pattern of ligands changes the conformations of the enzyme, through which the activity of the enzyme for the target reaction is controlled. Quantification of the activity changes of allosteric enzyme has been a central issue in biochemistry since the classical Monod-Wyman-Changeux model was first proposed \cite{monodj}\cite{wyman}. Since then, various theoretical and experimental methods have been used to predict the conformation changes due to substrate binding and modification by ligands \cite{berg}\cite{coppey}\cite{freire}\cite{gibson}\cite{gunasekaran}\cite{king}\cite{ming}\cite{han}\cite{segel}\cite{kampen}\cite{Zaslaver}.
83:
84: The regulatory mechanism of allosteric enzymes is in some sense similar to that of operons, in spite of the obvious functional and structural differences between them. Both of them are controlled by interactions between binding sites (catalytic sites for metabolic cases and \textit{cis}-regulatory elements for genetic cases) and binding factors (ligands and substrate in the metabolic case and transcription factors in the genetic case). In the metabolic case, interactions take place between proteins, while in the genetic case they take place between protein and DNA. It is interesting that although the allosteric protein is a player in both regulations, it plays totally different roles. In the metabolic control, it is controlled by ligands, while in the genetic control it acts as transcription factors to control genes.
85:
86: The protein/protein interactions of metabolic regulation and protein/DNA interactions in transcription control are weak and essentially reversible,
87: and they share many properties.
88: In fact, equilibrium statistical mechanics is used to quantify both
89: metabolic regulation \cite{gibson} and genetic regulation \cite{bintu}\cite{panh}\cite{wolfd}.
90: If we view the binding/unbinding processes taking place between binding sites and binding factors in both cases as a common stochastic process, it suggests a common theoretical framework to deal with metabolic and genetic regulations in a unified way.
91:
92: Along this line, we propose a common theoretical framework for describing biological regulators of metabolic and genetic regulations
93: to facillitate quantitative descriptions of intracellular bio-chemical processes.
94: The core of our approach is to describe protein/protein and protein/DNA interactions with a common Markov process focusing on the behavior of a single molecule rather than a population of molecules. This viewpoint enables us to describe phenomena with a finite state Markov process, which greatly simplifies
95: the argument by avoiding the difficulties associated with infinite dimensionality
96: \cite{munsky}.
97: All the quantification features associated with various complicated physical and biological phenomena
98: can then be condensed into transition probabilities.
99: We introduce a new notion of probability flow that offers a new insight
100: into the stationary distribution of the Markov process.
101: The stationary equation is regarded as representing the
102: conservation of probability flows at each node.
103: It is closely related to the Wegscheider condition derived a century ago.
104: Based on the notion of probability flow, we derive a representation of
105: the stationary probability distribution, which leads to a unified quantitative form
106: for the general biological regulator.
107: Our method of computing the stationary probability distribution
108: dramatically simplifies the classical King and Altman method \cite{king},
109: which is used for computing the stationary distribution of chemical processes.
110: It is expected to yield a new computational tool for operon regulation
111: that can deal with the increasing complexity of operons \cite{santillan}.
112:
113: Metabolic and genetic controls have been investigated relatively independently
114: because of their apparent large differences in modality.
115: Therefore, there have been few studies dealing with systems that combine metabolic and genetic regulations \cite{Zaslaver}. The unified framework described in this paper will enhance our understanding of intracellular biochemical processes.
116:
117: In the next section, we formulate a biological regulator in an abstract way as a Markov process. The abstract biological regulator is defined through the stationary solution of the Master Equation. In Section \ref{sec3}, we introduce the notion of probability flow and explain its relevance to the stationary distribution.
118: Sections \ref{sec4} and \ref{sec5} deal with the single loop case and introduce the Wegscheider condition, which guarantees that the probability flow vanishes. Section \ref{sec6} generalizes the results of the preceding section to multi-loop cases.
119: Section \ref{sec7} discusses the meaning of transition probabilities to deduce
120: biologically meaningful representations of biological regulations.
121:
122:
123: \section{Characterization of biological regulations as a
124: finite-state Markov process}
125:
126: The main stage of genetic control is transcription regulation
127: whose fundamental mechanism is to change the transcription rate
128: of the operon via binding/unbinding of transcription factors
129: to and from their \textit{cis}-regulatory elements.
130: The main agent of metabolic control is the allosteric enzyme
131: that determine the rate of the target chemical reaction
132: via binding/unbinding of substrate and ligands.
133: The fundamental common feature of both regulations lies in
134: the molecular interaction between sites of the macro-molecule and
135: molecular binding factors that work through binding/unbinding processes
136: (see Table~\ref{table:1}).
137:
138: \begin{table*}[bt]
139: \caption{
140: Comparison of genetic and metabolic regulations.}\label{table:1}
141: \begin{center}
142: \begin{tabular}{c|p{2.6cm}|p{2.6cm}|p{2.6cm}|p{2.6cm}}
143: & \raisebox{-0.5\normalbaselineskip}{Agent} &
144: \raisebox{-0.5\normalbaselineskip}{Site} &
145: \raisebox{-0.5\normalbaselineskip}{Factor} & Control\par Parameters \\
146: \hline \raisebox{-0.5\normalbaselineskip}{Genetic} &
147: \raisebox{-0.5\normalbaselineskip}{Operons} &
148: \textit{Cis}-regulatory elements & Transcription factors & Transcription rate \\
149: \hline \raisebox{-0.5\normalbaselineskip}{Metabolic} &
150: Allosteric \par enzymes & \raisebox{-0.5\normalbaselineskip}{Catalytic sites} &
151: Ligands and \par substrates & \raisebox{-0.5\normalbaselineskip}{Reaction rate}
152: \end{tabular}
153: \end{center}
154: \end{table*}
155:
156: %shown in Table~\ref{table:1}
157:
158: These processes are obviously random subject to certain thermodynamical
159: constraints.
160: Our idea is to establish a mathematical framework to describe
161: such regulatory mechanisms in a unified way.
162: We now formulate the biological regulation in a somewhat
163: abstract way. We assume that the macro-molecule, which is
164: the main agent of the regulation, has $n$ binding sites (BS),
165: each of which can bind some of $m$ ligands with different
166: affinities that depend on the binding patterns of sites.
167: Instead of the ligand, we use the word \textit{binding factors} (BF)
168: to emphasize the bilateral symmetry of sites to be bound
169: and factors to bind.
170: Each BS can bind only one BF at each time.
171: The binding sites are denoted by $b_{1},b_{2},\cdots,b_{n}$,
172: while the binding factors by $U_{1},U_{2},\cdots,U_{m}$.
173: The state of the regulator is denoted by $n$-tuples of $m$
174: alphabet $U_{1},U_{2},\cdots,U_{m}$ denoting the BFs being bound
175: and $\phi $ which represents the empty (unoccupied) site.
176: If the regulator has three BSs, for example,
177: $S=(U_{1}\phi U_{2})$ denotes the state with
178: $b_{1}$ and $b_{3}$ being bound by $U_{1}$ and $U_{2}$,
179: respectively, and with $b_{2}$ empty.
180: Let $N$ be the total number of non-empty states,
181: which is obviously finite.
182: Since we have the empty state $S_{0}=(\phi\phi\cdots\phi)$,
183: the total number of states is $N+1$.
184:
185: The state of the regulator changes as its binding pattern
186: changes, i.e., as BFs bind to and dissociate from BSs.
187: These are clearly stochastic phenomena and are legitimately
188: described as stochastic process.
189: Let the transition probability from the state $S_{j}$ to $S_{i}$
190: during an infinitesimal time duration $\Delta t$ be denoted by
191: $q_{ij}\Delta t$.
192: All the quantitative features of the complex biochemical
193: process associated with protein/protein interactions and
194: protein/DNA interactions including origomerization \cite{han}\cite{wyman},
195: conformational change \cite{freire}\cite{panh} and
196: target localization \cite{coppey} can be adequately
197: represented in terms of $q_{ij}$.
198: For the sake of clarity, we assume that binding or
199: unbinding of only one BF to and from a BS occurs during
200: the infinitesimal time duration $\Delta t$.
201: Thus, the transition from $(U_{1}\phi U_{2})$ to
202: $(U_{1}\phi U_{3})$ is regarded as consequtive
203: transitions from $(U_{1}\phi U_{2})$ to $(U_{1}\phi \phi)$
204: and from $(U_{1}\phi \phi)$ to $(U_{1}\phi U_{3})$.
205: The transition probability for unit time $q_{ij}$ represents the rate
206: of reaction from $S_{j}$ to $S_{i}$, and sometimes
207: it is regarded as identical to the rate constant
208: of the corresponding reaction.
209: We define ${\bf A}_{i}$ to be the set of states that are accessible
210: from $S_{i}$ during the infinitesimal time duration $\Delta t$.
211: We call it the \textit{adjacent set} of $S_{i}$.
212:
213: Let $p_{i}(t)$ be the probability that the regulator is
214: in a state $S_{i}$ at time $t$.
215: Then, the stochastic time-evolution of the regulator is
216: described by the \textit{chemical master equation} (CME),
217: \begin{equation}
218: \frac{dp_{i}}{dt}=\sum_{j\in \bf{A}_{i}}q_{ij}p_{j}-\sum_{j\in \bf{A}_{i}}q_{ji}p_{i}.
219: \quad i=0,1,\cdots,N.
220: \label{eq2-1}
221: \end{equation}
222: The notation $\sum_{j\in {\bf A}_{i}}$ denotes the sum of all $j$
223: such that $S_{j}\in {\bf A}_{i}$.
224: The first term of the right-hand side of (\ref{eq2-1})
225: represents the net probability coming from adjacent states to $S_{i}$
226: while the second term the net probability
227: leaving $S_{i}$ to adjacent states.
228: The CME (\ref{eq2-1}) is written in the matrix form as
229: \begin{equation}
230: \frac{dp(t)}{dt}=Qp(t)
231: \label{eq2-2}
232: \end{equation}
233: where $p(t)$ denotes the $(N+1)$-dimensional vector whose $(i+1)$-th
234: component is $p_{i}(t)$.
235: The matrix $Q$ in (\ref{eq2-2}) is called a
236: \textit{transition matrix}.
237:
238: A finite Markov process can be described by the state-transition
239: diagram.
240: In Figure~\ref{fig:1} shows examples of state-transition diagrams.
241: The most salient feature of the Markov process for our purpose
242: is the reflection property that $q_{ij}\neq 0$
243: always implies $q_{ji}\neq 0$ because of the reversibility of the molecular
244: interactions we are interested in this paper.
245: In terms of the adjacent set ${\bf A}_{i}$,
246: $S_{j}\in {\bf A}_{i}$ always implies $S_{i}\in{\bf A}_{j}$.
247: \begin{figure}[tb]
248: \begin{center}
249: \includegraphics[width=0.4\textwidth]{fig1.eps}
250: \end{center}
251: \caption{State transition diagram of
252: (a) Example 1, (b) Example 2,(c) Example 3.}
253: \label{fig:1}
254: \end{figure}
255:
256:
257: It is known \cite{kampen} that the solution of (\ref{eq2-2})
258: converges to a unique equilibrium stationary solution $p$
259: that satisfies
260: \begin{equation}
261: Qp=0,
262: \label{eq2-3}
263: \end{equation}
264: or equivalently,
265: \begin{equation}
266: \sum_{j\in {\bf A}_{i}}q_{ij}p_{j}-\sum_{j\in {\bf A}_{i}}q_{ji}p_{i}=0.
267: \label{eq2-4a}
268: \end{equation}
269: Equaiton (\ref{eq2-3}) or (\ref{eq2-4a}) is called
270: the \textit{stationary state equation} (SSE).
271: Usually, the convergence is faster than the chemical reactions
272: controlled by the regulator.
273: Therefore, we can assume that the regulator is always in a
274: stationary distribution satisfying (\ref{eq2-3}),
275: together with the normalization constraint,
276: \begin{equation}
277: \sum^{N}_{i=0}p_{i}=1.
278: \label{eq2-4}
279: \end{equation}
280:
281: Let $\gamma_{i}$ be the activity of the regulator at state $S_{i}$.
282: The overall activity $\gamma$ of the regulator is then
283: defined as the average activity with respect to the
284: stationary distribution $p$, i.e.,
285: \begin{equation}
286: \gamma=\sum^{N}_{i=0}\gamma_{i}p_{i}.
287: \label{eq2-5}
288: \end{equation}
289: The examples that follow show that the above definition of the regulation
290: activity adequately describes both genetic and metabolic regulations.
291:
292: \begin{ex}\upshape
293: (Stochastic version of Michaelis-Menten equation)\\
294: Consider a molecule of an enzyme $E$ that catalyses the reaction
295: of producing a product $P$ from a substrate $S$, i.e.,
296: \begin{equation}
297: E+S \overset{k_{1}}{\underset{k_{-1}}{\rightleftarrows}} ES
298: \overset{k_{2}}{\rightarrow} E+P.
299: \label{eq2-6}
300: \end{equation}
301: An enzyme molecule can be regarded as a biological regulator
302: with one BS and one BF (the substrate).
303: The enzyme has the two state $S_{0}$ and $S_{1}$,
304: the empty state and the state occupied by a substrate molecule,
305: respectively.
306: The transition diagram is shown in Fig.~\ref{fig:1}(a).
307: The transition matrix $Q$ in this case is written as
308: \begin{equation}
309: Q=\begin{bmatrix}-q_{10}&q_{01}\\
310: q_{10}&-q_{01}\end{bmatrix},
311: \label{eq2-7}
312: \end{equation}
313: and the stationary distribution is given by
314: \begin{equation}
315: p_{0}=\frac{1}{1+\frac{q_{10}}{q_{01}}},\quad p_{1}=\frac{\frac{q_{10}}{q_{01}}}{1+\frac{q_{10}}{q_{01}}}.
316: \label{eq2-8}
317: \end{equation}
318: Since $q_{01}$ represents the probability of unbinding the substrate $S$ from
319: the BS during unit time, we can identy it with $k_{-1}$.
320: In addition, the probability of binding $S$ to the BS is proportional
321: to the concentration $\left[S \right]$ of the substrate with $k_{1}$
322: being its rate coefficient. Thus,
323: \begin{equation}
324: q_{01}=k_{-1},\quad q_{10}=k_{1}\left[S \right],
325: \label{eq2-9}
326: \end{equation}
327: which gives
328: \begin{equation}
329: p_{0}=\frac{K_{1}}{K_{1}+\left[S \right]}, \quad
330: p_{1}=\frac{\left[S \right]}{K_{1}+\left[S \right]}, \quad
331: K_{1}=\frac{k_{-1}}{k_{1}}.
