q-bio0512041/noise.tex
1: \documentclass[twocolumn,pre,superscriptaddress,floatfix,amsmath]{revtex4}
2: %\documentclass[twocolumn,prl,superscriptaddress,showpacs,floatfix]{revtex4}
3: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: \usepackage{graphicx}
5: 
6: \newcommand{\myav}[1]{\langle{#1}\rangle}
7: \newcommand{\mymat}[1]{{\mathbf #1}}
8: \newcommand{\myorder}[1]{\{#1\}}
9: \newcommand{\zero}{\myorder{0}}
10: \newcommand{\one}{\myorder{1}}
11: \newcommand{\mysum}[1]{\underset{\scriptstyle\alpha\in #1}{\textstyle\sum}}
12: \newcommand{\sumzero}{\mysum\zero}
13: \newcommand{\sumone}{\mysum\one}
14: \newcommand{\torate}[1]{\overset{#1}{\to}}
15: \newcommand{\tofromrate}[2]{\underset{#2}{\overset{#1}{\rightleftharpoons}}}
16: %\newcommand{\torate}[1]{\stackrel{#1}{\longrightarrow}}
17: %\newcommand{\longtorate}[1]{\overset{#1}{\longrightarrow}}
18: \newcommand{\latin}[1]{\emph{#1}}
19: \newcommand{\species}[1]{\emph{#1}}
20: %\newcommand{\eqref}[1]{(\ref{#1})}
21: 
22: \begin{document}
23: 
24: \title{Exact results for noise power spectra in linear biochemical
25: reaction networks}
26: 
27: \author{Patrick B. Warren}
28: %
29: \affiliation{Unilever R\&D Port Sunlight, Bebington, Wirral, CH63 3JW, UK.}
30: %
31: \author{Sorin {T\u anasa-Nicola}}
32: %
33: \affiliation{FOM Institute for Atomic and Molecular Physics,
34: Kruislaan 407, 1098 SJ Amsterdam, The Netherlands.}
35: %
36: \author{Pieter Rein ten Wolde}
37: %
38: \affiliation{FOM Institute for Atomic and Molecular Physics,
39: Kruislaan 407, 1098 SJ Amsterdam, The Netherlands.}
40: 
41: \date{December 2005}
42: 
43: \begin{abstract}
44: %
45: We present a simple method for determining the exact noise power
46: spectra in linear chemical reaction networks.  We apply the method to
47: networks which are representative of biochemical processes such as
48: gene expression and signal detection.  Our results clarify how noise
49: is transmitted by signal detection motifs, and indicate how to
50: coarse-grain networks by the elimination of fast reactions.
51: %
52: \end{abstract}
53: 
54: \maketitle
55: 
56: \section{Introduction}
57: %
58: \label{sec:intro}
59: %
60: It is now clear that chemical noise can sometimes play a significant
61: role in the functioning of biochemical reaction networks \cite{RWA}.
62: Such noise is implicated in the spontaneous flipping of genetic
63: switches \cite{ARMcA}, and has been shown to have an adverse effect on
64: the functioning of a synthetic chemical oscillator \cite{Elowitz}.  In
65: some circumstances it has also been argued that noise can have a
66: beneficial effect, for example in stochastic focusing \cite{PBE}.
67: 
68: Chemical noise in reaction networks can be characterised by the root
69: mean square (rms) deviation of number of molecules from the mean.
70: Noise is important if the rms deviation is a significant fraction of
71: the mean.  This is typically associated with systems where the mean
72: number of molecules is rather small.  Such situations are
73: characteristic of gene regulatory networks where the concentrations of
74: regulatory proteins can be small enough for there to be only a few
75: molecules in the cell volume.  For example, for the lac operon in
76: \species{E. coli}, the lac repressor is active at concentrations where
77: there are only 10--20 molecules present in the cell volume
78: \cite{SanMAc}.
79: 
80: The mean and the rms deviation are `point statistics', and have
81: frequently been used to characterise the stochastic properties of
82: biochemical reaction networks.  In the present paper, we focus on the
83: noise power spectra, which are a more refined characterisation of the
84: stochastic properties.  We also focus on linear reaction networks for
85: which the noise power spectra and point statistics can be calculated
86: exactly.  In the next section, we develop a general theory for
87: computing the noise power spectra in linear reaction networks.  We
88: relegate technical details to the Appendix.  In the following sections
89: we apply the general theory to increasingly elaborate networks which
90: model basic processes of biochemical interest, such as gene expression
91: and signal detection.
92: 
93: \section{General theory}
94: \label{sec:gen}
95: %
96: We define a \emph{linear reaction network} to be a network of chemical
97: reactions which does not involve bimolecular or higher order reactions.
98: For such a network, the chemical rate equations can be written as
99: %
100: \begin{equation}
101: \frac{d N_i}{dt} = {\textstyle\sum_j}K_{ij} N_j + b_i\label{eq:cre}
102: \end{equation}
103: %
104: where $N_i$ is the number of molecules of species $i$, $K_{ij}$ is a
105: matrix of rate coefficients, and $b_i$ are source terms.
106: Eq.~\eqref{eq:cre} is a set of linear ordinary differential equations,
107: which explains the origin of the phrase `linear reaction network'.
108: Appendix \ref{sec:appkb} indicates how $K_{ij}$ and $b_i$ are related
109: to the stoichiometry matrix and reaction rates which specify the
110: network.
111: 
112: Let $N_i(t)$ be the instaneous number of molecules of species $i$ in
113: the system.  Define the mean value of $N_i$ in steady state to be
114: $\myav{N_i}$, and the mean value out of steady state to be
115: $\myav{N_i}_t$.  By taking moments of the chemical master equation
116: (see Appendix), one can prove that
117: %
118: \begin{equation}
119: \frac{d\myav{N_i}_t}{dt} = {\textstyle\sum_j}K_{ij}\myav{N_i}_t 
120: + b_i.\label{eq:nit} 
121: \end{equation}
122: %
123: Therefore, the chemical rate equations are \emph{exact} for a linear
124: reaction network, provided we work with the mean quantities
125: $\myav{N_i}_t$.  In steady state, the mean values therefore solve
126: %
127: \begin{equation}
128: {\textstyle\sum_j}K_{ij}\myav{N_i} + b_i = 0.\label{eq:ni}
129: \end{equation}
130: %
131: This is a system of linear simultaneous equations which can be readily
132: solved. 
133: 
134: We now define the deviation away from the steady state mean values to be
135: %
136: \begin{equation}
137: \Delta N_i(t) = N_i(t) - \myav{N_i}
138: \end{equation}
139: %
140: and introduce the steady-state variance-covariance matrix
141: %
142: \begin{equation}
143: S_{ij}=\myav{\Delta N_i\,\Delta N_j}.
144: \end{equation}
145: %
146: The diagonal elements of this are the variances
147: %
148: \begin{equation}
149: \sigma_i^2=S_{ii}=\myav{\Delta N_i^2}.
150: \end{equation}
151: %
152: The rms deviation mentioned in the introduction is $\sigma_i$.  It is
153: often the case that $\sigma_i^2\sim\myav{N_i}$ and in fact some
154: authors define the `noise strength' to be $\sigma_i^2/\myav{N_i}$
155: although we will not explicitly use this in the present paper.  This
156: scaling means that the relative rms deviation $\sigma_i/\myav{N_i}\sim
157: 1/\surd\myav{N_i}$ which indicates that the relative important of
158: noise decreases inversely with the square root of the number of
159: molecules involved.  It is the basic reason why noise is important for
160: systems where the number of molecules is small.