332: \label{eq2-10}
333: \end{equation}
334: Since $p_{1}$ denotes the probability of the state of the enzyme with $S$
335: being bound (ES),
336: the total number of ES molecules is given by $p_{1}\left[E \right]_{T}$,
337: where $\left[E \right]_{T}$ denotes the concentration of the total enzyme,
338: which is assumed to be constant.
339: Therefore, since the maximum reaction rate is given by
340: $k_{2}\left[E \right]_{T}$, the ratio of the reaction rate $\upsilon$ to its
341: maximum $\upsilon_{\max}$ is given by
342: \begin{equation}
343: \frac{\upsilon }{\upsilon_{\max}}=\frac{k_{2}p_{1}\left[E \right]_{T}}{k_{2}\left[E \right]_{T}}=p_{1}=\frac{\left[S \right]}{K_{1}+\left[S \right]},
344: \label{eq2-11}
345: \end{equation}
346: which is the celebrated \textit{Michaelis-Menten equation}.
347: Here, the rapid equilibrium assumption used in
348: deriving the Michaelis-Menten equation \cite{segel} has been replaced
349: with the notion of rapid
350: convergence of the probability distribution to a stationary one.
351: Note that we consider the behavior of a single enzyme molecule
352: rather than the collective behavior of the enzyme and enzyme-substrate complex.
353: In this way, a stochastic formulation of biological regulation gives an alternative interpretation of the Michaelis-Menten equation.
354: It should be noted that we have not assumed the enzyme concentration
355: is constant, which is required to derive
356: Michaelis-Menten equation in the traditional way.
357:
358: It is not difficult to see that the standard deviation
359: $\sigma_{\upsilon}=\sqrt{\bar{\upsilon^{2}}-\bar{\upsilon}^{2}}$ is given by
360: \[
361: \sigma _{\upsilon}=k_{2}\frac{\sqrt{K_{1}\left[S \right]}}{K_{1}+\left[S \right]},
362: \]
363: which gives a rough estimate of the precision of the approximation (\ref{eq2-11}).
364: This is a bonus of the stochastic version of the Michaelis-Menten equation
365: that we have derived.
366: \end{ex}
367:
368: \begin{ex}\upshape
369: (Allosteric enzyme with an activator \cite{segel})\\
370: An allosteric enzyme $e$ with one binding site for
371: an activator $A$ in addition to the one for a substrate $S$ is described as
372: \begin{align}
373: \label{ex2-1}
374: &E\ + \ S\ \overset{k_{1}}{\underset{k_{-1}}{\rightleftarrows}} ES \overset{k_{p}}{\rightarrow}\ E\ +\ P \nonumber\\
375: &+ \hspace*{19mm} + \nonumber\\
376: &A \hspace*{21mm} A \\
377: {\scriptstyle k_{-2}}\!&\! \uparrow\downarrow \!{\scriptstyle k_{2}} \qquad\ \ {\scriptstyle k_{-3}}\! \uparrow\downarrow \!{\scriptstyle k_{3}}
378: \nonumber\\
379: E&A\ +\ S\ \overset{k_{4}}{\underset{k_{-4}}{\rightleftarrows}}\ ESA \ \overset{k_{pA}}{\rightarrow}\ EA\ +\ P
380: \nonumber
381: \end{align}
382: The enzyme has now the two binding sites, one for the substrate
383: and the other for the activator $A$.
384: The enzyme has four states $S_{0}=(\phi\phi)$, $S_{1}=(S\phi)$,
385: $S_{2}=(SA)$, $S_{3}=(\phi A)$.
386: The state transition diagram is shown in Fig.~\ref{fig:1}(b).
387: We can associate the transition probabilities with
388: the rate constants in (\ref{ex2-1}) as
389: \begin{align}
390: &q_{10}=k_{1}\left[S \right],\ q_{01}=k_{-1}, \ q_{21}=k_{3}\left[A \right],
391: \ q_{12}=k_{-3}, \nonumber \\
392: &q_{32}=k_{-4}, \ q_{23}=k_{4}\left[S \right],
393: \ q_{03}=k_{-2}, \ q_{30}=k_{2}\left[A \right],\label{eq2-13}
394: \end{align}
395: where $\left[S \right]$ and $\left[A \right]$ denote the
396: concentrations of the substrate and the activator, respectively.
397: The transition matrix is given by
398: \begin{equation}
399: Q=\begin{bmatrix}
400: D_{0}&q_{01}&0&q_{03}\\
401: q_{10}&D_{1}&q_{12}&0\\
402: 0&q_{21}&D_{2}&q_{23}\\
403: q_{30}&0&q_{32}&D_{3}
404: \end{bmatrix}.
405: \end{equation}
406: where $D_{0}=-(q_{10}+q_{30})$, $D_{1}=-(q_{01}+q_{21})$,
407: $D_{2}=-(q_{12}+q_{32})$, $D_{3}=-(q_{03}+q_{23})$.
408: The rate of the target chemical reactions producing $P$ is given by
409: \[
410: \gamma =k_{p}p_{1}+k_{pA}p_{2}
411: \]
412: %shown in Fig.~\ref{fig:1}
413:
414: \noindent
415: The actual computation of the stationary probability distribution
416: for this example is done in Example 6, where you see the result is
417: very complex.
418: \end{ex}
419:
420: \begin{ex}\upshape
421: (Allosteric enzyme with both activator and inhibitor)\\
422: The reaction scheme is written as
423: \begin{align}
424: \label{ex3-16}
425: &EI\ +\ S\ \overset{k_{7}}{\underset{k_{-7}}{\rightleftarrows}}\ ESI \ \overset{k_{pI}}{\rightarrow}\ EI\ +\ P \nonumber \\
426: {\scriptstyle k_{6}}\!&\! \uparrow\downarrow \!{\scriptstyle k_{-6}} \qquad\quad\ {\scriptstyle k_{5}}\! \uparrow\downarrow \!{\scriptstyle k_{-5}}
427: \nonumber\\
428: &\ I \hspace*{24mm} I \nonumber \\
429: &+ \hspace*{22mm} + \nonumber \\
430: &E\ +\ S\ \ \overset{k_{1}}{\underset{k_{-1}}{\rightleftarrows}}\ \ ES \ \overset{k_{p}}{\rightarrow}\ E\ +\ P \\
431: &+ \hspace*{22mm} + \nonumber \\
432: &A \hspace*{24mm} A \nonumber \\
433: {\scriptstyle k_{-2}}\!&\! \uparrow\downarrow \!{\scriptstyle k_{2}} \qquad\quad\ {\scriptstyle k_{-3}}\! \uparrow\downarrow \!{\scriptstyle k_{3}}
434: \nonumber\\
435: &\! EA\ +\ S \ \overset{k_{4}}{\underset{k_{-4}}{\rightleftarrows}}\ ESA \ \overset{k_{pA}}{\rightarrow}\ EA\ +\ P \nonumber
436: \end{align}
437: where I denotes the inhibitor.
438: The enzyme has six states, namely,
439: $S_{0}=(\phi\phi)$, $S_{1}=(\phi A)$, $S_{2}=(SA)$,
440: $S_{3}=(S \phi)$, $S_{4}=(SI)$, $S_{5}=(\phi I)$.
441: The state transition diagram is shown in Fig.~\ref{fig:1}(c).
442: The right part is identical to the diagram of Fig.~\ref{fig:1}(a).
443: The state transition matrix is given by
444: \begin{equation}
445: Q=\begin{bmatrix}
446: D_{0}&q_{01}&0&q_{03}&0&q_{05}\\
447: q_{10}&D_{1}&q_{12}&0&0&0\\
448: 0&q_{21}&D_{2}&q_{23}&0&0\\
449: q_{30}&0&q_{32}&D_{3}&q_{34}&0\\
450: 0&0&0&q_{43}&D_{4}&q_{45}\\
451: q_{50}&0&0&0&q_{54}&D_{5}
452: \end{bmatrix},
453: \end{equation}
454: where $D_{0}=-(q_{10}+q_{30}+q_{50})$,
455: $D_{1}=-(q_{01}+q_{21})$, $D_{2}=-(q_{12}+q_{32})$,
456: $D_{3}=-(q_{03}+q_{23}+q_{43})$, $D_{4}=-(q_{34}+q_{54})$,
457: $D_{5}=-(q_{05}+q_{45})$.
458: The rate of the target chemical reactions is given by
459: \[
460: \gamma=k_{pA}p_{2}+k_{p}p_{3}+k_{pI}p_{4}
461: \]
462: The computation of the stationary probability distribution is
463: done in Example 6.
464: \end{ex}
465:
466: \begin{ex}\upshape
467: ({\it Lac} operon)\\
468: The {\it lac} operon, which has been a subject of
469: genetic molecular biology for more than fifty years
470: (\cite{bliss}\cite{santillan}\cite{wong}\cite{Zaslaver}), can be
471: regarded as a biological regulator.
472: A number of different models have been proposed for it including
473: a very complicated model with seven \textit{cis}-regulatory elements \cite{santillan}.
474: The standard model has three binding sites \cite{Zaslaver}.
475: One is the promoter with RNAP as its unique binding factor.
476:
477: There are two binding factors associated with the {\it lac} operon,
478: cAMP-CRP and LacI, which have their own
479: binding sites. They are independent of each other.
480: cAMP-CRP is an activator of the operon, while LacI
481: is a repressor.
482: Binding of LacI at its site prevents the other factors from
483: binding to their sites.
484:
485: Denote the RNAP, cAMP-CRP and LacI by $U_{1}$, $U_{2}$ and $U_{3}$, respectively.
486: Then, the {\it lac} operon has the following five states;
487: \begin{align*}
488: &S_{0}=(\phi\phi\phi),\ S_{1}=(U_{1}\phi\phi),\ S_{2}=(U_{1}U_{2}\phi),\\
489: &S_{3}=(\phi U_{2}\phi),\ S_{4}=(\phi\phi U_{3}).
490: \end{align*}
491: The transition diagram is shown in Fig.~\ref{fig:2}.
492: % shown in Fig~\ref{fig:2}
493: The transcription rate is given by
494: \[ \gamma=\gamma_{1}p_{1}+\gamma_{2}p_{2}, \]
495: where $\gamma_{1}$ and $\gamma_{2}$ are the transcription rates corresponding to $S_{1}$ and $S_{2}$.
496: \end{ex}
497: \begin{figure}[h]
498: \begin{center}
499: \includegraphics[width=0.3\textwidth]{fig2.eps}
500: \end{center}
501: \caption{State transition diagram of {\it lac} operon.}
502: \label{fig:2}
503: \end{figure}
504:
505: Recently, Santillan et al. proposed a lac operon model with
506: five binding sites \cite{santillan}.
507: The number of binding patterns of its \textit{cis}-regulatory units is
508: at least 50.
509: Figure~\ref{fig:3} shows the transition diagram which is very complicated.
510: We need an efficient computational theory to handle complex operons
511: like this.
512: % shown in Fig~\ref{fig:3}
513: \begin{figure}[h]
514: \begin{center}
515: \includegraphics[width=0.5\textwidth]{fig3.eps}
516: \end{center}
517: \caption{Transition diagram of {\it lac} operon with 50 states.}
518: \label{fig:3}
519: \end{figure}
520:
521: \section{Conservation of probability flows in stationary distribution
522: and loop-free transition diagram}
523: \label{sec3}
524:
525: In this section, we introduce the notion of probability flow and discuss its meaning
526: in solving the SSE (\ref{eq2-4a}).
527: The SSE (\ref{eq2-4a}) can be written as
528: \begin{equation}
529: \sum_{j\in {\bf A}_{i}}(q_{ij}p_{j}-q_{ji}p_{i})=0,\quad i=0,1,\cdots,N.
530: \label{eq8}
531: \end{equation}
532: The term $q_{ij}p_{j}-q_{ji}p_{i}$ represents the net probability
533: of the state transition from $S_{j}\in {\bf A}_{i}$ to $S_{i}$.
534: Hence, it is reasonable to call it the \textit{probability flow from
535: $S_{j}$ to $S_{i}$}, which is denoted by
536: \begin{equation}
537: \rho_{ij}=q_{ij}p_{j}-q_{ji}p_{i}.
538: \label{eq9}
539: \end{equation}
540: If $\rho_{ij}>0$, the probability flows from $S_{j}$ to $S_{i}$
541: and if $\rho_{ij}<0$, it flows oppositely.
542: Clearly, it is \textit{skew-symmetric}, i.e.,
543: \begin{equation}
544: \rho_{ij}+\rho_{ji}=0.
545: \label{eq10}
546: \end{equation}
547: The SSE (\ref{eq8}) is represented in terms of probability flow as
548: \begin{equation}
549: \sum_{j\in {\bf A}_{i}}\rho_{ij}=0, \quad \forall i,
550: \label{eq11}
551: \end{equation}
552: which implies that the probability flows are conserved
553: at each state node.
554: In other words, the net probability flow incoming to $S_{i}$
555: (the sum of positive $\rho _{ij}$) is equal to the outgoing
556: flow from $S_{i}$ (the sum of negative $\rho_{ij}$), as
557: is illustrated in Figure~\ref{fig:4}.
558: Here, it is important to notice that the probability flow is directed.
559: % shown in Fig.~\ref{fig:4}
560:
561:
562: \begin{figure}[h]
563: \begin{center}
564: \includegraphics[width=0.3\textwidth]{fig4.eps}
565: \end{center}
566: \caption{Conservation of probability flows.}
567: \label{fig:4}
568: \end{figure}
569:
570:
571: In chemical kinetics, one usually assumes that the equilibrium state
572: satisfies the relations
573: \begin{equation}
574: q_{ij}p_{j}=q_{ji}p_{i},\quad \forall j\in {\bf A}_{i},\quad \forall i,
575: \label{eq12}
576: \end{equation}
577: if we interpret $q_{ij}$ as the kinetic rate coefficient of the reaction
578: $S_{j}\to S_{i}$.
579: The relations (\ref{eq12}) are called \textit{detailed balance} in
580: chemical kinetics \cite{heinrich}.
581: Due to (\ref{eq9}), the detailed balance holds if and only if
582: $\rho_{ij}=0$, $\forall i$, $j\in {\bf A}_{i}$; i.e., all the probability
583: flows vanish.
584: The relation (\ref{eq12}) is written as
585: \begin{equation}
586: p_{i}=r_{ij}p_{j}
587: \label{eq13a}
588: \end{equation}
589: where $r_{ij}$ is called the \textit{transition ratio} from
590: $S_{j}$ to $S_{i}$ and is defined as
591: \begin{equation}
592: r_{ij}=\frac{q_{ij}}{q_{ji}}.