161: 
162: As many authors have noticed \cite{Berg, McAA, TvO}, the
163: variance-covariance matrix at steady state can also be obtained by
164: taking moments of the chemical master equation.  The details are
165: unimportant for our arguments, and are relegated to Appendix
166: \ref{sec:appkb}.  For the remainder of the main text, we assume that the
167: means and variance-covariance matrix have been computed.  We will say
168: that these comprise the \emph{point statistics} of the network.
169: 
170: To refine the description of the steady state beyond the point
171: statistics, we introduce the set of correlation functions
172: %
173: \begin{equation}
174: C_{ij}(t)=\myav{\Delta N_i(0)\,\Delta N_j(t)}.
175: \end{equation}
176: %
177: In steady state, the value chosen for the time origin is unimportant
178: (therefore the average could be regarded as a time average over
179: starting times, instead of an ensemble average).  The correlation
180: functions have the properties
181: %
182: \begin{equation}
183: \begin{array}{l}
184: C_{ij}(0) \to S_{ij}\quad\text{as}\quad t\to0,\\[6pt]
185: C_{ij}(t) \to 0\quad\text{as}\quad t\to\infty,\\[6pt]
186: C_{ij}(t) = C_{ji}(-t)
187: \end{array}
188: \end{equation}
189: %
190: (the first of these is not always true; see section \ref{sec:fast}).
191: The last property shows that the autocorrelation functions are
192: time-symmetric, $C_{ii}(t) = C_{ii}(-t)$, but it does not follow that
193: the cross-correlation functions are time-symmetric, as is illustrated
194: by the example in section \ref{sec:gem}.
195: 
196: Closely related to the autocorrelation functions are the noise power
197: spectra, which are the central topic of study in the present paper.
198: These are defined to be
199: %
200: \begin{equation}
201: P_i(\omega) = \myav{|\Delta N_i(\omega)|^2} 
202: = 2{\int_0^\infty}\!\!dt\,
203: \cos\omega t \,C_{ii}(t)\label{eq:power}
204: \end{equation}
205: %
206: where $\Delta N_i(\omega)$ is the Fourier transform of $\Delta
207: N_i(t)$.  We shall use the second form in Eq.~\eqref{eq:power} to
208: compute the noise power spectra.  
209: 
210: Usually the noise power spectra obey the sum rule (for an exception,
211: see section \ref{sec:fast}),
212: %
213: \begin{equation}
214: \sigma_i^2 = \frac{1}{2\pi} {\int_0^\infty}\!\!
215: d\omega\,P_i(\omega).\label{eq:sumrule}
216: \end{equation}
217: 
218: Now we turn to the computation of the correlation functions and the
219: noise power spectra.  In Appendix \ref{sec:appc} we prove that the
220: correlation functions satisfy
221: %
222: \begin{equation}
223: \frac{dC_{ki}}{dt} = {\textstyle\sum_j}K_{ij}C_{kj}.\label{eq:corr} 
224: \end{equation}
225: %
226: If we subtract Eq.~\eqref{eq:ni} from Eq.~\eqref{eq:nit}, we find that
227: the mean deviation out of steady state obeys
228: %
229: \begin{equation}
230: \frac{d\myav{\Delta N_i}_t}{dt} 
231: = {\textstyle\sum_j}K_{ij}\myav{\Delta N_i}_t.\label{eq:dnit} 
232: \end{equation}
233: %
234: This shows that the correlation functions decay in exactly the same
235: way as deviations away from steady state, so that Eq.~\eqref{eq:corr}
236: is an example of a regression theorem \cite{gardinerbook}.
237: 
238: Eq.~\eqref{eq:corr} is to be solved with the initial conditions
239: $C_{ij}(0) = S_{ij}$.  A convenient approach is by the use of
240: the Laplace transform, which allows the initial conditions to be
241: automatically incorporated.  Taking the Laplace transform of
242: Eq.~\eqref{eq:corr} shows that
243: %
244: \begin{equation}
245: s \tilde C_{ki} - S_{ki} 
246: = {\textstyle\sum_j}K_{ij}\tilde C_{kj}\label{eq:lcorr}
247: \end{equation}
248: %
249: where
250: %
251: \begin{equation}
252: \tilde C_{ij}(s) = {\int_0^\infty}\!\!dt\,e^{-st}
253: C_{ij}(t).\label{eq:laplace}
254: \end{equation}
255: %
256: Eq.~\eqref{eq:lcorr} is a set of linear simultaneous equations which
257: can readily be solved for $\tilde C_{ij}(s)$.
258: 
259: The power spectra now follow almost for free.  Comparing
260: Eq.~\eqref{eq:laplace} with Eq.~\eqref{eq:power} shows that
261: %
262: \begin{equation}
263: P_i(\omega) = \tilde C_{ii}(i\omega) + \tilde C_{ii}(-i\omega)
264: ,\label{eq:l2p}
265: \end{equation}
266: %
267: in other words $P_i(\omega)$ is twice the real part of the analytic
268: continuation of $\tilde C_{ii}(s)$ to $s = i\omega$.
269: 
270: Eqs.~\eqref{eq:lcorr} and \eqref{eq:l2p} are the key results of this
271: section.  They indicate how exact results for the power spectra can be
272: obtained for an arbitrary linear reaction network by purely algebraic
273: methods.  In Appendix \ref{sec:appl} we prove that the same results
274: can be obtained from the chemical Langevin equations
275: described by Gillespie \cite{GillCLE}.
276: 
277: It is interesting to note that the power spectra are simpler to obtain
278: than the correlation functions themselves, which require an inverse
279: Fourier or Laplace transform.  To undertake such a step requires in
280: general that a polynomial be factorised, which is not always possible.
281: In fact the route to the power spectra, via Eqs~\eqref{eq:ni},
282: \eqref{eq:lcorr} and \eqref{eq:app:vcm} in the Appendix, just involves
283: the solution of linear simultaneous equations.
284: 
285: \begin{figure}
286: %
287: \begin{center}
288: \includegraphics{fig1.eps}
289: \end{center}
290: %
291: \caption[?]{Power spectra for the gene expression models in
292: Eq.~\eqref{eq:A} (Eq.~\eqref{eq:pnk}; dashed line) and
293: Eq.~\eqref{eq:MA} (Eq.~\eqref{eq:MApa}; solid line).  Parameters for
294: the explicit mRNA model of Eq.~\eqref{eq:MA} are $k = 2.76\,
295: \mathrm{min}^{-1}$, $\lambda = 0.12\, \mathrm{min}^{-1}$, $\rho =
296: 3.2\, \mathrm{min}^{-1}$, and $\gamma = 0.016\, \mathrm{min}^{-1}$.
297: These correspond to cro in a recent model of phage lambda
298: \cite{SanMac2}.  Parameters for the implicit mRNA model of
299: Eq.~\eqref{eq:A} are $k = 1.36\, \mathrm{min}^{-1}$, $n = 48.1$,
300: $\gamma = 0.0142\, \mathrm{min}^{-1}$.  These are chosen so that the
301: two models have matched point statistics, $\myav{N_A} = 4.6\times10^3$
302: and $\sigma_A / \myav{N_A} = 0.07$, and matching values of $P_A(0) =
303: 1.6\times10^7\,\mathrm{min}$.\label{fig:fig1}}
304: %
305: \end{figure}
306: 
307: \section{Applications}
308: %
309: \subsection{Gene expression models}
310: \label{sec:gem}
311: %
312: We first consider a very simple model of gene expression
313: %
314: \begin{equation}
315: \torate{k}n\,\text{A},\qquad \text{A}\torate{\gamma}
316: \label{eq:A}
317: \end{equation}
318: %
319: which represents a birth-death process in which $n$
320: copies of A are generated each time the first (birth) reaction fires.