593: \label{eq14a}
594: \end{equation}
595: The transition ratio (TR) is directed as shown in Fig.~\ref{fig:5}(a)
596: and corresponds to the equilibrium coefficient of the chemical reaction
597: $S_{j}\leftrightarrows S_{i}$.
598: It is sometimes convenient to describe the transition
599: diagram in terms of the transition ratios (TR), instead of
600: transition probabilities, as is shown in Fig.~\ref{fig:5}(b).
601: We call such a diagram a \textit{TR diagram} to distinguish it from the usual transition
602: diagram. Note that
603: \begin{equation}
604: r_{ij}=r^{-1}_{ji}.
605: \label{e24q}
606: \end{equation}
607: % shown in Fig.~\ref{fig:5}
608:
609: \begin{figure}[h]
610: \begin{center}
611: \includegraphics[width=0.35\textwidth]{fig5.eps}
612: \end{center}
613: \caption{TR Diagrams.
614: (a) Description of TR, (b) TR diagram of Fig.~\ref{fig:1}(b).}
615: \label{fig:5}
616: \end{figure}
617:
618:
619: \noindent
620: The direction of an edge can be freely assigned in TR diagrams,
621: but it must be consistent with the direction of the TR,
622: in the sense that an edge with $r_{ij}$ as its TR must be directed
623: from $S_{j}$ to $S_{i}$.
624: Now, we shall prove that the probability flows vanish at all
625: edges unless the transition diagram has a loop.
626: Here, a loop is defined in the context of the TR diagram.
627: In other words, if the transition diagram does not have a loop
628: (loop-free), all the probability flows vanish.
629: To see this remarkable fact, assume that there exists
630: an edge $e$ with a non-zero probability flow and a state $S_{j}$ is
631: connected to this edge.
632: Then, due to the conservation of probability flow at $S_{j}$, there exists
633: another edge connecting $S_{j}$ to another state $S_{i}$ with
634: a non-zero probability flow.
635: The same reasoning for $S_{i}$ leads one to conclude that $S_{i}$ must
636: be connected to a new state $S_{k}$ by an edge with a
637: non-zero probability flow (Fig.~\ref{fig:6}).
638: Repeating the procedure creates a sequence of states
639: $S_{j}\to S_{i}\to S_{k}\to \cdots $, which are connected by
640: edges with non-zero probability flows.
641: Since the number of states is finite, the sequence must visit
642: a state which has already appeared in the sequence.
643: Hence, the graph must contain a loop.
644:
645: % shown in Fig.~\ref{fig:6}
646: \begin{figure}[h]
647: \begin{center}
648: \includegraphics[width=0.4\textwidth]{fig6.eps}
649: \end{center}
650: \caption{Sequence of edges with non-zero probability flow.}
651: \label{fig:6}
652: \end{figure}
653:
654:
655: For a while, we concentrate on the loop-free case
656: where all the probability flows vanish.
657: Take two arbitrary states $S_{i}$ and $S_{j}$.
658: If there are two different paths connecting $S_{i}$ and $S_{j}$,
659: it means that there is a loop composed of a path from $S_{j}$ to
660: $S_{i}$ and one from $S_{i}$ to $S_{j}$.
661: Therefore, if the transition diagram is loop-free, the state
662: $S_{i}$ is connected to $S_{j}$ through a unique path.
663: Let this path be composed of $l$ state:
664: $S_{j}=S_{i_{1}}\to S_{i_{2}}\to \cdots \to S_{i_{l}}=S_{i}$.
665: We can define the transition probability from $S_{j}$ to $S_{i}$ as
666: \begin{equation}
667: \bar{q}_{ij}=q_{ii_{l-1}}q_{i_{l-1}i_{l-2}}\cdots q_{i_{2}j},
668: \label{eq13}
669: \end{equation}
670: where the overbar is used to distinguish the transition probability
671: computed along a path from those between adjacent nodes.
672: In this way, we can compute the transition probability for any state
673: pair $S_{i}$ and $S_{j}$ which are not necessarily adjacent to each other.
674:
675: We can extend the detailed balance (\ref{eq12}) to any pair of states
676: by using the extended transition probabilities (\ref{eq13}).
677:
678: To see this, assume that three states $S_{0}$, $S_{1}$, and $S_{2}$ are
679: connected by a path, i.e., $S_{1}\in {\bf A}_{0}$ and $S_{2} \in {\bf A}_{1}$.
680: Then, the detailed balance (\ref{eq12}) implies
681: $q_{10}p_{0}=q_{01}p_{1}$ and $q_{21}p_{1}=q_{12}p_{2}$.
682: Multiplying each side yields
683: $q_{10}p_{0}q_{21}p_{1}=q_{01}p_{1}q_{12}p_{2}$.
684: Cancelling $p_{1}$ from both sides gives $q_{21}q_{10}p_{0}=q_{01}q_{12}p_{2}$,
685: or equivalently, $\bar{q}_{20}p_{0}=\bar{q}_{02}p_{2}$.
686: This relation suggests that the detailed balance holds
687: in terms of the generalized transition probability (\ref{eq13}) even
688: for state pairs which are not adjacent to each other.
689: It is not difficult to see that this is indeed the case
690: by repeatedly applying it through the unique
691: path connecting the two states, i.e.,
692: \begin{equation}
693: \bar{q}_{ij}p_{j}=\bar{q}_{ji}p_{i},\quad \forall i,j.
694: \label{eq14}
695: \end{equation}
696: The above relation includes the detailed balance (\ref{eq12})
697: as a special case.
698: Therefore, we call the relation (\ref{eq14}) the
699: \textit{generalized detailed balance}.
700: Taking $j=0$ in (\ref{eq14}), we can represent $p_{j}$ as
701: \begin{equation}
702: p_{i}=\bar{r}_{i0}p_{0},
703: \label{eq15}
704: \end{equation}
705: where $\bar{r}_{ij}$ is defined as a generalization of (\ref{eq14a}) with
706: $q_{ij}$ being replaced by $\bar{q}_{ij}$ given by (\ref{eq13}), i.e.,
707: \begin{equation}
708: \bar{r}_{ij}=\frac{\bar{q}_{ij}}{\bar{q}_{ji}}=r_{ii_{l-1}}r_{i_{l-1}i_{l-2}}\cdots r_{i_{2}j}
709: \label{eq16}
710: \end{equation}
711: Here, the overbar is again used to distinguish it from $r_{ij}$ given
712: in (\ref{eq14a}) for state pairs adjacent to each other.
713: $\bar{r}_{ij}$ is also called the transition ratio (TR) from $S_{j}$ to $S_{i}$.
714: Note that the relation (\ref{e24q}) is extended to
715: \[
716: \bar{r}_{ij}=\bar{r}^{-1}_{ji}.
717: \]
718: From the normalization constraint (\ref{eq2-4}) and (\ref{eq15}),
719: the statitonary probability distribution is
720: simply given by
721: \begin{equation}
722: p_{i}=\frac{\bar{r}_{i0}}{\sum^{N}_{j=0}\bar{r}_{j0}},\quad
723: i=0,1,\cdots,N
724: \label{eq17}
725: \end{equation}
726: where $\bar{r}_{00}=1$.
727:
728: We sum up the discussion in this section as follows:\\
729: \textit{If the transition diagram is loop-free,
730: the following facts hold: \\
731: (1) the probability flows vanish at every edge. \\
732: (2) The stationary distribution is given by (\ref{eq17}).\\
733: (3) The generalized detailed balance (\ref{eq14}) holds for
734: the stationary distribution.}
735:
736: \begin{ex}\upshape
737: (Loop-free Transition Diagram)\\
738: Consider a transition diagram of Fig.~\ref{fig:7}, which is loop-free.
739: According to (\ref{eq15}), we have $p_{1}=q_{10}p_{0}$,
740: $p_{2}=r_{21}r_{10}p_{0}$, $p_{3}=r_{31}r_{10}p_{0}$,
741: $p_{4}=r_{43}r_{31}p_{0}$, $p_{5}=r_{50}p_{0}$ with
742: \[
743: p_{0}=\frac{1}{1+r_{10}(1+r_{21}+r_{31}+r_{43}r_{31})+r_{50}}
744: \]
745:
746: % shown in Fig.~\ref{fig:7}
747: \begin{figure}[h]
748: \begin{center}
749: \includegraphics[width=0.3\textwidth]{fig7.eps}
750: \end{center}
751: \caption{A loop-free transition diagram.}
752: \label{fig:7}
753: \end{figure}
754:
755:
756:
757: The above reasoning can be applied to a loop-free subgraph of
758: the general (not loop-free) transition diagram.
759: If a loop-free subgraph is attached to a state $S_{i}$ which
760: belongs to a loop, we can write down the probability of the state
761: of that subgraph according to (\ref{eq17}) with $p_{0}$ being replaced by $p_{i}$.
762: As an example, consider the transition diagram of Fig.~\ref{fig:8}.
763: The state $S_{j}$ is a part of a loop which will be discussed
764: in the next section.
765: If the state $S_{i}$ is not in any loop, then it is a part of a path
766: connecting $S_{i}$ and the terminal state $S_{k}$. Thus,
767: \begin{equation}
768: p_{i}=\bar{r}_{ij}p_{j},
769: \label{eq31new}
770: \end{equation}
771: because the probability flow at any edge on the path connecting
772: $S_{j}$ and $S_{k}$ vanishes.
773: This can be shown by directly applying the argument of this section
774: to the subgraph of Fig.~\ref{fig:8}.
775: \begin{figure}[h]
776: \begin{center}
777: \includegraphics[width=0.4\textwidth]{fig8.eps}
778: \end{center}
779: \caption{A loop-free subgraph.}
780: \label{fig:8}
781: \end{figure}
782:
783:
784: % shown in Fig.~\ref{fig:8}
785: \end{ex}
786:
787: \section{Wegscheider condition}
788: \label{sec4}
789:
790: In the preceding section, we showed that the probability
791: flows vanish if the transition diagram is loop-free.
792: In that case, the stationary distribution is simply calculated from (\ref{eq17}).
793: There are cases where the probability flows vanish
794: even if the transition diagram contains loops.
795:
796: In order to discuss this issue, we consider the case
797: where the whole transition diagram is a loop, as shown in
798: Fig.~\ref{fig:9}(a).
799: The corresponding TR diagram is shown in Fig.~\ref{fig:9}(b).
800: The state transition matrix is given by
801: \begin{equation}
802: \label{eq5-1eq}
803: Q=\begin{bmatrix}D_{0}&q_{01}&0&\cdots&0&q_{0N}\\
804: q_{10}&D_{1}&q_{12}&\cdots&0&0\\
805: 0&q_{21}&D_{2}&\cdots&0&0\\
806: &\cdots&\cdots&&&\\
807: q_{N0}&0&0&\cdots&q_{N,N-1}&D_{N}
808: \end{bmatrix}
809: \end{equation}
810: with $D_{i}=-(q_{i+1,i}+q_{i-1,i})$, $i=1,2,\cdots,N-1$,
811: $D_{0}=-(q_{10}+q_{N0})$, $D_{N}=-(q_{0N}+q_{N-1,N})$.
812: \begin{figure}[t]
813: \begin{center}
814: \includegraphics[width=0.4\textwidth]{fig9.eps}
815: \end{center}
816: \caption{Loop transition Diagram.
817: (a) Transition probability description, (b) transition ratio description.}
818: \label{fig:9}
819: \end{figure}
820:
821:
822: Each state has only two edges, one of which
823: receives the probability flow, the other dispatches it.
824: Due to the conservation of probability flows (\ref{eq11}),
825: the incoming probability flow and the outgoing one are equal.
826: Hence, each edge of the loop has the same amount of probability flow,
827: which is denoted by $\rho$.
828: Taking its direction to be clockwise, that is, taking the
829: probability flow from $S_{i-1}$ to $S_{i}$ to be positive,
830: the probability flow is given by
831: \begin{equation}
832: \rho=q_{i,i-1}p_{i-1}-q_{i-1,i}p_{i},\quad i=1,2,\cdots,N.
833: \label{5-eq29}
834: \end{equation}
835: Thus, we have the following recursion:
836: \begin{align}
837: \label{eq18}
838: p_{i}&=r_{i,i-1}p_{i-1}-\frac{\rho}{q_{i-1,i}},\quad
839: i=1,2,\cdots,N, \\
840: p_{0}&=r_{0N}p_{N}-\frac{\rho}{q_{N,0}}. \nonumber
841: \end{align}
842:
843: In order to obtain a general representation of stationary
844: probabilities, we introduce the numbers $\xi_{0},\xi_{1},\cdots,\xi_{N}$
845: defined sequentially by
846: \begin{align}
847: \xi_{0}&=0 \nonumber\\
848: \xi_{i}&=r_{i,i-1}\xi_{i-1}-\frac{1}{q_{i-1,i}},\quad
849: i=1,2,\cdots,N
850: \label{eq23}
851: \end{align}
852: Then, due to (\ref{eq18}) and (\ref{eq23}),
853: $p_{i}=r_{i,i-1}p_{i-1}+\rho(\xi_{i}-r_{i,i-1}\xi_{i-1})$,
854: Therefore, we have
855: \[ p_{i}-\rho\xi_{i}=r_{i,i-1}(p_{i-1}-\rho\xi_{i-1}), \]
856: which implies $p_{i}-\rho\xi_{i}=\bar{r}_{i0}p_{0}$.
857: Here, $\bar{r}_{i0}=r_{i,i-1}r_{i-1,i-2}\cdots r_{10}$.
858: Thus, $p_{i}$ is represented as
859: \begin{equation}
860: p_{i}=\bar{r}_{i0}p_{0}+\rho \xi_{i}, \quad i=1,\cdots,N.
861: \label{eq24}
862: \end{equation}
863:
864: If $\rho=0$, the above relations are identical to (\ref{eq15}).
865: Thus, equation (\ref{eq24}) represents the stationary
866: distribution as a sum of the probability distribution for the case
867: with zero probability flow and a correction term due to
868: non-zero probability flow.
869: Since $\rho=q_{0N}p_{N}-q_{N0}p_{0}$, the relation (\ref{eq24})
870: for $i=N$ implies $\rho=q_{0N}(\bar{r}_{N0}p_{0}+\rho\xi_{N})-q_{N0}p_{0}=q_{N0}(r_{0N}\bar{r}_{N0}-1)p_{0}+\rho q_{0N}\xi_{N}$.