321: This represents the `burstiness' of gene expression in prokaryotes
322: \cite{TvO}.  The point statistics are
323: %
324: \begin{equation}
325: \myav{N_A}=\frac{nk}{\gamma},\qquad\sigma_A^2=\frac{n+1}{2}\myav{N_A}
326: \label{eq:nk}
327: \end{equation}
328: %
329: The chemical rate equation is $dN_A/dt = nk - \gamma N_A$ so the 
330: autocorrelation function obeys $dC_{AA}/dt = - \gamma C_{AA}$.  The
331: solution, and its Laplace transform, are 
332: %
333: \begin{equation}
334: C_{AA}(t)=\sigma_A^2 e^{-\gamma t},\quad
335: \tilde C_{AA}(s)=\frac{\sigma_A^2}{s+\gamma}.
336: \end{equation}
337: %
338: The corresponding power spectrum is
339: %
340: \begin{equation}
341: P_A(\omega)=\frac{2\gamma\sigma_A^2}{\omega^2+\gamma^2}.\label{eq:pnk}
342: \end{equation}
343: 
344: Now we turn to a more realistic representation of gene expression
345: which explicitly includes mRNA, namely
346: %
347: \begin{equation}
348: \torate{k}\text{M},\quad \text{M}\torate{\lambda},
349: \quad \text{M}\torate{\rho}\text{M}+\text{A},\quad
350: \text{A}\torate{\gamma}.\label{eq:MA}
351: \end{equation}
352: %
353: The first two reactions represent generation and degradation of M
354: which corresponds to mRNA.  The second two reactions represent
355: generation of the protein A from M and the degradation of A.  This is
356: a model which has been analysed by other workers \cite{Berg, McAA,
357: TvO}, and we recover previously known results for the point statistics.
358: 
359: The mean values in steady state are
360: %
361: \begin{equation}
362: \myav{N_M}=\frac{k}{\lambda},\quad
363: \myav{N_A}=\frac{\rho}{\gamma}\myav{N_M}=\frac{bk}{\gamma}.
364: \label{eq:bk}
365: \end{equation}
366: %
367: In this, $b=\rho/\lambda$ is the mean number of copies of A produced
368: in the lifetime of one mRNA and can be taken to represent the
369: `burstiness' of gene expression.  Comparing the last of
370: Eq.~\eqref{eq:bk} with the first of Eq.~\eqref{eq:nk}, it appears that
371: $b$ in the present model corresponds to $n$ in the previous implicit
372: mRNA model.  As we shall see though, this interpretation does not
373: carry through to the other statistical properties.
374: 
375: The steady state variances and covariance are 
376: (see Appendix \ref{sec:appkb}) 
377: %
378: \begin{equation}
379: \begin{array}{l}
380: \displaystyle
381: \sigma_M^2=\myav{N_M},\quad
382: \sigma_A^2=\Bigl(1+\frac{\rho}{\gamma+\lambda}\Bigr)\,\myav{N_A},\\[18pt]
383: \displaystyle
384: S_{M\!A}=\frac{\rho\myav{N_M}}{\gamma+\lambda}
385: =\frac{\gamma\myav{N_A}}{\gamma+\lambda}.\label{eq:MAvcm}
386: \end{array}
387: \end{equation}
388: %
389: In this case, comparing Eq.~\eqref{eq:MAvcm} with Eq.~\eqref{eq:nk},
390: we see that the correspondence $b\leftrightarrow n$ fails for
391: $\sigma_A^2$.
392: 
393: Now we turn to the computation of the correlation functions and the
394: power spectra.  The chemical rate equations are 
395: %
396: \begin{equation}
397: \frac{d N_M}{dt} = k - \lambda N_M,\quad
398: \frac{d N_A}{dt} = \rho N_M - \gamma N_A.
399: \end{equation}
400: %
401: The correlation functions obey the homogeneous form of these
402: equations.  Taking the Laplace transform as indicated in the previous
403: section, we find
404: %
405: \begin{equation}
406: \begin{array}{l}
407: s\tilde C_{M\!M} - \sigma_M^2 = - \lambda\tilde C_{M\!M},\\[6pt]
408: s\tilde C_{AM} - S_{AM} = - \lambda\tilde C_{AM},\\[6pt]
409: s\tilde C_{M\!A} - S_{M\!A} = \rho \tilde C_{M\!M}
410:  - \gamma \tilde C_{M\!A},\\[6pt]
411: s\tilde C_{AA} - \sigma_A^2 = \rho \tilde C_{AM} - \gamma \tilde C_{AA}.
412: \end{array}
413: \end{equation}
414: %
415: Because $C_{M\!A}(t)\ne C_{AM}(t)$, there are four different
416: Laplace-transformed correlation functions (note though that
417: $S_{M\!A}=S_{AM}$).  The solutions are
418: %
419: \begin{equation}
420: \begin{array}{l}
421: \displaystyle
422: \tilde C_{M\!M}=\frac{\sigma_M^2}{s+\lambda},\qquad
423: \tilde C_{AM}=\frac{S_{AM}}{s+\lambda},\\[12pt]
424: \displaystyle
425: \tilde C_{M\!A}=\frac{S_{M\!A}}{s+\gamma}+
426: \frac{\rho\sigma_M^2}{(s+\lambda)(s+\gamma)},\\[12pt]
427: \displaystyle
428: \tilde C_{AA}=\frac{\sigma_A^2}{s+\gamma}+
429: \frac{\rho S_{AM}}{(s+\lambda)(s+\gamma)}.
430: \end{array}
431: \end{equation}
432: %
433: In this particular case, the inverse Laplace transforms can easily be
434: found by the method of partial fractions.  We give only the result
435: for the cross-correlation function, which exhibits the interesting time
436: asymmetry mentioned above,
437: %
438: \begin{equation}
439: C_{M\!A}(t)=
440: %\Bigl\{
441: \left\{
442: \begin{array}{ll}
443: \displaystyle 
444: S_{M\!A} \,e^{-\gamma t}+
445: \rho\sigma_M^2\,
446: \frac{e^{-\gamma t}-e^{-\lambda t}}{\lambda-\gamma}
447: & (t\ge0)\\
448: \displaystyle 
449: S_{M\!A} \,e^{-\lambda|t|} 
450: \phantom{\frac{e^{-\gamma t}-e^{-\lambda t}}{\lambda-\gamma}}
451: & (t\le0)
452: \end{array}
453: \right.
454: \end{equation}
455: %
456: (using $C_{M\!A}(-t)=C_{AM}(t)$ for the second line).
457: 
458: Finally, the power spectra are obtained.  The power spectrum for M is
459: $P_M(\omega) = {2\lambda\myav{N_M}}/{(\omega^2+\lambda^2)}$.  This
460: might have been anticipated as a special case of the previous model
461: with $n=1$, since as far as M is concerned it is undergoing a simple
462: birth-death process.  The power spectrum for A is
463: %
464: \begin{equation}
465: P_A(\omega)=\frac{2\gamma\myav{N_A}(\omega^2+\lambda(\lambda+\rho))}%
466: {(\omega^2+\lambda^2)(\omega^2+\gamma^2)}.\label{eq:MApa}
467: \end{equation}
468: %
469: With a suitable choice of $n$ and rate cofficients $k$ and $\mu$, the
470: point statistics of the previous model in Eq.~\eqref{eq:A} could be
471: matched to the present model in Eq.~\eqref{eq:MA}.  However, it is in
472: general impossible to match the power spectra since the $\omega$
473: dependence is different.  Fig.~\ref{fig:fig1} shows the power spectra
474: for the two models with matched point statistics.  The presence of two
475: correlation times in the explicit mRNA model compared to a single
476: relaxation time for the implicit mRNA model is clearly seen.
477: 
478: We have gone through the calculations in some detail for these two
479: model, but for the subsequent calculations we will only give the final
480: results.