871: Therefore, we get
872: \begin{equation}
873: \rho(1-q_{0N}\xi_{N})=-(1-\eta)q_{N0}p_{0}
874: \label{eq35eq}
875: \end{equation}
876: where $\eta$ is given by
877: \begin{equation}
878: \eta=r_{0N}\bar{r}_{N0}=r_{0N}r_{N,N-1}\cdots r_{10}
879: \label{eq36eq}
880: \end{equation}
881: Since $\xi_{1}=-1/q_{01}<0$ and $r_{ij}>0$, $\xi_{N}<0$.
882: Hence, $1-q_{0N}\xi_{N}>0$. Therefore, we conclude that
883: the probability flow vanishes, i.e., $\rho=0$, if and only if $\eta=1$.
884:
885: The number $\eta$ is the product of all the
886: TRs along the loop. If we use the analogy between the TR and
887: the equilibrium coefficient in chemical kinetics, the condition
888: $\eta=1$, i.e.,
889: \begin{equation}
890: r_{10}r_{21}\cdots r_{0N}=1,
891: \label{eq37eq}
892: \end{equation}
893: corresponds to the condition derived almost a century ago
894: by Wegscheider. The condition (\ref{eq37eq}) is referred to
895: as the \textit{Wegscheider condition} \cite{karmarkar}.
896: We call the product $\eta$ of (\ref{eq36eq})
897: the \textit{Wegscheider product} for a loop as in Fig.~\ref{fig:9}.
898: The Wegscheider product is in some sense directed.
899: The representation (\ref{eq36eq}) defines the Wegscheider product
900: clockwise in the context of Fig.~\ref{fig:9}.
901: It can also be defined in the opposite direction
902: $\eta'=r_{N0}r_{N-1,N}\cdots r_{01}$.
903: Due to (\ref{e24q}), $\eta'=\eta^{-1}$.
904: Since $\eta=1$ implies $\eta'=1$, the Wegscheider condition can be
905: represented by computing the product in either clockwise
906: or counter clockwise direction.
907:
908: The Wegscheider condition is a property of a loop that
909: requires the transition ratio along the total loop to be unity.
910: This condition is equivalently written in terms of transition
911: probabilities as
912: \begin{equation}
913: q_{10}q_{21}\cdots q_{0N}=q_{N0}q_{N-1,N}\cdots q_{01}.
914: \label{eq21}
915: \end{equation}
916: The left-hand side represents the transition probability from
917: a state to itself along a clockwise contour, while
918: the right-hand side represents that along a counter clockwise contour.
919: The identity (\ref{eq21}) gives a stochastic interpretation of the Wegscheider condition.
920:
921: Taking the normalization condition (\ref{eq2-4}) into account, we have
922: \begin{equation}
923: p_{i}=\frac{\bar{r}_{i0}+\rho\xi_{i}}{\sum^{N}_{j=0}\bigl(\bar{r}_{j0}+\rho\xi_{j} \bigr)},
924: \quad i=0,1,\cdots,N,
925: \label{eq27}
926: \end{equation}
927: where $r_{00}=1$, $\xi_{0}=0$.
928: If the Wegscheider condition holds, then $\rho=0$ and the stationary
929: distribution becomes identical to (\ref{eq17}).
930:
931: The most common method of quantifying the regulatory activities of operons
932: is based on thermal equilibrium theory \cite{berg}\cite{bintu}, which assigns a Gibbs free energy to each binding pattern of the binding sites.
933: This corresponds to assigning a stationary probability distribution
934: a priori, rather than constructing a Markov process by assigning
935: the transition probabilities between states.
936: The transition ratio may then be defined as the ratio between the state probabilities.
937: In that case, the Wegscheider condition obviously holds.
938: To see this, consider a loop $S_{0}\to S_{1}\to S_{2}\to S_{0}$.
939: Then, $r_{10}=p_{1}/p_{0}$, $r_{21}=p_{2}/p_{1}$, $r_{02}=p_{0}/p_{2}$.
940: Thus, $r_{10}r_{21}r_{02}=1$.
941: This proves that the Wegscheider condition is consistent with thermal
942: equilibrium theory.
943: The relation (\ref{eq27}), however, suggests more general types of
944: equilibrium.
945:
946:
947: \section{Edge removal and modifier}
948: \label{sec5}
949:
950: In this section, we consider the meaning of the numbers $\xi_{i}$
951: in (\ref{eq23}) and try to interpret (\ref{eq24}) in the
952: context of edge removal.
953: From (\ref{eq23}), $q_{i-1,i}\xi_{i}-q_{i,i-1}\xi_{i-1}=-1$ and
954: $q_{i,i+1}\xi_{i+1}-q_{i+1,i}\xi_{i}=-1$.
955: These relations yield $-q_{i,i-1}\xi_{i-1}-q_{i,i+1}\xi_{i+1}+(q_{i+1,i}+q_{i-1,i})\xi_{i}=0$.
956: Since $\xi_{0}=0$ and $\xi_{1}=-1/q_{01}$, we have, from (\ref{eq5-1eq}),
957: \begin{equation}
958: Q\xi=\begin{bmatrix}-1\\0\\\vdots\\0\\1\end{bmatrix}(1-q_{0N}\xi_{N}),
959: \label{eq38eq}
960: \end{equation}
961: where $\xi=\begin{bmatrix}\xi_{0}&\xi_{1}&\cdots&\xi_{N}\end{bmatrix}^{T}$.
962: To further investigate the meaning of the vector $\xi$,
963: we consider a transition diagram which is obtained from the
964: original diagram of Fig.~\ref{fig:9} by eliminating the edge
965: connecting $S_{0}$ and $S_{N}$ (Fig.~\ref{fig:10}(a)).
966: The modified transition diagram has no loop, and
967: hence, the stationary distribution is given by $\bar{p}_{i}=\bar{r}_{i0}p_{0}$,
968: following the discussion in the preceding section.
969: The transition matrix $Q'$ corresponding to the modified diagram
970: of Fig.~\ref{fig:10}(a) is obtained by taking $q_{0N}=q_{N0}=0$ in
971: (\ref{eq5-1eq}).
972: From (\ref{eq38eq}), we have
973: \begin{equation}
974: Q'\xi=\begin{bmatrix}-1\\0\\ \vdots\\0\\1\end{bmatrix}.
975: \label{eq39eq}
976: \end{equation}
977: We call the vector $\xi$ satisfying (\ref{eq39eq}) with $\xi_{0}=0$
978: \textit{the modifier} of the loop corresponding to the edge $S_{N}\to S_{0}$.
979: It is computed through the recursion formula (\ref{eq23}).
980: We shall generalize it in the sequel.
981: \begin{figure}[h]
982: \begin{center}
983: \includegraphics[width=0.5\textwidth]{fig10.eps}
984: \end{center}
985: \caption{Reduced transition diagrams.
986: (a) Edge $S_{N}\to S_{0}$ is eliminated, (b) edge $S_{\nu}\to S_{\nu+1}$ is eliminated.}
987: \label{fig:10}
988: \end{figure}
989:
990: If we let $p'$ be the vector whose $(j+1)$th element $p_{i}$
991: is given by (\ref{eq15}), which represents
992: the stationary probability distribution corresponding to
993: the modified loop-free diagram of Fig.~\ref{fig:10}(a),
994: equation (\ref{eq24}) can be written as
995: \begin{equation}
996: p=p'+\rho \xi.
997: \label{eq40eq}
998: \end{equation}
999: This is the general form of the stationary probability
1000: distribution. It holds for a loop transition diagram as well as for general
1001: transition diagram with multiple loops, as is discussed in the next section.
1002: Note that the stationary distribution is represented by a
1003: stationary distribution $p'$ corresponding to the modified loop-free
1004: diagram with a correcting term $\rho\xi$ due to non-zero probability flow.
1005:
1006: Equation (\ref{eq40eq}) is obtained by cutting the edge $S_{N}\to S_{0}$
1007: as in Fig.~\ref{fig:10}(a).
1008: Here, we can show that equation (\ref{eq40eq}) is also
1009: derived by eliminating any edge $S_{\nu}\to S_{\nu+1}$ of
1010: the loop (Fig.~\ref{fig:10}(b)).
1011: Since $\bar{r}_{i0}$ is defined as the product of each TR
1012: along the path connecting $S_{i}$ to $S_{0}$, we have
1013: \begin{equation}
1014: \bar{r}_{i0}=\left\{%
1015: \begin{array}{lll}
1016: r_{10}r_{21}\cdots r_{i,i-1} & \mbox{if} & i\le \nu. \\
1017: r_{N0}r_{N-1,N}\cdots r_{i,i-1} & \mbox{if} & i\ge \nu+1.
1018: \end{array}\right.
1019: \label{eq41eq}
1020: \end{equation}
1021: Now, the modifier is defined as
1022: \begin{align}
1023: &\mbox{[\textit{Forward Recursion}]}\nonumber\\
1024: &\xi_{0}=0 \nonumber \\
1025: &\xi_{i}=r_{i,i-1}\xi_{i-1}-\frac{1}{q_{i-1,i}}, \quad i=1,2,\cdots,\nu
1026: \label{eq42eq}\\
1027: &\mbox{[\textit{Backward Recursion}]}\nonumber\\
1028: &\xi_{N+1}=0. \quad q_{N+1,N}=q_{0N} \nonumber \\
1029: &\xi_{i}=r_{i,i+1}\xi_{i+1}+\frac{1}{q_{i+1,i}}, \quad i=N,N-1,\cdots,\nu+1.
1030: \label{eq44plus}
1031: \end{align}
1032: The recursion formula (\ref{eq42eq}) computes the modifier elements of
1033: the path connecting $S_{0}$ to $S_{\nu}$, while the recursion formula
1034: (\ref{eq44plus}) computes those of the path connecting $S_{0}$ to
1035: $S_{\nu+1}$.
1036: The paths are oppositely directed, but both recursions are
1037: essentially the same except the signs of the added term.
1038: We call (\ref{eq42eq}) \textit{forward recursion}, while (\ref{eq44plus})
1039: \textit{backward recursion}.
1040: It is straightforward to see that
1041: $\xi=\begin{bmatrix}\xi_{0}&\xi_{1}&\cdots&\xi_{N}\end{bmatrix}^{T}$
1042: satisfies
1043: \[
1044: Q\xi=\begin{bmatrix}0&0\\\vdots&\vdots\\0&0\\-q_{\nu+1,\nu}&q_{\nu,\nu+1}\\
1045: q_{\nu+1,\nu}&-q_{\nu,\nu+1}\\0&0\\\vdots&\vdots\\0&0\end{bmatrix}
1046: \begin{bmatrix}\xi_{\nu}\\\xi_{\nu+1}\end{bmatrix}+\begin{bmatrix}0\\\vdots\\0\\1\\-1\\0\\\vdots\\0\end{bmatrix}.
1047: \]
1048: Since $q_{\nu+1,\nu}=q_{\nu,\nu+1}=0$ in the modified transition
1049: diagram, we have
1050: \begin{equation}
1051: \begin{array}{rcll}
1052: \ldelim[{6}{0.1mm}[]&0& \rdelim]{6}{0.1mm}[]\\
1053: &\vdots& \\
1054: Q'\xi= &1& &\cdots \nu+1 \\
1055: &-1& &\cdots \nu+2 \\
1056: &\vdots& \\
1057: &0& \\
1058: \end{array}
1059: \label{eq44eq}
1060: \end{equation}
1061: It is not difficult to see that the stationary distribution is given by
1062: (\ref{eq24}) with $\bar{r}_{i0}$ and $\xi_{i}$ being represented by
1063: (\ref{eq41eq}) and (\ref{eq42eq})(\ref{eq44plus}), respectively,
1064: by applying
1065: the arguments of the preceding section to the forward and
1066: backward recursions separately.
1067:
1068: % shown in Fig.~\ref{fig:9}
1069:
1070: The modifier is extended to any edge $S_{\nu}\to S_{\nu+1}$ and can be
1071: computed through (\ref{eq42eq})(\ref{eq44plus}).
1072: The stationary probability distribution is given in this case by (\ref{eq40eq}),
1073: where $p'$ corresponds to the stationary probability
1074: distribution of the reduced loop-free diagram.
1075: Since $\rho=0$ holds under Wegscheider condition, we obtain the following result:
1076:
1077: \textit{The stationary distribution of a loop is unchanged
1078: if an edge is removed, provided that the Wegscheider condition
1079: (\ref{eq37eq}) or (\ref{eq21}) holds}. \\
1080: In other words, the computation of the stationary distribution
1081: for a loop transition diagram is reduced to the loop-free case
1082: where a very simple form (\ref{eq17}) is already available by eliminating
1083: an edge, provided that the Wegsheider condition holds.
1084: We shall extend this remarkable property to the general transition
1085: diagram with multiple loops in the next section.
1086:
1087: % shown in Fig.~\ref{fig:10}
1088:
1089: \begin{ex}\upshape
1090: (Stationary Distribution of Example 2)\\
1091: This is the case of $N=3$ in the above algorithm.
1092: Remove the edge $e: S_{2}\to S_{3} (\nu=2)$.
1093: The numbers $\xi_{i},i=0,1,2,3,$ are given respectively by
1094: forward recursion
1095: \begin{align}
1096: \xi_{0}&=0, \ \xi_{1}=-\frac{1}{q_{01}},\nonumber \\
1097: \xi_{2}&=-\frac{r_{21}}{q_{01}}-\frac{1}{q_{12}}=-\frac{q_{21}+q_{01}}{q_{01}q_{12}} \label{eq:ex6-48}
1098: \end{align}
1099: and by backward recursion
1100: \[
1101: \xi_{3}=\frac{1}{q_{03}}
1102: \]
1103: The stationary probability distribution is calculated to be
1104: \begin{align}
1105: \label{eq5-1}
1106: p_{1}&=r_{10}p_{0}-\frac{1}{q_{01}}\rho \nonumber\\
1107: p_{2}&=\bar{r}_{20}p_{0}-\rho\xi_{2}=r_{21}r_{10}p_{0}-\frac{q_{01}+q_{21}}{q_{01}q_{12}}\rho \\
1108: p_{3}&=r_{30}p_{0}+\frac{1}{q_{03}}\rho. \nonumber
1109: \end{align}
1110: The Wegscheider product $\eta$ is given by
1111: \[ \eta=r_{03}r_{32}r_{21}r_{10}=\frac{q_{03}q_{32}q_{21}q_{10}}{q_{30}q_{23}q_{12}q_{01}}. \]
1112: The probability flow $\rho=q_{32}p_{2}-q_{23}p_{3}$ is given by
1113: \[ \rho=\frac{q_{03}q_{32}q_{21}q_{10}-q_{30}q_{01}q_{12}q_{23}}{q_{01}q_{12}(q_{03}+q_{32})+q_{03}q_{23}(q_{01}+q_{21})}p_{0} \]
1114: \end{ex}
1115:
1116:
1117: \section{Eliminating loops from the transition diagram and derivation of
1118: the flow equation}
1119: \label{sec6}
1120:
1121: Now, we consider the general case where multiple loops exist,
1122: and derive a method of computing probability flows.