481: 
482: \subsection{Signal detection}
483: %
484: A very simple model of signal detection is the reaction
485: %
486: \begin{equation}
487: \text{A}\torate{\nu}\text{B},\quad\quad 
488: \text{B}\torate{\mu}\label{eq:AB}
489: \end{equation}
490: %
491: where A is the input signal and B is the output signal.
492: Eq.~\eqref{eq:AB} has a chemical rate equation
493: %
494: \begin{equation}
495: \frac{d N_B}{dt}=\nu N_A-\mu N_B+\eta.\label{eq:ABcle}
496: \end{equation}
497: %
498: We have included an additional noise term $\eta$ in this, so it is
499: actually a chemical Langevin equation.  Taking the Fourier transform
500: yields
501: %
502: \begin{equation}
503: i\omega N_B=\nu N_A-\mu N_B + \eta,
504: \quad\text{or}\quad
505: N_B=\frac{\nu N_A+\eta}{i\omega+\mu}.
506: \end{equation}
507: %
508: The mean square modulus is then
509: %
510: \begin{equation}
511: \myav{|N_B|^2}=\frac{\nu^2\myav{|N_A|^2}+\myav{|\eta|^2}}
512: {\omega^2+\mu^2}.
513: \end{equation}
514: %
515: We have assumed that $\eta$ is uncorrelated with $N_A$.  The theory of
516: chemical Langevin equations \cite{GillCLE} (see also Appendix
517: \ref{sec:appl}) shows that $\myav{|\eta|^2} = 2\mu\myav{N_B}$, and for
518: $\omega>0$ one has $\myav{|N_i|^2} = \myav{|\Delta N_i|^2}$, thus
519: finally
520: %
521: \begin{equation}
522: P_B(\omega)=\frac{\nu^2}{\omega^2+\mu^2}\, P_A(\omega)
523: +\frac{2\mu\myav{N_B}}{\omega^2+\mu^2}.
524: \label{eq:psf}
525: \end{equation}
526: %
527: This result has quite a striking interpretation.  The total noise in
528: the output signal is made up of an \emph{extrinsic} contribution
529: (first term) plus an \emph{intrinsic} contribution (second term).
530: Moreover, the extrinsic noise is equal the input signal noise
531: multiplied by a low-pass filter function.  This is analogous to the
532: behaviour of a passive RC circuit element \cite{SAR}.  This result, in
533: one form or another, was derived by Paulsson \cite{Paulsson}, and by
534: Shibata and Fujimoto \cite{Shibata}.  We refer to it as the `spectral
535: addition rule'.  It is a potentially important result because it
536: indicates that there is a trade-off between signal amplification by
537: the detection motif and signal contamination by added intrinsic noise.
538: 
539: Our exact results for linear reaction networks enable us to test the
540: spectral addition rule.  We find that its validity is limited by the
541: assumption that the intrinsic noise $\eta$ is uncorrelated with the
542: input signal $N_A$.  In particular, if the detection motif consumes
543: molecules of the input signal, a correlation can arise which spoils
544: the spectral addition rule.
545: 
546: To demonstrate this, we now consider the effect of adjoining
547: Eq.~\eqref{eq:AB} onto the gene expression models in the previous
548: section.  We present results for the explicit mRNA model, but similar
549: conclusions can be drawn for the implicit mRNA model too \cite{TNWtW}.
550: We therefore solve
551: %
552: \begin{equation}
553: \torate{k}\text{M}\torate{\lambda},
554: \quad \text{M}\torate{\rho}\text{M}+\text{A},\quad
555: \text{A}\torate{\gamma},\quad
556: \text{A}\torate{\nu}\text{B}\torate{\mu}.
557: \label{eq:MAB}
558: \end{equation}
559: %
560: We find that
561: %
562: \begin{equation}
563: \frac{P_A(\omega)}{\myav{N_A}}=
564: \frac{2(\gamma+\nu)(\omega^2+\lambda(\lambda+\rho))}
565: {(\omega^2+\lambda^2)(\omega^2+(\gamma+\nu)^2)}
566: \end{equation}
567: %
568: and
569: %
570: \begin{equation}
571: \frac{P_B(\omega)}{\myav{N_B}}=
572: \frac{2\mu[(\omega^2+\lambda^2)(\omega^2+(\gamma+\nu)^2)
573: +\lambda\nu\rho(\gamma+\nu)]}
574: {(\omega^2+\lambda^2)(\omega^2+(\gamma+\nu)^2)(\omega^2+\mu^2)}
575: \end{equation}
576: %
577: with $\nu\myav{N_A}=\mu\myav{N_B}$.  It is not hard to demonstrate
578: that the spectral addition rule \emph{fails} for these expressions.
579: 
580: As an aside, one can show $\sigma_B^2/\myav{N_B} <
581: \sigma_A^2/\myav{N_A}$.  Since the $\text{A}\to\text{B}$ reaction in
582: Eq.~\eqref{eq:MAB} can be regarded as post-translational modification
583: step, this shows that post-translational modification can reduce the
584: noise associated with gene expression.  The reason is that the
585: post-translational modification reaction smoothes the `burstiness' of
586: gene expression by acting as a low-pass filter.
587: 
588: To appreciate the influence of coupling the detection reaction to the
589: input signal noise, we now change the detection scheme so that the
590: signal molecules A are \emph{not} consumed, by replacing
591: Eq.~\eqref{eq:AB} with
592: %
593: \begin{equation}
594: \text{A}\torate{\nu}\text{A}+\text{B},\quad\quad 
595: \text{B}\torate{\mu}.\label{eq:AAB}
596: \end{equation}
597: %
598: As far as B is concerned, it is important to note that the \emph{same}
599: chemical Langevin equation Eq.~\eqref{eq:ABcle} holds for this
600: detection scheme as for the previous scheme.  For this variant we find
601: $P_A(\omega)$ is as given in Eq.~\eqref{eq:MApa} and
602: %
603: \begin{equation}
604: \frac{P_B(\omega)}{\myav{N_B}}=
605: \frac{2\mu[(\omega^2+\lambda^2)(\omega^2+\gamma(\gamma+\nu))
606: +\gamma\nu\rho\lambda]}
607: {(\omega^2+\gamma^2)(\omega^2+\lambda^2)(\omega^2+\mu^2)}
608: \label{eq:pBMAB}
609: \end{equation}
610: %
611: One can check that in this case $P_A(\omega)$ and $P_B(\omega)$ do
612: obey the spectral addition rule.
613: 
614: Whilst we have only demonstrated the failure of the spectral addition rule
615: for one particular case, the result is suggestive of the general
616: conclusion that the spectral addition rule will only hold for detection
617: schemes which do not consume input signal molecules.  These
618: conclusions can also be reached by analysing the chemical Langevin
619: equations for the whole network \cite{TNWtW}.
620: 
621: \subsection{Fast reactions}
622: %
623: \label{sec:fast}
624: %
625: There is a growing literature on the adiabatic elimination of fast
626: reactions for stochastic chemical kinetics \cite{Kepler01,
627: Haseltine02, Shibata03a, Shibata03, Bundschuh, Rao03, Roussel04,
628: Chatterjee05, Cao05, Straube05}.  For example Kepler and Elston
629: \cite{Kepler01}, and Shibata \cite{Shibata03a, Shibata03}, have a
630: formalism that permits a systematic approach to the problem, involving
631: the identification of the fast and slow variables in the system.  In a
632: similar example, Bundschuh \latin{et al} present simulation results
633: which support the general strategy of elimination of fast variables in
634: terms of slow variables \cite{Bundschuh}.  Here we examine the
635: procedure for elimination of a fast equilibration reaction in the
636: context of the \emph{exact} results for a linear reaction network.