1123: If some edges are removed, the transition diagram becomes loop-free.
1124: Although the selection of such loops is not unique,
1125: the minimum number of such edges required to make the transition
1126: diagram loop-free is unique.
1127: As an example, consider a TR diagram of Fig.~\ref{fig:11}(a).
1128: The elimination of the edges $e_{1}: S_{2}\to S_{3}$ and
1129: $e_{2}: S_{4}\to S_{5}$ makes the diagram loop-free, as is
1130: illustrated in Fig.~\ref{fig:11}(b).
1131: We can eliminate, say, the edges $S_{0}\to S_{1}$ and $S_{3}\to S_{4}$
1132: to make the diagram loop-free.
1133: The elimination of two edges is necessary and
1134: sufficient to make the diagram loop-free in this case.
1135: \begin{figure}[h]
1136: \begin{center}
1137: \includegraphics[width=0.3\textwidth]{fig11.eps}
1138: \end{center}
1139: \caption{An example with three loops.
1140: (a) original diagram, (b) modified diagram.}
1141: \label{fig:11}
1142: \end{figure}
1143:
1144: Now, we recover the eliminated edges in the reduced loop-free
1145: diagram one by one. At each recovery step,
1146: at least one new loop is recovered and we select one such loop.
1147: These loops constitute \textit{basis loops}.
1148: In case of Fig.~\ref{fig:11}, the recovery of edge
1149: $e_{1}: S_{2}\to S_{3}$ recovers the loop $L_{1}: S_{0}\to S_{1}\to S_{2}\to S_{3}\to S_{0}$.
1150: Then, recovery of $e_{2}: S_{4}\to S_{5}$ recovers
1151: $L_{2}: S_{0}\to S_{3}\to S_{4}\to S_{5}\to S_{0}$ and
1152: $L_{3}: S_{0}\to S_{1}\to S_{2}\to S_{3}\to S_{4}\to S_{5}\to S_{0}$.
1153: Thus, we can choose $L_{1}$ and $L_{2}$ as basis loops.
1154: $L_{1}$ and $L_{3}$ can also be basis loops.
1155: In this respect, each edge selected to make the transition
1156: diagram loop-free corresponds to one basis loop.
1157:
1158: Now, let us consider an algebraic representation for edge removal.
1159: Consider an edge $e: S_{j}\to S_{i}$.
1160: The probability flow of the edge $e$ from $S_{j}$ to $S_{i}$
1161: is given by (\ref{eq9}), which is alternatively represented as
1162: \begin{equation}
1163: \rho_{ij}=\sigma (i,j)^{T}p,
1164: \label{eq28}
1165: \end{equation}
1166: where $\sigma (i,j)$ is an $(N+1)$-vector given by
1167: \begin{equation}
1168: \left[\sigma(i,j)\right]_{k}=\left\{%
1169: \begin{array}{lll}
1170: q_{ij}, & k=j+1 \\
1171: -q_{ji}, & k=i+1 \\
1172: 0, & \mbox{otherwise.}
1173: \end{array}\right.
1174: \label{eq29}
1175: \end{equation}
1176: Note that $p_{i}$ appears at the $(i+1)$-th place
1177: instead of at the $i$-th place in the vector $p$ because of
1178: the existence of $p_{0}$ as its first element.
1179: Removal of $e$ from the transition diagram
1180: amounts to removal of $\rho_{ij}$ and $\rho_{ji}=-\rho_{ij}$
1181: from the $(i+1)$-th and the $(j+1)$-th components of $Qp$,
1182: respectively.
1183: Therefore, removal of $e$ corresponds to removal of
1184: the matrix
1185: \begin{equation}
1186: \Delta Q:=-\delta(i,j)\sigma(i,j)^{T},
1187: \label{eq30}
1188: \end{equation}
1189: where $\delta(i,j)$ is an $(N+1)$-dimensional vector
1190: given by
1191: \begin{equation}
1192: \left[\delta(i,j)\right]_{k}=\left\{%
1193: \begin{array}{lll}
1194: -1, & k=i+1 \\
1195: 1, & k=j+1 \\
1196: 0 &
1197: \end{array}\right.
1198: \label{eq31}
1199: \end{equation}
1200: The reduced transition matrix $Q$ is given by
1201: \begin{equation}
1202: Q'=Q-\Delta Q.
1203: \label{eq32}
1204: \end{equation}
1205: Hence, from (\ref{eq30}) and (\ref{eq28}), we have
1206: \[
1207: Q'p=\bigl(Q+\delta (i,j)\sigma(i,j)^{T}\bigr)p=\rho_{ij}\delta(i,j).
1208: \]
1209: Assume that the transition diagram contains $l$ basis loops
1210: $L_{1},L_{2},\cdots,L_{l}$.
1211: We can assume that all these loops contain $p_{0}$
1212: without loss of generality.
1213: Now, choose a set of edges $e_{k}\in L_{k},k=1,2,\cdots,l$,
1214: such that the elimination of $e_{1},e_{2}\cdots,e_{l}$ makes
1215: the transition diagram loop-free.
1216: We assume that the edge $e_{k}$ connects the states
1217: $S_{\mu_{k}}$ and $S_{\nu_{k}}$ in the direction from
1218: $S_{\mu_{k}}$ to $S_{\nu_{k}}$, i.e., $e_{k}: S_{\mu_{k}}\to S_{\nu_{k}}$.
1219: Denote the probability flow associated with $e_{k}$ by $\rho_{k}$,
1220: i.e., $\rho_{k}=\rho_{\nu_{k}\mu_{k}}$.
1221: Then, from (\ref{eq28}), we get
1222: \begin{align}
1223: \rho_{k}&=q_{\nu_{k}\mu_{k}}p_{\mu_{k}}-q_{\mu_{k}\nu_{k}}p_{\nu_{k}} \nonumber\\
1224: &=\sigma(\nu_{k},\mu_{k})^{T}p.
1225: \label{eq6-6}
1226: \end{align}
1227: The transition matrix $Q'$ corresponding to the reduced loop-free
1228: transition diagram is thus
1229: \begin{equation}
1230: Q'=Q+\sum^{l}_{k=1}\delta(\nu_{k},\mu_{k})\sigma(\nu_{k},\mu_{k})^{T}.
1231: \label{eq33}
1232: \end{equation}
1233: Now we define a vector $\xi(\nu_{k},\mu_{k})$ satisfying
1234: \begin{align}
1235: &Q'\xi(\nu_{k},\mu_{k})=\delta(\nu_{k},\mu_{k}), \quad k=1,2,\cdots,l. \nonumber \\
1236: &\xi(\nu_{k},\mu_{k})_{0}=0.
1237: \label{eq34}
1238: \end{align}
1239: We call $\xi(\nu_{k},\mu_{k})$ the \textit{modifier}
1240: of $L_{k}$ corresponding to the edge $S_{\mu_{k}}\to S_{\nu_{k}}$,
1241: which is a generalization of $\xi$ introduced in the preceding section
1242: satisfying (\ref{eq39eq}) or (\ref{eq44eq}).
1243: Since $Q'$ corresponds to a loop-free transition matirx,
1244: its stationary probability distribution $p'$ is given by
1245: \begin{equation}
1246: p'_{i}=\bar{r}_{i0}p_{0}, \quad i=1,2,\cdots,N.
1247: \label{eq35}
1248: \end{equation}
1249: where $\bar{r}_{i0}$ denotes the transition ratio from
1250: $S_{0}$ to $S_{i}$ taken along a unique path connecting
1251: $S_{0}$ to $S_{i}$ in the reduced loop-free transition diagram.
1252: From (\ref{eq33}) and (\ref{eq34}), we get
1253: \[
1254: Q'\bigl(I-\sum^{l}_{k=1}\xi(\nu_{k},\mu_{k})\sigma(\nu_{k},\mu_{k})^{T}\bigr)=Q.
1255: \]
1256: From $Qp=0$,
1257: \[
1258: Q'\bigl(I-\sum^{l}_{k=1}\xi(\nu_{k},\mu_{k})\sigma(\nu_{k},\mu_{k})^{T}\bigr)p=0.
1259: \]
1260: From $Q'p'=0$ and the uniqueness of the stationary distribution,
1261: we have
1262: \[
1263: \bigl(I-\sum^{l}_{k=1}\xi(\nu_{k},\mu_{k})\sigma(\nu_{k},\mu_{k})^{T}\bigr)p=p'.
1264: \]
1265: Taking (\ref{eq28}) into account, we can now
1266: write the stationary probability distribution explicitly as
1267: \begin{equation}
1268: p=p'+\sum^{l}_{k=1}\rho_{k}\xi(\nu_{k},\mu_{k}),
1269: \label{eq36}
1270: \end{equation}
1271: where $p'$ denotes the stationary probability distribution of the reduced
1272: loop-free transition diagram given by (\ref{eq35}).
1273: This generalizes (\ref{eq40eq}).
1274: Equation (\ref{eq36}) clearly shows the importance
1275: of the probability flows reflecting the loop structure of
1276: the transition diagram.
1277: If the probability flows vanish for each loop,
1278: the stationary distribution is identical to that of reduced
1279: loop-free diagram given in (\ref{eq17}).
1280: The modifier $\xi(\nu_{k},\mu_{k})$ represents the dependence of
1281: the stationary distribution on the probability flow.
1282:
1283: It remains to compute the probability flow $\rho_{1},\rho_{2},\cdots,\rho_{l}$.
1284: Premultiplication of (\ref{eq36}) by $\sigma(\nu_{k},\mu_{k})^{T}$
1285: yields
1286: \begin{equation}
1287: \rho_{k}-\sum^{l}_{m=1}\sigma(\nu_{k},\mu_{k})^{T}\xi(\nu_{m},\mu_{m})\rho_{m}=\sigma(\nu_{k},\mu_{k})^{T}p'.
1288: \label{eq37}
1289: \end{equation}
1290: Now, from the definition of $\sigma(\nu_{k},\mu_{k})$ (\ref{eq29}),
1291: we have
1292: \begin{align*}
1293: \sigma(\nu_{k},\mu_{k})^{T}p'&=q_{\nu_{k}\mu_{k}}p'_{\mu_{k}}-q_{\mu_{k}\nu_{k}}p'_{\nu_{k}}\\
1294: &=q_{\mu_{k}\nu_{k}}(-p'_{\nu_{k}}+r_{\nu_{k}\mu_{k}}p'_{\mu_{k}})\\
1295: &=-q_{\mu_{k}\nu_{k}}\bar{r}_{\nu_{k}0}(1-\bar{r}_{0\nu_{k}}r_{\nu_{k}\mu_{k}}\bar{r}_{\mu_{k}0})p_{0}
1296: \end{align*}
1297: As is shown in Fig.~\ref{fig:12}, the term
1298: $\bar{r}_{0\nu_{k}}r_{\nu_{k}\mu_{k}}\bar{r}_{\mu_{k}0}$
1299: corresponds to the product of the transition ratios along $L_{k}$
1300: associated with the edge connecting $S_{\nu_{k}}$ and $S_{\mu_{k}}$,
1301: i.e., the Wegscheider product of $L_{k}$.
1302: We denote it by $\eta_{k}$.
1303: Equation (\ref{eq37}) can now be rewritten as
1304: \begin{equation}
1305: \rho_{k}+\sum^{l}_{m=1}\theta_{km}\rho_{m}=\alpha_{k}(1-\eta_{k})p_{0},
1306: \quad k=1,2,\cdots,l,
1307: \label{eqb62}
1308: \end{equation}
1309: where $\theta_{km}=-\sigma(\nu_{k},\mu_{k})^{T}\xi(\nu_{m},\mu_{m})$,
1310: $\alpha_{k}=-q_{\mu_{k}\nu_{k}}\bar{r}_{\nu_{k}0}$, and $\eta_{k}$ is the Wegscheider product of $L_{k}$.
1311: Equation (\ref{eqb62}) implies that the probability flows can be
1312: computed by solving only a linear equation of $l$ unknowns,
1313: instead of $(N+1)$ unknowns of SSE (\ref{eq2-3}) or (\ref{eq2-4a}).
1314: We call (\ref{eqb62}) the \textit{flow equation}.
1315: \begin{figure}[h]
1316: \begin{center}
1317: \includegraphics[width=0.3\textwidth]{fig12.eps}
1318: \end{center}
1319: \caption{Loop Elimination by Removal of an Edge.}
1320: \label{fig:12}
1321: \end{figure}
1322:
1323: Now, let us summarize the procedure of computing the stationary
1324: probability distribution.
1325: \begin{step}\upshape
1326: Find the minimum number of edges $e_{k}: S_{\mu_{k}}\to S_{\nu_{k}}$,
1327: $k=1,2,\cdots,l$ such that their elimination makes the transition
1328: diagram loop-free.
1329: \end{step}
1330: \begin{step}\upshape
1331: Compute the stationary distribution $p'$ via
1332: (\ref{eq35}) for the reduced loop-free diagram.
1333: \end{step}
1334: \begin{step}\upshape
1335: For each eliminated edge $e_{k}$, associate a loop $L_{k}$,
1336: which is recovered with the edge.
1337: \end{step}
1338: \begin{step}\upshape
1339: \label{step4}
1340: For $e_{k}$, compute the modifier
1341: $\xi(\nu_{k},\mu_{k})$ that satisfies (\ref{eq34}).
1342: \end{step}
1343: \begin{step}\upshape
1344: Solve the flow equation (\ref{eqb62}) to obtain the probability
1345: flows $\rho_{k}$, $k=1,2,\cdots,l$.
1346: \end{step}
1347: \begin{step}\upshape
1348: Compute $p$ from (\ref{eq36}).
1349: \end{step}
1350: \begin{step}\upshape
1351: Compute the exact probability distribution taking the
1352: normalization constraint (\ref{eq2-4}) into account.
1353: \end{step}
1354: The computation of the modifier (Step~\ref{step4}) is discussed
1355: in Appendix.
1356: \begin{ex}\upshape
1357: Consider the transition diagram shown in Fig.~\ref{fig:11}(a) with
1358: 6 states and 3 loops.
1359: To eliminate the loops, two edges, $e_{1}:S_{2}\to S_{3}$
1360: and $e_{2}:S_{4}\to S_{5}$, are removed.
1361: The modified transition diagram is shown in Fig.~\ref{fig:11}(b),
1362: which is loop-free. We choose
1363: $L_{1}:S_{0}\to S_{1}\to S_{2}\to S_{3}\to S_{0}$, which is created
1364: by recovering $e_{1}$ in the modified diagram and
1365: $L_{2}:S_{0}\to S_{3}\to S_{4}\to S_{5}\to S_{0}$ which is created
1366: by recovering $e_{2}$ in the modified diagram as basis loops.
1367: In terms of the notations introduced in this section,
1368: $\mu_{1}=2$, $\nu_{1}=3$, $\mu_{2}=4$, $\nu_{2}=5$,
1369: $\rho_{1}=\rho_{32}$ and $\rho_{2}=\rho_{54}$.
1370: The transition matrices $Q$ and $Q'$ for the original diagram
1371: and the modified diagram are given by
1372: \begin{align*}
1373: &Q=\begin{bmatrix}D_{0}&q_{01}&0&q_{03}&0&q_{05}\\
1374: q_{10}&D_{1}&q_{12}&0&0&0\\
1375: 0&q_{21}&D_{2}&\dashbox(15,10){$q_{23}$}&0&0\\
1376: q_{30}&0&\dashbox(15,10){$q_{32}$}&D_{3}&q_{34}&0\\
1377: 0&0&0&q_{43}&D_{4}&\dashbox(15,10){$q_{45}$}\\
1378: q_{50}&0&0&0&\dashbox(15,10){$q_{54}$}&D_{5}
1379: \end{bmatrix},\\
1380: &Q'=\begin{bmatrix}D_{0}&q_{01}&0&q_{03}&0&q_{05}\\
1381: q_{10}&D_{1}&q_{12}&0&0&0\\
1382: 0&q_{21}&D'_{2}&0&0&0\\
1383: q_{30}&0&0&D'_{3}&q_{34}&0\\
1384: 0&0&0&q_{43}&D'_{4}&0\\
1385: q_{50}&0&0&0&0&D'_{5}
1386: \end{bmatrix}
1387: \end{align*}
1388: where $D_{0}=-(q_{10}+q_{30}+q_{50})$, $D_{1}=-(q_{01}+q_{21})$,
1389: $D_{2}=-(q_{12}+q_{32})$, $D_{3}=-(q_{03}+q_{23}+q_{43})$,
1390: $D_{4}=-(q_{34}+q_{54})$, $D_{5}=-(q_{05}+q_{45})$,
1391: $D'_{2}=-q_{12}$, $D'_{3}=-(q_{03}+q_{43})$, $D'_{4}=-q_{34}$,
1392: $D'_{5}=-q_{05}$.
1393: The elements of $Q$ encircled by dashed boxes are eliminated in $Q'$.
1394: The modifiers, $\xi_{1}$ corresponding to $L_{1}$ for
1395: $e_{1}$ and $\xi_{2}$ corresponding to $L_{2}$ for
1396: $e_{2}$, are given as
1397: \begin{align}
1398: \xi_{1}&=\begin{bmatrix}0\\1/q_{01}\\-r_{21}/q_{01}-1/q_{12}\\
1399: 1/q_{03}\\r_{43}/q_{03}\\0\end{bmatrix}, %\nonumber \\
1400: \xi_{2}&=\begin{bmatrix}0\\0\\0\\-1/q_{03}\\-r_{43}/q_{03}-1/q_{34}\\1/q_{05}\end{bmatrix}.
1401: \label{eq64eq}
1402: \end{align}
1403: The detail of the above calculation is given in Appendix.
1404:
1405: According to (\ref{eqb62}), the flow equation is given by
1406: \begin{align*}
1407: &(1+\theta_{11})\rho_{1}+\theta_{12}\rho_{2}=\alpha_{1}(1-\eta_{1})\\
1408: &\theta_{21}\rho_{1}+(1+\theta_{22})\rho_{2}=\alpha_{2}(1-\eta_{2})
1409: \end{align*}
1410: where $\theta_{ij}=-\sigma^{T}_{i}\xi_{j}$, $i,j=1,2$, with
1411: \begin{align*}
1412: \sigma^{T}_{1}&=\sigma(3,2)^{T}=
1413: \begin{bmatrix}0&0&q_{32}&-q_{23}&0&0\end{bmatrix}\\
1414: \sigma^{T}_{2}&=\sigma(5,4)^{T}=\begin{bmatrix}0&0&0&q_{54}&-q_{45}&0\end{bmatrix}.
1415: \end{align*}
1416: and
1417: \[
1418: \alpha_{1}=-q_{23}\bar{r}_{30},\quad \alpha_{2}=-q_{45}\bar{r}_{50}.
1419: \]
1420: $\eta_{1}$ and $\eta_{2}$ are Wegscheider products
1421: corresponding to $L_{1}$ and $L_{2}$, respectively, and are
1422: given respectively:
1423: \[
1424: \eta_{1}=r_{10}r_{21}r_{32}r_{03}, \quad \eta_{2}=r_{30}r_{43}r_{54}r_{05}.
1425: \]
1426: \end{ex}
1427:
1428: % shown in Fig.~\ref{fig:11}
1429: % shown in Fig.~\ref{fig:12}
1430:
1431: \begin{ex}\upshape
1432: (Enzyme with three effectors)\\
1433: Consider a more complex block transition diagram shown in
1434: Fig.~\ref{fig:13}(a).
1435: This transition diagram describes an allosteric enzyme with
1436: three different effectors or an operon with three transcription factors.
1437: A description like (\ref{ex3-16}) for this example is very
1438: complicated and hence, is omitted.
1439: Now, we choose four edges $e_{1}: S_{1}\to S_{2}$, $e_{2}: S_{4}\to S_{5}$,
1440: $e_{3}: S_{6}\to S_{7}$ and $e_{4}: S_{7}\to S_{8}$ to make the
1441: diagram loop-free.
1442: The reduced diagram is shown in Fig.~\ref{fig:13}(b).
1443: A basis set of loops is composed of the following four loops,
1444: $L_{1}: S_{0}\to S_{1}\to S_{2}\to S_{3}\to S_{0}$,
1445: $L_{2}: S_{0}\to S_{3}\to S_{4}\to S_{5}\to S_{0}$,
1446: $L_{3}: S_{0}\to S_{1}\to S_{6}\to S_{7}\to S_{0}$ and
1447: $L_{4}: S_{0}\to S_{7}\to S_{8}\to S_{5}\to S_{0}$,
1448: as is shown in Fig.~\ref{fig:13}(a).
1449: The stationary probability distribution corresponding to the reduced
1450: loop-free diagram Fig.~\ref{fig:13}(b) is easily calculated according
1451: to (\ref{eq35}), i.e.,
1452: \begin{align*}
1453: &p'_{1}=r_{10}p_{0},\ p'_{2}=\bar{r}_{20}p_{0}=r_{23}r_{30}p_{0},\ p'_{3}=r_{30}p_{0},\\
1454: &p'_{4}=\bar{r}_{40}p_{0}=r_{43}r_{30}p_{0}, \ p'_{5}=r_{50}p_{0},\\
1455: &p'_{6}=\bar{r}_{60}p_{0}=r_{61}r_{10}p_{0},\ r'_{7}=r_{70}p_{0}, \\
1456: &p'_{8}=\bar{r}_{80}p_{0}=r_{85}r_{50}p_{0}.
1457: \end{align*}
1458: The $\sigma $-vectors defined by (\ref{eq29}) in this case are
1459: $\sigma_{1}=\sigma(2,1)$, $\sigma_{2}=\sigma(5,4)$, $\sigma_{3}=\sigma(7,6)$,
1460: $\sigma_{4}=\sigma(8,7)$, i.e.,
1461: \begin{align*}
1462: &\sigma_{1}=\begin{bmatrix}0\\q_{21}\\-q_{12}\\0\\0\\0\\0\\0\\0\end{bmatrix},
1463: \ \sigma_{2}=\begin{bmatrix}0\\0\\0\\0\\q_{54}\\-q_{45}\\0\\0\\0\end{bmatrix},
1464: \ \sigma_{3}=\begin{bmatrix}0\\0\\0\\0\\0\\0\\q_{76}\\-q_{67}\\0\end{bmatrix},%\\
1465: &\sigma_{4}=\begin{bmatrix}0\\0\\0\\0\\0\\0\\0\\q_{87}\\-q_{78}\end{bmatrix}.
1466: \end{align*}
1467: The modifiers $\xi_{i}$, $i=1,2,3,4$ for $L_{1}$, $L_{2}$, $L_{3}$, $L_{4}$
1468: are computed in Appendix; they are
1469: \begin{align}
1470: \xi_{1}&=\begin{bmatrix}0\\-1/q_{01}\\r_{23}/q_{03}+1/q_{32}\\1/q_{03}\\r_{43}/q_{03}\\0\\-r_{61}/q_{01}\\0\\0\end{bmatrix},\
1471: \xi_{2}=\begin{bmatrix}0\\0\\r_{23}/q_{03}\\-1/q_{03}\\-r_{43}/q_{03}-1/q_{34}\\1/q_{05}\\0\\0\\r_{85}/q_{05}\end{bmatrix},\
1472: \nonumber \\
1473: \xi_{3}&=\begin{bmatrix}0\\-1/q_{01}\\0\\0\\0\\0\\-r_{61}/q_{01}-1/q_{16}\\1/q_{07}\\0\end{bmatrix}, \
1474: \xi_{4}=\begin{bmatrix}0\\0\\0\\0\\0\\1/q_{05}\\0\\-1/q_{07}\\r_{85}/q_{05}+1/q_{54}\end{bmatrix}
1475: \label{ex8-63}
1476: \end{align}
1477: The flow equation (\ref{eqb62}) is given by
1478: {\small
1479: \begin{align}
1480: &\begin{bmatrix}
1481: 1+\theta_{11}&\theta_{12}&\theta_{13}&0\\
1482: \theta_{21}&1+\theta_{22}&0&\theta_{24}\\
1483: \theta_{31}&0&1+\theta_{33}&\theta_{34}\\
1484: 0&\theta_{42}&\theta_{43}&1+\theta_{44}\end{bmatrix}
1485: \begin{bmatrix}\rho_{1}\\\rho_{2}\\\rho_{3}\\\rho_{4}\end{bmatrix} \nonumber \\
1486: &=\begin{bmatrix}\alpha_{1}(1-\eta_{1})\\\alpha_{2}(1-\eta_{2})\\\alpha_{3}(1-\eta_{3})\\\alpha_{4}(1-\eta_{4})\end{bmatrix}
1487: \label{ex8-64}
1488: \end{align}
1489: }
1490: where $\theta_{ij}=-\sigma^{T}_{i}\xi_{j}$, $i=1,2,3,4$, $j=1,2,3,4$,
1491: $\alpha_{1}=-q_{12}\bar{r}_{20}$, $\alpha_{2}=-q_{45}\bar{r}_{50}$,
1492: $\alpha_{3}=-q_{67}\bar{r}_{70}$, $\alpha_{4}=-q_{78}\bar{r}_{80}$, and
1493: $\eta_{i}$, $i=1,2,3,4$, are Wegscheider products for $L_{i}$,
1494: and are given by $\eta_{1}=r_{03}r_{32}r_{21}r_{10}$,
1495: $\eta_{2}=r_{05}r_{54}r_{43}r_{30}$, $\eta_{3}=r_{07}r_{76}r_{61}r_{10}$,
1496: and $\eta_{4}=r_{05}r_{58}r_{87}r_{70}$.
1497: The anti-diagonal elements $\theta_{14}$, $\theta_{23}$,
1498: $\theta_{32}$ and $\theta_{41}$ vanish from the loop structure of
1499: Fig.~\ref{fig:13}.
1500:
1501: % shown in Fig.~\ref{fig:13}
1502: \end{ex}
1503: \begin{figure}[h]
1504: \begin{center}
1505: \includegraphics[width=0.25\textwidth]{fig13.eps}
1506: \end{center}
1507: \caption{
1508: An Example of Transition Diagram.
1509: (a) Original Diagram, (b) Reduced Loop-free Diagram.
1510: Transition probabilities are omitted.
1511: }
1512: \label{fig:13}
1513: \end{figure}
1514:
1515:
1516:
1517: \section{Specification of transition probabilities}
1518: \label{sec7}
1519:
1520: In order to obtain a useful representation of the activity of
1521: the biological regulator discussed so far, we must give a way
1522: to connect transition probabilities $q_{ij}$ with more workable
1523: parameters with physical and/or biological meanings.
1524: The state transition caused by binding a BF to a BS and that
1525: caused by releasing a BF from a BS are essentially different
1526: in their nature.
1527: The former transition is usually proportional to the concentration of
1528: the BF to be bound, while the latter does not depend on concentration of
1529: BFs but possibly depends on the occupation pattern
1530: of other BSs.
1531: Therefore, it is natural to assume the following characterization
1532: of transition probabilities:
1533: \begin{equation}
1534: q_{ij}=\left\{%
1535: \begin{array}{ll}
1536: \alpha _{ij}u_{k},& \mbox{if $S_{j}\to S_{i}$ is the binding of $U_{k}$},\\
1537: \alpha_{ij}, & \mbox{if it is a releasing of a BF}
1538: \end{array}\right.
1539: \label{eqary}
1540: \end{equation}
1541: for each $j$ and $i\in {\bf A}_{j}$, where $\alpha_{ij}$ denotes a
1542: coefficient and $u_{k}$ the concentration of $U_{k}$.
1543: The transition probabilities used in Example 1 (equation (\ref{eq2-9}))
1544: and these in Example 2 (equation (\ref{eq2-13}))
1545: are special cases of (\ref{eqary}).
1546:
1547: From the definition of TR in (\ref{eq14a}), we have
1548: \begin{equation}
1549: r_{ij}=\left\{%
1550: \begin{array}{ll}
1551: \beta_{ij}u_{k}, & \mbox{if $S_{j}\to S_{i}$ is the binding of $U_{k}$},\\
1552: \beta_{ij}u^{-1}_{k}, & \mbox{if it is the releasing of $U_{k}$}.
1553: \end{array}\right.
1554: \end{equation}
1555: where $\beta_{ij}=\alpha_{ij}/\alpha_{ji}$.
1556: We can derive a concrete form for the overall activity of our biological
1557: regulator based on (\ref{eqary}),
1558: for the case that the Wegscheider condition holds, i.e., the stationary
1559: distribution given by (\ref{eq17}).