637: Our results shed light on the way in which fast equilibration
638: reactions can be systematically eliminated from a reaction network,
639: whilst preserving the noise attributes.
640: 
641: To make a concrete example, we suppose that the signal detection motif
642: of the previous section consists of a fast
643: binding-unbinding step, followed by a slower detection step.  We
644: replace $\text{A}\to\text{B}$ by
645: $\text{A}\rightleftharpoons\text{A}^*\to\text{B}$ where A$^*$
646: represents the bound state.  We suppose that the substrate which binds
647: A is in excess, so that our reaction network is still a linear
648: network.  The reaction scheme in its totality is now
649: %
650: \begin{equation}
651: \begin{array}{ll}
652: \torate{k}\text{M}\torate{\lambda},
653: &{}\quad \text{M}\torate{\rho}\text{M}+\text{A},\\[6pt]
654: \text{A}\torate{\gamma},
655: &{}\quad
656: \text{A}\tofromrate{k_{\!f}}{k_b}\text{A}^*\torate{\nu}\text{B}\torate{\mu}.
657: \end{array}
658: \label{eq:MAAsB}
659: \end{equation}
660: %
661: The noise statistics for this network can be solved but the
662: expressions are rather lengthy.  We therefore focus on the limit of a
663: fast binding-unbinding reaction.  
664: 
665: The equilibrium constant $K$ for the binding-unbinding reaction is
666: defined via
667: %
668: \begin{equation}
669: {k_{\!f}}=K\,{k_b}.
670: \end{equation}
671: %
672: To take the limit of fast equilibration, we keep $K$ finite and allow
673: $k_b\to\infty$.  The results are as follows.  Let us first consider
674: the final product species B.  We find $P_B(\omega)$ is given by
675: Eq.~\eqref{eq:pBMAB} but with `renormalised' reaction rates
676: %
677: \begin{equation}
678: \gamma_R=\frac{\gamma}{K+1},\quad\nu_R=\frac{K\nu}{K+1}
679: \label{eq:renorm}
680: \end{equation}
681: %
682: The same is true of the variance $\sigma_B^2$.  
683: 
684: We can rationalise this as follows.  The chemical rate equations which
685: involve the species in the fast equilibration reaction (A and A$^*$)
686: are
687: %
688: \begin{equation}
689: \begin{array}{l}
690: dN_A/dt = \rho N_M-(\gamma+k_{\!f}) N_A+k_b N_{A^*},\\[6pt]
691: dN_{A^*}/dt = k_{\!f}N_A-(\nu+k_b)N_{A^*},\\[6pt]
692: dN_B/dt = \nu N_{A^*}-\mu N_B.
693: \end{array}
694: \end{equation}
695: %
696: We now make a linear transformation from $N_A$ and $N_{A^*}$ to new
697: variables $N_X$ and $N_Y$ :
698: %
699: \begin{equation}
700: \begin{array}{l}
701: N_X = N_A + N_{A^*},\\[6pt]
702: N_Y = K N_A - N_{A^*}.
703: \end{array}
704: \label{eq:lintran}
705: \end{equation}
706: %
707: These are chosen because $N_X$ is a \emph{slow} variable which is
708: conserved by the equilibration reaction, and $N_Y$ is a \emph{fast}
709: variable which vanishes in steady state.  In terms of these
710: concentration variables, the chemical rate equations become
711: %
712: \begin{equation}
713: \begin{array}{l}
714: \displaystyle
715: dN_X/dt = \rho N_M - \frac{\gamma+K\nu}{K+1}\,N_X-
716: \frac{\gamma-\nu}{K+1}\,N_Y,\\[12pt]
717: \displaystyle
718: dN_Y/dt +(K+1) k_b N_Y = K\rho N_M-\frac{\gamma-\nu}{K+1}\,N_X,\\[12pt]
719: \displaystyle
720: dN_B/dt = \frac{K\nu}{K+1}\,N_X-\frac{\nu}{K+1}\,N_Y-\mu N_B.
721: \end{array}
722: \label{eq:cXYB}
723: \end{equation}
724: %
725: The second of these shows that $N_Y$ relaxes at a rate $\sim k_b$ to a
726: constant $\sim 1/k_b$, for large $k_b$.  This formalises the
727: separation of timescales between $N_Y$ and all the other concentration
728: variables.  In the large $k_b$ limit therefore, $N_Y=0$ and
729: Eqs.~\eqref{eq:cXYB} become
730: %
731: \begin{equation}
732: \begin{array}{l}
733: dN_X/dt = \rho N_M - (\gamma_R+\nu_R) N_X,\\[6pt]
734: dN_B/dt = \nu_R N_X - \mu N_B.
735: \end{array}
736: \end{equation}
737: %
738: with the reaction rates given by Eq.~\eqref{eq:renorm}.  These are (a
739: subset of) the chemical rate equations for the following reaction
740: network
741: %
742: \begin{equation}
743: \torate{k}\text{M}\torate{\lambda},
744: \quad \text{M}\torate{\rho}\text{M}+\text{X},\quad
745: \text{X}\torate{\gamma_R},\quad
746: \text{X}\torate{\nu_R}\text{B}\torate{\mu}.
747: \label{eq:MXB}
748: \end{equation}
749: %
750: This is the same as the previously discussed scheme in
751: Eq.~\eqref{eq:MAB}, with the replacement $\text{A}\leftrightarrow
752: \text{X}$.  The power spectra follow from Eqs.~\eqref{eq:pBMAB}
753: accordingly.  The important point to note is that this scheme not only
754: has the correct chemical rate equations but also has the correct noise
755: power spectra.
756: 
757: Now let us turn to the power spectra for A and A$^*$ in the original
758: network.  We find that
759: %
760: \begin{equation}
761: P_A(\omega) = \frac{1}{(K+1)^2}\,P_X(\omega),\quad
762: P_{A^*}(\omega) = K^2 P_A(\omega).\label{eq:pAAs}
763: \end{equation}
764: %
765: Again these results are easy to rationalise, since inverting
766: Eqs.~\eqref{eq:lintran} and setting $N_Y = 0$ shows that 
767: %
768: \begin{equation}
769: N_A = \frac{1}{K+1}\,N_X,\quad
770: N_{A^*} = K\,N_A
771: \end{equation}
772: %
773: (the proportionality factors become squared when computing the noise
774: power spectra).
775: 
776: For the variances, $\sigma_A^2$ and $\sigma_{A^*}^2$, we find a
777: different story though.  In fact, in the limit $k_b\to\infty$, we
778: find a \emph{breakdown} of the sum rule in Eq.~\eqref{eq:sumrule}.
779: The reason is that
780: %
781: \begin{equation}
782: \lim_{k_b\to\infty} \int_0^\infty\!\!d\omega\,
783: P_i(\omega)\ne
784: \int_0^\infty\!\!d\omega\,
785: \lim_{k_b\to\infty} P_i(\omega)
786: \label{eq:limorder}
787: \end{equation}
788: %
789: for $i=\text{A}$ and $\text{A}^*$.  The sum rule always holds for the
790: left hand side of Eq.~\eqref{eq:limorder}, but the limiting power
791: spectra are defined in terms of the right hand side of
792: Eq.~\eqref{eq:limorder}.  (As an aside, we find that the sum rule is
793: satisfied for $P_{A+A^*}(\omega)$, which is equal to $P_X(\omega)$,
794: confirming that the effect is confined to the individual species in
795: the fast equilibration reaction.)  To handle this kind of situation,
796: we define a `sum rule deficit'
797: %
798: \begin{equation}
799: \Delta\sigma^2_i \equiv \sigma_i^2-\frac{1}{2\pi}\int_0^\infty\!\!d\omega\,
800: P_i(\omega).