1560:
1561: Let $S_{i}$ be a state where $l$ sites are occupied by
1562: $U_{i_{1}}$, $U_{i_{2}}$, $\cdots$, $U_{i_{l}}$
1563: and the remaining $n-l$ sites are empty.
1564: Then, according to (\ref{eq15}) and (\ref{eqary}),
1565: the probability $p_{i}$ is
1566: $p_{i}=a_{i}u_{i_{1}}u_{i_{2}}\cdots u_{i_{l}}p_{0}$,
1567: where $a_{i}$ denotes the product of $\beta_{ij}s$ associated with
1568: the binding of $U_{i_{1}}$, $U_{i_{2}}$, $\cdots$, $U_{i_{l}}$.
1569: Note that the binding and unbinding of other BFs cannot take
1570: place because the path connecting $S_{0}$ to $S_{i}$ does not contain
1571: any loop.
1572: Now, the normalization constraint (\ref{eq2-4}) yields the
1573: overall activity $\gamma$ defined in (\ref{eq2-5}) as
1574: \begin{equation}
1575: \gamma=\frac{\gamma_{0}+
1576: \Sigma^{N}_{i=1}\gamma_{i}a_{i}u_{i_{1}}u_{i_{2}}
1577: \cdots u_{i_{l}}}
1578: {1+\Sigma^{N}_{i=1}a_{i}u_{i_{1}}u_{i_{2}}
1579: \cdots u_{i_{l}}}.
1580: \label{eq:b}
1581: \end{equation}
1582: This is a familiar form for describing operon
1583: regulation, which appears frequently in the literature e.g., \cite{mochizuki}
1584: \cite{ozbudak}\cite{segel}.
1585: For the Lac operon of Fig.~\ref{fig:2}, the transition probabilities
1586: given by (\ref{eqary}) implies
1587: $q_{10}=\alpha_{10}u_{1}$, $q_{01}=\alpha_{01}$, $q_{21}=\alpha_{21}u_{2}$,
1588: $q_{12}=\alpha_{12}$, $q_{30}=\alpha_{30}u_{2}$, $q_{03}=\alpha_{03}$,
1589: $q_{40}=\alpha_{40}u_{3}$ and $q_{04}=\alpha_{04}$.
1590: Since only the states bound by $U_{1}$ can initiate transcription,
1591: $\gamma_{0}=\gamma_{3}=\gamma_{4}=0$ in (\ref{eq:b}).
1592: Thus, we have
1593: \begin{equation}
1594: \gamma=\frac{
1595: \gamma_{1}K_{1}u_{1}+
1596: \gamma_{2}K_{1}K_{2}u_{1}u_{2}}
1597: {1+K_{1}u_{1}+K_{1}K_{2}u_{1}u_{2}+K_{3}u_{2}+K_{4}u_{3}},
1598: \label{gamfra}
1599: \end{equation}
1600: where $K_{1}=\alpha_{10}/\alpha_{01}$, $K_{2}=\alpha_{21}/\alpha_{12}$,
1601: $K_{3}=\alpha_{30}/\alpha_{03}$ and $K_{4}=\alpha_{40}/\alpha_{04}$.
1602:
1603: The general form (\ref{eq27}), as well as its specification (\ref{eq:b}),
1604: enables us to derive a variety of complicated formulae representing
1605: enzymic actions (e.g., \cite{segel}) almost immediately.
1606: Equation (\ref{eq:b}) also generalizes the classical MWC model \cite{monodj}
1607: to heterotropic cases.
1608:
1609: It is important to notice that the Wegscheider product is
1610: constant and does not depend on concentration of any transcription
1611: factor given the specification (\ref{eqary}).
1612: To see this, notice that
1613: if the loop contains an edge associated with the binding of $U_{k}$,
1614: it must contain an edge associated with unbinding of $U_{k}$,
1615: because the loop must recover the starting state along the path.
1616: If the binding transition contains a TR with $u_{k}$, then it must contain TR
1617: associated with unbinding of $U_{k}$ which contains $u^{-1}_{k}$, so that they
1618: cancel out in the Wegscheider product.
1619: Thus, we have shown that \textit{the Wegscheider product is
1620: always constant} and does not contain the concentrations of BF.
1621:
1622: \section{Conclusions}
1623:
1624: The metabolic process is controlled by enzymes whose expressions are controlled by genetic regulations. On the other hand, genetic regulation is controlled by the products of metabolism that determine the cell state. In this sense, the genetic and metabolic regulations are closely linked together to form a huge and complex network of intracellular regulations. It has been desirable to establish a common framework for quantitatively describing these regulations in a unified way.
1625: For that purpose, we used the analogy between genetic and metabolic regulations shown in Table~\ref{table:1}. The core of the analogy is that the actual computations of the control action are performed through molecular interactions at the regulatory sites between the sites to be bound and the factors to bind or to dissociate.
1626:
1627: We formulated a finite Markov process describing both the genetic and metabolic regulations. The most salient feature of this Markov process is its reciprocity in transition probability; that is, for each pair of state, if the transition probability from one state to another is positive, the opposite transition probability is also positive. This property reflects the reversibility of the molecular interactions we are dealing with, and it is the source of many interesting properties of the Markov process we have formulated. The parameters that determine the action of regulations is represented as the average rate of the stationary probability distribution of that Markov process, based on the assumption that the dynamics of molecular interactions that compute the control actions are much faster than the reaction they are regulating. This assumption is analogous to the fast equilibrium assumption used in deriving the classical Michaelis-Menten equation \cite{segel}. Actually, our approach can derive a stochastic version of Michaelis-Menten formula which suggests its legitimacy. Moreover, we can derive a quantitative estimate of how precise the Michaelis-Menten equation is by supplying the variance of the stationary distribution.
1628:
1629: We introduced a new notion of probability flow associated with each edge of the transition diagram to represent the loop structure of the transition diagram.
1630: The state transition diagram is nothing but a representation of the conservation of probability flows at each node. Based on this observation, we derived many interesting properties of the stationary probability distributions.
1631:
1632: We proved a simple result that the probability flows vanish if the graph has no loop. This immediately implies that the detailed balance holds for pairs of adjacent states. This was generalized to include pairs of states which are not necessarily adjacent to each other. A very simple graphical method of computing the stationary distribution was derived.
1633:
1634: We derived a condition which guarantees the detailed balance even if the transition diagram has a loop. This condition is not new; Wegscheider, a German chemist, discovered this condition a century ago. The condition, which we call the Wegsheider condition in this paper, requires that the product of all the transition ratios around a loop be unity. We combined this condition with the probabilistic flow and directly showed that the probability flows vanish under it.
1635: This result again gives a probabilistic interpretation of this classical result. The detailed balance has been accepted as an obvious fact in literature of chemical kinetics and enzymology, For instance, a famous book of Segel assumed this condition without even mentioning Wegscheider's name \cite{segel}.
1636: We gave a stochastic interpretation of this condition.
1637:
1638: We derived a simple method of computing the stationary probability distribution for general cases with multiple loops. We showed that if the Wegsheider condition holds for a loop, we can eliminate any of the edges contained in the loop without changing the stationary probability distribution. In other words, if the Wegscheider condition holds for a loop, we can consider a transition diagram with that loop by removing an arbitrary edge within that loop. The stationary probability distribution of the reduced diagram is identical to that of the original diagram.
1639:
1640: We derived a purely graphical method of computing the stationary probability distribution based on the notions of probability flow and modifiers. We derived a simple equation for computing the probability flow. Our method dramatically simplifies the classical King and Altman method \cite{king}.
1641:
1642: Our result suggests that there are two kinds of stationary probability
1643: distributions: the one which satisfies the detailed balance,
1644: and one which does not.
1645: The former class is characterized by zero probability flow, the latter
1646: by non-zero probability flows.
1647: In the latter case, the probability distribution is fixed, but continuous
1648: flows of probability exist in loops, which perhaps reflects
1649: a sort of thermal irreversibility of molecular interactions.
1650: Exploitation of the bio-chemical characterization of the probability flow
1651: is an interesting issue of theoretical biology.
1652:
1653: The theoretical framework developed in this paper is based on the
1654: observation that the intracellular regulations are embedded in a
1655: homogeneous computational medium of molecular interactions.
1656: This view is not entirely new, but no serious attempt has been made
1657: so far to mathematically formulate it, within the best of our knowledge.
1658: A finite Markov process model proposed in this paper captures some
1659: essential features of the computational medium and explains the versatility, evolvability and flexibility of the intracellular regulations
1660: and various signal transductions.
1661: It offers a potential capability to deal with systems in which
1662: metabolism and gene expressions are linked and integrated together.
1663: We are now exploiting an analytical tool for investigating
1664: such systems with greater complexity based on our framework
1665: presented in this paper.
1666:
1667:
1668: \begin{thebibliography}{99}
1669:
1670: \bibitem{ackers}
1671: Ackers, G.K., Johnson, A.D., Shea, M.A., 1982.
1672: Quantitative model for gene regulation by $\lambda $ phage repressor.
1673: Proc. Nat. Acad. Sci. 79, 1129-1133.
1674:
1675: \bibitem{beckett}
1676: Beckett, D., 2005.
1677: Multilevel regulation of protein-protein interactions in biological circuitry.
1678: Phys. Biol. 2, S67-73.
1679:
1680: \bibitem{berg}
1681: Berg, O.G., von Hippel, P.H., 1987.
1682: Selection of DNA binding sites by regulatory proteins. Statistical-mechanical theory and application to operators and promoters.
1683: J. Mol. Biol. 193, 723-750.
1684:
1685: \bibitem{bintu}
1686: Bintu, L., Buchler, N.E., Garcia, H.G., Gerland, U., Hwa, T., Kondev, J., Phillips, R., 2005.
1687: Transcriptional regulation by the numbers: models.
1688: Curr. Opin. Genet. Dev. 15, 116-24.
1689:
1690: \bibitem{bliss}
1691: Bliss, R.D., Painter, P.R., Marr, A.G., 1982.
1692: Role of feedback inhibition in stabilizing the classical operon.
1693: J. Theor. Biol. 97, 177-193.
1694:
1695: \bibitem{bolouri}
1696: Bolouri, H., Davidson, E.H., 2002.
1697: Modeling DNA sequence-based cis-regulatory gene networks.
1698: Dev. Biol. 246, 2-13.
1699:
1700: \bibitem{chen}
1701: Chen, K.C., Csikasz-Nagy, A., Gyorffy, B., Val, J., Novak, B., Tyson, J.J.,
1702: 2000. Kinetic analysis of a molecular model of the budding yeast cell cycle.
1703: Mol. Biol. Cell 11, 369-91.
1704:
1705: \bibitem{cherry}
1706: Cherry, J.L., Adler, F.R., 2000.
1707: How to make a biological switch. J. Theor. Biol. 203, 117-133.
1708:
1709: \bibitem{coppey}
1710: Coppey, M., Benichou, O., Voituriez, R., Moreau, M., 2004.
1711: Kinetics of target site localization of a protein on DNA: a stochastic approach.
1712: Biophys. J. 87, 1640-1649.
1713:
1714: \bibitem{dekel}
1715: Dekel, E., Alon, U., 2005.
1716: Optimality and evolutionary tuning of the expression level of a protein.
1717: Nature 436, 588-592.
1718:
1719: \bibitem{endy}
1720: Endy, D., Brent, R., 2001.
1721: Modelling cellular behaviour.
1722: Nature 409, 391-395.
1723:
1724: \bibitem{mca}
1725: Fell, D., 1997.
1726: Understanding the Control of Metabolism.
1727: Portland Press, London.
1728:
1729: \bibitem{freire}
1730: Freire E., 2000.
1731: Can allosteric regulation be predicted from structure?
1732: Proc. Natl. Acad. Sci. USA. 97, 11680-11682.
1733:
1734: \bibitem{gardner}
1735: Gardner, T.S., Cantor, C.R., Collins, J.J., 2000.
1736: Construction of a genetic toggle switch in \textit{Escherichia coli}.
1737: Nature 403, 339-342.
1738:
1739: \bibitem{gibson}
1740: Gibson, K.M., Giren, J.A., Head, M.S., 1997.
1741: A new class of models for computing receptor-ligand binding affinities.
1742: Chem. Biol. 4, 87-92.
1743:
1744: \bibitem{gilles}
1745: Gillespie, D.T., 1977.
1746: Stochastic simulation of coupled chemical reactions.
1747: J. Phys. Chem. 81, 2341-2361.
1748:
1749: \bibitem{griffith}
1750: Griffith, J.S., 1968.
1751: Mathematics of cellular control processes. I. Negative feedback
1752: to one gene, II. Positive feedback to one gene.
1753: J. Theor. Biol. 20, 202-208, 209-16.
1754:
1755: \bibitem{gunasekaran}
1756: Gunasekaran, K., Ma, B., Nussinov, R., 2004.
1757: Is allostery an intrinsic property of all dynamic proteins?
1758: Proteins 57, 433-43.
1759:
1760: \bibitem{heinrich}
1761: Heinrich, H., Schuster, S., 1996.
1762: The regulation of Cellular Systems.
1763: Chapman and Hall, New York.
1764:
1765: \bibitem{istrail}
1766: Istrail, S., Davidson, E.H., 2005.
1767: Logic functions of the genomic cis-regulatory code.
1768: Proc. Natl. Acad. Sci. USA. 102, 4954-4959.
1769:
1770: \bibitem{kaern}
1771: Kaern, M., Elston, T.C., Blake, W.J., Collins, J.J., 2005.
1772: Stochasticity in gene expression: from theories to phenotypes.
1773: Nat. Rev. Genetics 6, 451-464.
1774:
1775: \bibitem{karmarkar}
1776: Karmarkar, R., Bose, I., 2004.
1777: Graded and binary responses in stochastic gene expressions.
1778: Phys. Biol. 1, 197-204.
1779:
1780: \bibitem{killer}
1781: Killer, A.D., 1995. Model genetic circuit encoding autoregulatory
1782: transcription factors.
1783: J. Theor. Biol. 172, 169-185.
1784:
1785: \bibitem{king}
1786: King, E.L., Altman, C., 1956.
1787: A schematic method of deriving the rate laws for enzyme-catalyzed reactions.
1788: J. Phys. Chem. 60, 1375-1378.
1789:
1790: \bibitem{kitano}
1791: Kitano, H., 2004. Biological robustness.
1792: Nat. Rev. Genet. 5, 826-837.
1793:
1794: \bibitem{lloyd}
1795: Lloyd, G., Landini, P., Busby, S., 2001.
1796: Activation and repression of transcription initiation in bacteria.
1797: Essays Biochem. 37, 17-31.