801: \end{equation}
802: %
803: For the present situation, we find that 
804: %
805: \begin{equation}
806: \Delta\sigma^2_A = \Delta\sigma^2_{A^*} = \frac{\myav{N_{A^*}}}{K+1}.
807: \end{equation}
808: %
809: Interestingly, the same result is found for \emph{all} situations
810: involving a fast equilibration reaction which we have examined.
811: 
812: The origin of the sum rule deficit lies in the noise generated by the
813: fast equilibration reaction.  Formally, at any finite $k_b$ we can
814: define the contribution of the equilibration reaction to the total
815: noise of species $i$ to be
816: %
817: \begin{equation}
818: \Delta P_i(\omega)=P_i(\omega)-\lim_{k_b\to\infty} P_i(\omega).
819: \end{equation}
820: %
821: For $i=\text{A}$ and A$^*$, one has that $\Delta P_i(\omega)\to0$ as
822: $k_b\to\infty$, at any finite value of $\omega$, but that
823: $\int_0^\infty \!\Delta P_i(\omega)\to 2\pi\Delta\sigma^2_i$.  The
824: reason is that, whilst $\Delta P_i(\omega)$ vanishes as $1/k_b$, it
825: extends over a frequency range $0<\omega\alt k_b$, so the total
826: integrated contribution does not vanish in the limit $k_b\to\infty$.
827: In words, the equilibration reaction contributes high-frequency noise
828: which vanishes at any particular finite frequency in the limit of
829: infinitely fast equilibration, but makes a net non-zero contribution
830: to the total integrated noise.
831: 
832: The universal magnitude of the sum rule deficit can be understood by
833: analysing the simplest of all equilibration reactions, namely
834: %
835: \begin{equation}
836: \text{A}\tofromrate{k_{\!f}}{k_b}\text{A}^*.\label{eq:AAs}
837: \end{equation}
838: %
839: Solving for the point statistics in this system one has $K\myav{N_A} =
840: \myav{N_{A^*}}$ and, from Eqs.~\eqref{eq:app:vcm} in the Appendix, 
841: %
842: \begin{equation}
843: \begin{array}{l}
844: -K\sigma_A^2+S_{AA^*}+\myav{N_{A^*}}=0\\[6pt]
845: -K S_{AA^*}+\sigma_{A^*}^2
846: +K\sigma_A^2-S_{AA^*}-2\myav{N_{A^*}}=0\\[6pt]
847: K S_{AA^*}-\sigma_{A^*}^2+\myav{N_{A^*}}=0
848: \end{array}
849: \label{eq:aasvcm}
850: \end{equation}
851: %
852: These equations are linearly dependent and cannot by themselves be
853: solved for the elements of the variance-covariance matrix.  The reason
854: is that the slow variable $N_A+N_{A^*}$ is exactly conserved by the
855: reaction scheme in Eq.~\eqref{eq:AAs}.  However this implies that
856: $\Delta N_A=-\Delta N_{A^*}$ and so $\sigma_A^2 = \sigma_{A^*}^2 =
857: -S_{AA^*}$.  These are consistent with Eqs.~\eqref{eq:aasvcm}, which
858: can now be solved to get
859: %
860: \begin{equation}
861: \sigma_A^2=\sigma_{A^*}^2=-S_{AA^*}=\frac{\myav{N_{A^*}}}{K+1}.
862: \end{equation}
863: %
864: This explains the universal value found for the sum rule deficit
865: above.
866: 
867: \begin{figure}
868: %
869: \begin{center}
870: \includegraphics{fig2.eps}
871: \end{center}
872: %
873: \caption[?]{Correlation function for the model in Eq.~\eqref{eq:AAs2},
874: with $\gamma_R = K\gamma / (K+1) = 0.12\, \mathrm{min}^{-1}$ and $K =
875: 1$.  The lines are for $k_b = 0.2$, 1.0, $5.0\,\mathrm{min}^{-1}$, and
876: $k_b\to\infty$ (dashed line).\label{fig:fig2}}
877: %
878: \end{figure}
879: 
880: To complete the discussion, we turn finally to the correlation
881: functions where the analogue of Eq.~\eqref{eq:limorder} is
882: %
883: \begin{equation}
884: \lim_{k_b\to\infty}\;\lim_{t\to0}\;C_{ii}(t) \ne
885: \lim_{t\to0}\;\lim_{k_b\to\infty}C_{ii}(t)
886: \end{equation}
887: %
888: for $i=\text{A}$ and $\text{A}^*$.  The left hand side is always equal
889: to $\sigma_i^2$.  One can show that the difference between the two
890: sides gives an alternative expression for the sum rule deficit,
891: %
892: \begin{equation}
893: \Delta\sigma_i^2 = \sigma_i^2-\lim_{t\to0} C_{ii}(t).
894: \end{equation}
895: %
896: This can be made clear by another example.  Consider
897: %
898: \begin{equation}
899: \torate{k}\text{A}\tofromrate{k_{\!f}}{k_b}\text{A}^*\torate{\gamma}.
900: \label{eq:AAs2}
901: \end{equation}
902: %
903: The steady state point statistics, which hold for arbitrary $k_b$, are
904: $\myav{N_A} = (k_b+\gamma)\myav{N_{A^*}}/k_f$,
905: $\myav{N_{A^*}} = k/\gamma$, $\sigma_A^2 / \myav{N_A} =
906: \sigma_{A^*}^2 / \myav{N_{A^*}} = 1$, $S_{AA^*}=0$.  We solve for the
907: correlation functions using the methods described in section \ref{sec:gen}.
908: Similar results are obtained for both A and $\text{A}^*$ so we report
909: results only for species A.  The Laplace-transformed autocorrelation
910: function is
911: %
912: \begin{equation}
913: \tilde C_{AA}(s) = 
914: \frac{\sigma_A^2(s+\gamma+k_b)}
915: {s^2+(\gamma+k_f+k_b)s+\gamma k_f}.
916: \end{equation}
917: %
918: Writing
919: %
920: \begin{equation}
921: s^2+(\gamma+k_f+k_b)s+\gamma k_f = (s+k_+)(s+k_-)
922: \end{equation}
923: %
924: allows us to perform the inverse Laplace transform by the method of
925: partial fractions.  The decay rates $k_\pm$ are given by
926: %
927: \begin{equation}
928: k_\pm = \frac{(\gamma+k_f+k_b)\pm
929: \sqrt{(\gamma+k_f+k_b)^2-4\gamma k_f}}{2}.
930: \end{equation}
931: %
932: The net result is 
933: %
934: \begin{equation}
935: C_{AA}(t) = \sigma_A^2 ( Ae^{-k_+t}+Be^{-k_-t})
936: \end{equation}
937: %
938: with
939: %
940: \begin{equation}
941: A=\frac{(\gamma+k_b)-k_-}{k_+-k_-},\quad B=1-A.
942: \end{equation}
943: %
944: For any finite value of $k_b$, it is clear from this that
945: $C_{AA}(0)=\sigma_A^2$ and there is no sum rule deficit.
946: 
947: Now let us take the limit $k_b\to\infty$ with $K=k_f/k_b$ being held
948: constant.  To leading order we have
949: %
950: \begin{equation}
951: \begin{array}{ll}
952: \displaystyle
953: k_+\to(K+1)\,k_b,\quad{}
954: &\displaystyle
955: k_-\to\frac{K\gamma}{K+1},\\[12pt]
956: \displaystyle
957: A\to\frac{K}{K+1},
958: &\displaystyle
959: B\to\frac{1}{K+1},
960: \end{array}
961: \end{equation}
962: %
963: and thus
964: %
965: \begin{equation}
966: \lim_{k_b\to\infty}
967: C_{AA}(t)
968: = \frac{\sigma_A^2}{K+1}\,
969: \exp\Bigl[-\frac{K\gamma t}{K+1}\Bigr].
970: \end{equation}
971: %
972: Apart from the amplitude, this correlation function is characteristic
973: of $\torate{k}\text{X}\torate{\gamma_R}$ with $\gamma_R = K\gamma /
974: (K+1)$.  This is entirely in accord with rewriting the original scheme
975: in terms of the relevant slow variable.
976: 
977: The analysis shows that in the limit $k_b\to\infty$, we have
978: $C_{AA}(0)<\sigma_A^2$.  Thus there is a non-zero sum rule deficit
979: %
980: \begin{equation}
981: \Delta\sigma_A^2=\sigma_A^2\Bigl(1-\frac{1}{K+1}\Bigr)=
982: \frac{\myav{N_{A^*}}}{K+1}
983: \end{equation}
984: %
985: where we have used $\sigma_A^2 = \myav{N_A} = \myav{N_{A^*}}/K$.
986: Again, the universal value for the sum rule deficit is recovered.
987: 
988: Our analysis shows what happens to the correlation functions in the
989: limit of fast equilibration (similar results are found for
990: $C_{A^*\!A^*}$).  At any finite $k_b$ there are two decay rates in the
991: correlation function.  In the limit $k_b\to\infty$ one of these decay
992: rates diverges whilst the other saturates to a finite value.  The
993: diverging decay rate carries with it a finite amplitude and it is this
994: that gives rise to the sum rule deficit.  The phenomenology is
995: illustrated in Fig.~\ref{fig:fig2} which shows typical results for
996: $C_{AA}(t)$ at several increasing values of the equilibration rate.
997: 
998: The results in this section are supportive of the general strategy of
999: elimination of the fast variables in terms of the appropriate slow
1000: variables.  To be specific, the strategy is to rewrite the chemical
1001: rate equations in terms of fast and slow variables, eliminate the fast
1002: variables, and re-interpret the reduced rate equations in terms of a
1003: reduced reaction network, possibly involving new chemical species (in
1004: the present example, X replaces A and A$^*$).  Our study has provided
1005: examples where one can rigorously prove that this strategy is
1006: successful.
1007: 
1008: \section{Discussion}
1009: %
1010: Firstly, let us make some general remarks about the use of noise power
1011: spectra to characterise the stochastic properties of chemical reaction
1012: networks.  As we have shown, the noise power spectra are
1013: straightforward to calculate once one has the point statistics (the
1014: mean values and the variance-covariance matrix), for a linear network,
1015: or for a linearisation of a non-linear network.  In fact, the power
1016: spectra are \emph{easier} to calculate than the autocorrelation
1017: functions, which in general require the factorisation of a polynomial,
1018: unless one is satisfied with stopping at the Laplace-transformed
1019: autocorrelation functions.  Because the power spectra are functions of
1020: frequency, they are a more refined measure than the point statistics,
1021: and can be used for instance as a sensitive test of whether two
1022: reaction networks can be mapped onto one another.
1023: 
1024: Let us comment briefly on the situation for experiments and
1025: simulations.  Both face similar difficulties in measuring the power
1026: spectra.  A signal of sufficiently long duration needs to be captured
1027: at sufficiently fine resolution in order to be able to estimate the
1028: power spectrum.  Event-driven simulations though, such as those based
1029: on the Gillespie algorithm \cite{GillKMC}, have an advantage since in
1030: principle the exact moments when a signal changes its value are known.
1031: 
1032: Secondly, we discuss the biological relevance of our results.  Our
1033: results indicate that the way noise is transmitted through a signal
1034: detection motif may be more complicated than previously thought
1035: \cite{Paulsson, Shibata, TNWtW}.  In particular, if the detection
1036: motif consumes the input signal molecules, a correlation can be set up
1037: between the intrinsic noise generated by the detection motif, and the
1038: input signal noise.  We have examined this in the context of a
1039: post-translational modification reaction attached to the output of a
1040: gene expression module.  Our analysis shows incidentally that
1041: post-translational modification can ameliorate the noise of gene
1042: expression, by smoothing out the `burstiness' of the translation step.
1043: 
1044: We also examined the effect of a fast equilibration reaction
1045: interposed in the network.  We find that, in the limit of infinitely
1046: fast equilibration, such a network can be exactly mapped onto a
1047: reduced or coarse-grained network through the use of suitably chosen
1048: slow variables.  This result supports the use of slow variables as a
1049: general strategy for model-order reduction by elimination of fast
1050: reactions.
1051: 
1052: This work was supported by the Amsterdam Centre for Computational
1053: Science (ACCS).  The work is part of the research programme of the
1054: ``Stichting voor Fundamenteel Onderzoek der Materie (FOM)'', which is
1055: financially supported by the ``Nederlandse organisatie voor
1056: Wetenschappelijk Onderzoek (NWO)''.  PBW acknowledges the hospitality
1057: of Trey Ideker and the California Institute for Telecommunications and
1058: Information Technology (Calit2) at the University of California, San
1059: Diego, where part of this work was written up.
1060: 
1061: %%%\cleardoublepage
1062: 
1063: \appendix
1064: 
1065: \section{Technical details}
1066: %
1067: Technical details and proofs are relegated to this Appendix.  For
1068: further reading we recommend Gardiner \cite{gardinerbook}, van Kampen
1069: \cite{kampenbook}, Gillespie \cite{GillCLE, GillKMC}, and Swain
1070: \cite{Swain}.
1071: %
1072: \subsection{Point statistics}
1073: \label{sec:appkb}
1074: %
1075: We first describe the chemical master equation and present a useful
1076: moment relation.  As a preliminary step we need to define some basic
1077: quantities.  The stoichiometry matrix is a non-square matrix
1078: $\nu_{i\alpha}$ equal to the change in species $i$ due to the firing
1079: of reaction $\alpha$ in the network.  We shall use roman indices to
1080: label chemical species and the greek symbol $\alpha$ to label
1081: reactions.  In terms of $\nu_{i\alpha}$, the chemical rate equations
1082: can be written
1083: %
1084: \begin{equation}
1085: \frac{d N_i}{dt}={\textstyle\sum_\alpha} \nu_{i\alpha} a_\alpha
1086: \label{eq:app:cre}
1087: \end{equation}
1088: %
1089: where $a_\alpha$ is the flux through reaction channel $\alpha$.  The quantity
1090: $a_\alpha$ is usually dependent on the current values of $N_i$ and is
1091: referred to by Gillespie as the `propensity function'.  The propensity
1092: functions for a linear reaction network will be described shortly.
1093: 
1094: In these terms, the probability
1095: $P_{N_i}$ of being in the state characterised by $N_i$ molecules of
1096: species $i$ obeys
1097: %
1098: \begin{equation}
1099: \frac{\partial P_{N_i}}{\partial t} =
1100: \sum_\alpha
1101: \Bigl\{a_\alpha(N_i-\nu_{i\alpha})\,P_{N_i-\nu_{i\alpha}}
1102: -a_\alpha(N_i)\,P_{N_i}\Bigr\}.\label{eq:app:cme}
1103: \end{equation}
1104: %
1105: If a reaction channel is impossible, for instance where
1106: $N_i-\nu_{i\alpha}<0$, we set $a_\alpha = 0$.  Eq.~\eqref{eq:app:cme}
1107: is the chemical master equation.
1108: 
1109: Multiplying Eq.~\eqref{eq:app:cme} by a general function $f(N_i)$ and
1110: summing over all $N_i$ yields
1111: %
1112: \begin{equation}
1113: \frac{d\myav{f}_t}{dt} =
1114: {\textstyle\sum_\alpha}
1115: \myav{a_\alpha(N_i)\,[f(N_i+\nu_{i\alpha})-f(N_i)]}_t.