1798:
1799: \bibitem{mackey}
1800: Mackey, M.C., Santillan, M., Yildirim, N., 2004.
1801: Modeling operon dynamics: the tryptophan and lactose operons as paradigms.
1802: C.R. Biol. 327, 211-24.
1803:
1804: \bibitem{mcadams}
1805: McAdams, H.H., Arkin, A., 1997.
1806: Stochastic mechanisms in gene expression.
1807: Proc. Nat. Acad. Sci. 94, 814-819.
1808:
1809: \bibitem{ming}
1810: Ming, D., Wall, M.E., 2005.
1811: Quantifying allosteric effects in proteins.
1812: Proteins 59, 697-707.
1813:
1814: \bibitem{mochizuki}
1815: Mochizuki, A., 2005.
1816: An analytical study of the number of steady states in gene regulatory networks.
1817: J. Theor. Biol. 236, 291-310.
1818:
1819: \bibitem{monod}
1820: Monod, J., 1972. Chance and Necessity, Paperback.
1821:
1822: \bibitem{monodj}
1823: Monod, J., Wyman, J., Changeux, J.P., 1965.
1824: On the nature of allosteric transitions: a plausible model.
1825: J. Mol. Biol. 12, 88-118.
1826:
1827: \bibitem{munsky}
1828: Munsky, B., Khammash, M., 2006.
1829: The finite state projection algorithm for the solution of the chemical
1830: master equation.
1831: J. Chem. Phys. 124, 044104
1832:
1833: \bibitem{ozbudak}
1834: Ozbudak, E.M., Thattal, M., Lim, H.M., Shralman, B.I., von Oudenaarden, A.,
1835: 2004. Multistability in the lactose utilization network of Escherichia coli.
1836: Nature 427, 737-740
1837:
1838: \bibitem{panh}
1839: Pan, H., Lee, J.C., Hilser, V.J., 2000.
1840: Binding sites in Escherichia coli dihydrofolate reductase communicate by modulating
1841: the conformational ensemble.
1842: Proc. Natl. Acad. Sci. USA 97, 12020-12025.
1843:
1844: \bibitem{han}
1845: Perutz, M.F., 1989.
1846: Mechanisms of cooperativity and allosteric regulation in proteins.
1847: Q. Rev. Biophys. 22, 139-237.
1848:
1849: \bibitem{ptashne}
1850: Ptashne, M., 1992.
1851: A Genetic Switch, Cell \& Blackwell Scientific, Cambridge, MA.
1852:
1853: \bibitem{santillan}
1854: Santillan, M., Mackey, M., 2004.
1855: Influence of catabolite repression and inducer exclusion on the bistable behavior of the lac operon.
1856: Biophys. J. 86, 75-84.
1857:
1858: \bibitem{savageau}
1859: Savageau, M.A., 1985. A theory of alternative designs for
1860: biochemical control systems.
1861: Biomed. Biochim. Acta. 44, 875-880.
1862:
1863: \bibitem{segel}
1864: Segel, I.H., 1993.
1865: Enzyme Kinetics: Behaviour and analysis of rapid equilibrium and
1866: steady-state enzyme systems, John Wiley and Sons, INC., New York.
1867:
1868: \bibitem{smolen}
1869: Smolen, P., Boxter, D.A., Byrne, J.H., 2000.
1870: Modeling transcriptional control in gene networks--methods, recent results, and future directions.
1871: Bull. Math. Biol. 62, 247-292.
1872:
1873: \bibitem{tanaka}
1874: Tanaka, R.J., Okano, H., Kimura, H., 2006.
1875: Mathematical description of gene regulatory units.
1876: Biophys. J. 91, 1235-1247.
1877:
1878: \bibitem{ordenaa}
1879: Thattai M., van Oudenaarden, A., 2001.
1880: Intrinsic noise in gene regulatory networks.
1881: Proc. Natl. Acad. Sci. USA 98, 8614-8619.
1882:
1883: \bibitem{thomas}
1884: Thomas, R., Thieffry, D. Kauffman, M., 1995.
1885: Dynamical behaviour of biological regulatory network -- I.
1886: Biological role of feedback loops and practical use of the concept
1887: of the loop-characteristics state.
1888: Bull. Math. Biol. 57, 247-276.
1889:
1890: \bibitem{tyson}
1891: Tyson, J.J., 1978.
1892: The dynamics of feedback control circuits in biochemical pathways.
1893: Prog. Theol. Biol. 5, 1-61.
1894:
1895: \bibitem{kampen}
1896: van Kampen, N.G., 1992.
1897: Stochastic Process in Physics and Chemistry. Elsevier Science, Amsterdam.
1898:
1899: \bibitem{wolfd}
1900: Wolf, D.M., Eeckman, F.H., 1998.
1901: On the relationship between genomic regulatory element organization and gene regulatory dynamics.
1902: J. Theor. Biol. 195, 167-186.
1903:
1904:
1905: \bibitem{wong}
1906: Wong, P., Gladney, S., Keasling, J.D., 1997.
1907: Mathematical model of the lac operon: inducer exclusion, catabolite repression, and diauxic growth on glucose and lactose.
1908: Biotechnol. Prog. 13, 132-143.
1909:
1910: \bibitem{wyman}
1911: Wyman, J., 1968.
1912: Linked functions and reciprocal effects in hemoglobin: A second look.
1913: Adv. Protein Chem. 19, 223-286.
1914:
1915: \bibitem{yanofsky}
1916: Yanofsky, C., 1992. Transcriptional regulation: Elegance in design
1917: and discovery. Transcription Regulation I.
1918: Cold Spring Harbor Lab. Press, pp.3-24.
1919:
1920: \bibitem{yi}
1921: Yi, T.M., Huang, Y., Simon, M.I., Doyle, J., 2000.
1922: Robust perfect adaptation in bacterial chemotaxis through integral feedback control.
1923: Proc. Natl. Acad. Sci. USA. 97, 4649-4653.
1924:
1925: \bibitem{Zaslaver}
1926: Zaslaver, A., Mayo, A.E., Rosenberg, R., Bashkin, P., Sberro, H., Tsalyuk, M., Surette, M.G., Alon, U., 2004
1927: Just-in-time transcription program in metabolic pathways.
1928: Nat. Genet.36, 486-491.
1929:
1930: \end{thebibliography}
1931:
1932:
1933: \section*{Appendix}
1934:
1935: Computation of the modifier can be done through the simple recursions
1936: presented in Section~\ref{sec4}.
1937: The modifier corresponding to the loop $L_{k}$ is an
1938: $(N+1)$-dimensional vector $\xi_{k}=\xi(\nu_{k},\mu_{k})$ satisfying
1939: (\ref{eq34}).
1940: For the case that the transition diagram itself is a loop,
1941: we have seen that the recursions formulae (\ref{eq42eq}) and (\ref{eq44plus})
1942: give a vector $\xi$ that satisfies (\ref{eq44eq}),
1943: which corresponds to the special case $\mu_{k}=\nu$, $\nu_{k}=\nu+1$.
1944: The computation of the modifier for a general loop is essentially
1945: reduced to (\ref{eq42eq})(\ref{eq44plus}).
1946: To avoid notational complication,
1947: we assume that $L_{k}$ is the loop $S_{0}\to S_{1}\to S_{2}\to \cdots \to S_{\pi}\to S_{0}$.
1948: Since $e_{k}: S_{\mu_{k}}\to S_{\nu_{k}}$ is in the loop,
1949: we assume $\mu_{k}=\nu$, $\nu_{k}=\nu+1$, $\nu\le \pi-1 $.
1950: Then, the components of the modifier corresponding to the states inside
1951: the loop $L_{k}$ are given by (\ref{eq42eq}) and
1952: (\ref{eq44plus}).
1953: Thus, we have a procedure for computing the components of
1954: the modifier corresponding to the states inside the loop.
1955: The components of $\xi_{k}$ outside the loop
1956: $L_{k}$ are computed as follows: let $S_{j}$ be a state outside
1957: $L_{k}$. Since the reduced diagram is loop-free,
1958: there exists a unique path connecting $S_{j}$ to $S_{0}$.
1959: If the path does not have a common state with $L_{k}$ except
1960: $S_{0}$, let the component of $\xi_{k}$ corresponding to
1961: $S_{j}$ be zero.
1962: If the path has some states in common with $L_{k}$, there is a
1963: state $S_{k_{j}}$ nearest $S_{j}$ in $L_{k}$.
1964: Then, the component of $\xi_{k}$ corresponding to $S_{j}$ is
1965: given by
1966: \[
1967: (\xi_{k})_{j}=\bar{r}_{jk_{j}}\xi_{k_{j}} \tag{A1}
1968: \]
1969: where $\bar{r}_{jk_{j}}$ is the TR from the state $S_{k_{j}}$ to
1970: $S_{j}$ and $\xi_{k_{j}}$ is the component of the modifier
1971: corresponding to $S_{k_{j}}$ which has already been computed.
1972: The justification of (A1) is based on the remark at the end of
1973: Section~\ref{sec3} (equation (\ref{eq31new})).
1974:
1975: \vspace*{5mm}
1976: \noindent
1977: {\bf Example A1.}
1978: We compute $\xi_{1}$ and $\xi_{2}$ in Fig.~\ref{fig:11}.
1979: For $L_{1}$, the forward recursion corresponding to
1980: (\ref{eq42eq}) is given by
1981: \begin{align*}
1982: &\xi_{10}=0 \\
1983: &\xi_{11}=-\frac{1}{q_{01}} \\
1984: &\xi_{12}=r_{21}\xi_{11}-\frac{1}{q_{12}}=-r_{21}\frac{1}{q_{01}}-\frac{1}{q_{12}}
1985: \end{align*}
1986: and the backward recursion corresponding to (\ref{eq44plus})
1987: is given by
1988: \[ \xi_{13}=\frac{1}{q_{03}}. \]
1989: From the graph of Fig.~\ref{fig:11}(b), the path connecting
1990: $S_{4}$ to $S_{0}$ meets $L_{1}$ at $S_{3}$,
1991: while the path connecting $S_{5}$ to $S_{0}$ directly reaches $S_{0}$.
1992: Hence, we have
1993: \[
1994: \xi_{14}=r_{43}\xi_{13}=\frac{r_{43}}{q_{03}},\quad \xi_{15}=0.
1995: \]
1996: For $L_{2}$, the forward recursion is given by
1997: \begin{align*}
1998: &\xi_{20}=0 \\
1999: &\xi_{23}=-\frac{1}{q_{03}}\\
2000: &\xi_{24}=r_{43}\xi_{23}-\frac{1}{q_{34}}=-\frac{r_{43}}{q_{03}}-\frac{1}{q_{34}}
2001: \end{align*}
2002: and the backward recursion is given by
2003: \[ \xi_{25}=\frac{1}{q_{05}}. \]
2004: For the states outside $L_{2}$, $S_{2}$ and $S_{1}$ are directly
2005: connected to $S_{0}$ without meeting $L_{2}$.
2006: Hence, $\xi_{22}=\xi_{21}=0$.
2007: Thus, (\ref{eq64eq}) has been confirmed.
2008:
2009: \vspace*{5mm}
2010: \noindent
2011: {\bf Example A2.}
2012: We compute $\xi_{1}$, $\xi_{2}$, $\xi_{3}$ and $\xi_{4}$
2013: corresponding to the loops $L_{1}$, $L_{2}$, $L_{3}$ and $L_{4}$ of Example 8.
2014:
2015: For $L_{1}$,
2016: the forward recursion $(S_{0}\to S_{1})$ gives
2017: \[
2018: \xi_{11}=-1/q_{01},
2019: \]
2020: and the backward recursion $(S_{0}\to S_{3}\to S_{2})$
2021: gives
2022: \begin{eqnarray*}
2023: \xi_{13}&=&1/q_{03},\\
2024: \xi_{12}&=&r_{23}\xi_{13}+1/q_{32}.
2025: \end{eqnarray*}
2026: The probabilities of the states outside $L_{1}$ are
2027: given by
2028: \begin{eqnarray*}
2029: \xi_{14}&=&r_{43}\xi_{13},\\
2030: \xi_{1i}&=&\bar{r}_{i0}\xi_{10}=0,\, i=5,6,7,8.
2031: \end{eqnarray*}
2032:
2033: For $L_{2}$, the
2034: forward recursion $(S_{0}\to S_{3}\to S_{4})$
2035: gives
2036: \begin{eqnarray*}
2037: \xi_{23}=-1/q_{03},\\
2038: \xi_{24}=r_{43}\xi_{23}-1/q_{34},
2039: \end{eqnarray*}
2040: and the backward recursion $(S_{0}\to S_{5})$
2041: gives
2042: \[\xi_{25}=1/q_{05}.\]
2043: The probabilities outside $L_{2}$
2044: are given by
2045: \begin{eqnarray*}
2046: \xi_{22}&=&r_{23}\xi_{23}, \\
2047: \xi_{28}&=&r_{85}\xi_{25}, \\
2048: \xi_{2i}&=&\bar{r}_{i0}\xi_{20}=0,\, i=1,6,7.
2049: \end{eqnarray*}
2050:
2051: For $L_{3}$, the forward recursion
2052: $(S_{0}\to S_{1}\to S_{6})$ gives
2053: \begin{eqnarray*}
2054: \xi_{31}&=&-1/q_{01},\\
2055: \xi_{32}&=&r_{41}\xi_{31}-1/q_{14},
2056: \end{eqnarray*}
2057: and the
2058: backward recursion $(S_{0}\to S_{7})$
2059: gives
2060: \[\xi_{37}=1/q_{07}\]. The probabilities outside $L_{3}$
2061: is given by
2062: \[\xi_{3i}=\bar{r}_{i0}\xi_{20}=0,\, i=2,3,4,5,8\].
2063:
2064: For $L_{4}$, the
2065: forward recursion $(S_{0}\to S_{7})$
2066: gives \[\xi_{47}=-1/q_{07}\], and the
2067: backward recursion $(S_{0}\to S_{5}\to S_{8})$
2068: gives
2069: \begin{eqnarray*}
2070: \xi_{45}&=&1/q_{05},\\
2071: \xi_{48}&=&r_{85}\xi_{45}+1/q_{54}.
2072: \end{eqnarray*}
2073: The probabilities outside $L_{4}$
2074: is given by \[\xi_{4i}=\bar{r}_{i0}\xi_{40}=0, \, i=1,2,3,4,6.\]
2075:
2076: The above scheme verifies (\ref{ex8-63}).
2077:
2078:
2079:
2080:
2081:
2082:
2083:
2084:
2085:
2086: \end{document}