1116: \label{eq:app:mcme}
1117: \end{equation}
1118: %
1119: This is particularly useful for deriving equations for moments.
1120: 
1121: We now indicate how the point statistics and in particular the
1122: variance-covariance matrix can be found for an arbitrary linear
1123: reaction network.  The results follow as particular cases of
1124: Eq.~\eqref{eq:app:mcme}.
1125: 
1126: First we need to specify in more detail the propensity functions for a
1127: linear reaction network.  We distinguish between zeroth- and
1128: first-order reactions in the network.  The former are pure generating
1129: reactions of the form `$\to\text{products}$'.  The latter are
1130: monomolecular reactions and comprise the rest of the network (in a
1131: linear reaction network, all reactions are either zeroth- or
1132: first-order).  The set of zeroth-order reactions is denoted by $\zero$
1133: and the set of first-order reactions by $\one$.
1134: 
1135: The propensity function for a zeroth-order reaction is
1136: %
1137: \begin{equation}
1138: a_\alpha=k_\alpha,\quad \alpha\in\zero.\label{eq:app:v0}
1139: \end{equation}
1140: %
1141: where $k_\alpha$ is the reaction rate.  For the first order reactions,
1142: we introduce an indicator matrix $\epsilon_{i\alpha}$, such that
1143: $\epsilon_{i\alpha} = 1$ if the reaction $\alpha$ involves species $i$
1144: as the reactant, and $\epsilon_{i\alpha} = 0$ otherwise.  Armed with
1145: this, the propensity function for a first-order reaction is
1146: %
1147: \begin{equation}
1148: a_\alpha=k_\alpha{\textstyle\sum_i}\epsilon_{i\alpha}N_i
1149: \quad \alpha\in\one.\label{eq:app:v1}
1150: \end{equation}
1151: %
1152: Inserting Eqs.~\eqref{eq:app:v0} and \eqref{eq:app:v1} into
1153: Eq.~\eqref{eq:app:cre} obtains Eq.~\eqref{eq:cre} in the main text,
1154: with
1155: %
1156: \begin{equation}
1157: b_i=\sumzero k_\alpha \nu_{i\alpha}\,,\quad
1158: K_{ij}=\sumone k_\alpha \nu_{i\alpha}\epsilon_{j\alpha}\,.
1159: \end{equation}
1160: %
1161: 
1162: The first moment of the chemical master equation is found by setting
1163: $f(N_i)=N_i$ in Eq.~\eqref{eq:app:mcme}.  It is easy to show that this
1164: gives Eq.~\eqref{eq:nit} in the main text, with the above definitions
1165: of $b_i$ and $K_{ij}$.
1166: 
1167: The second moment of the chemical master equation is found by setting
1168: $f(N_i)=N_iN_j$ in Eq.~\eqref{eq:app:mcme}.  This gives a closed set
1169: of equations for the elements of the variance-covariance matrix.  In
1170: steady state, these equations can be reduced to
1171: %
1172: \begin{equation}
1173: {\textstyle\sum_k}(K_{ik} S_{kj}+K_{jk} S_{ki})+H_{ij}=0
1174: \label{eq:app:vcm}
1175: \end{equation}
1176: %
1177: where
1178: %
1179: \begin{equation}
1180: \textstyle
1181: H_{ij}=\sumzero k_\alpha \nu_{i\alpha} \nu_{j\alpha}+\sum_k\Bigl(
1182: \sumone k_\alpha \nu_{i\alpha} \nu_{j\alpha} \epsilon_{k\alpha}
1183: \Bigr)\myav{N_k}
1184: \label{eq:app:h}
1185: \end{equation}
1186: %
1187: Eq.~\eqref{eq:app:vcm} is a set of linear simultaneous equations for
1188: the elements $S_{ij}$ of the variance-covariance matrix.
1189: 
1190: As an example of the application of this machinery, consider the model
1191: for gene expression in Eq.~\eqref{eq:MA} in the main text.  There are
1192: two species, M and A, and four reactions.  The stoichiometry matrix is
1193: %
1194: \begin{equation}
1195: \nu_{i\alpha}=\Bigl(
1196: \begin{array}{c}1\\0\end{array}\Big|
1197: \begin{array}{ccc}-1&0&0\\0&1&-1\end{array}\Bigr).
1198: \end{equation}
1199: %
1200: The reaction rates are
1201: %
1202: \begin{equation}
1203: k_\alpha=(\;k\;|\;\lambda\;\;\rho\;\;\gamma\;).
1204: \end{equation}
1205: %
1206: The indicator matrix is
1207: %
1208: \begin{equation}
1209: \epsilon_{i\alpha}=\Bigl(
1210: \begin{array}{c}\cdot\\\cdot\end{array}\Big|
1211: \begin{array}{ccc}1&1&0\\0&0&1\end{array}\Bigr).
1212: \end{equation}
1213: %
1214: In these, the reactions to the left of the vertical line correspond to 
1215: $\alpha\in\zero$ and the reactions to the right of the line correspond to
1216: $\alpha\in\one$ (the indicator matrix is only defined for the latter
1217: reactions). 
1218: 
1219: From these one computes
1220: %
1221: \begin{equation}
1222: K_{ij}=\Bigl(\begin{array}{cc}
1223: -\lambda & 0\\ \rho & -\gamma \end{array}\Bigr).
1224: \end{equation}
1225: %
1226: Obviously this could have been written down by inspection. Less
1227: trivially, one also finds
1228: %
1229: \begin{equation}
1230: \begin{array}{ll}
1231: \displaystyle
1232: H_{ij}&=
1233: \Bigl(\begin{array}{cc}k+\lambda\myav{N_M} & 0\\
1234: 0 & \rho\myav{N_M}+\gamma\myav{N_A}\end{array}\Bigr)\\[18pt]
1235: \displaystyle
1236: &=\Bigl(\begin{array}{cc}2\lambda\myav{N_M} & 0\\
1237: 0 & 2\gamma\myav{N_A}\end{array}\Bigr).
1238: \end{array}
1239: \end{equation}
1240: %
1241: Eqs.~\eqref{eq:app:vcm} simplify to
1242: %
1243: \begin{equation}
1244: \begin{array}{l}
1245: -2\lambda\sigma_M^2+2\lambda\myav{N_M}=0,\\[6pt]
1246: \rho\sigma_M^2-\gamma S_{M\!A}-\lambda S_{M\!A} = 0,\\[6pt]
1247: 2\rho S_{M\!A}-2\gamma\sigma_A^2+2\gamma\myav{N_A}=0.
1248: \end{array}
1249: \end{equation}
1250: %
1251: There are only three equations since Eq.~\eqref{eq:app:vcm} is
1252: symmetric in $i$ and $j$.  These are solved to get
1253: Eqs.~\eqref{eq:MAvcm} in the main text.
1254: 
1255: \subsection{Regression theorem}
1256: \label{sec:appc}
1257: %
1258: The proof of the regression theorem Eq.~\eqref{eq:corr} in the main
1259: text is straightforward and parallels the development in Gardiner
1260: \cite{gardinerbook}.  We start by writing out an explicit expression for
1261: the correlation function
1262: %
1263: \begin{equation}
1264: \begin{array}{l}
1265: \displaystyle
1266: C_{ij}(t)=\int d\{N_{i,0}\}\,d\{N_{i,t}\}\;
1267: \Delta N_{i,0} \;\Delta N_{j,t}\\[6pt]
1268: {}\hspace{8em}{}\times P_s(N_{i,0})\;P(N_{i,t}|N_{i,0}; t).
1269: \end{array}
1270: \label{eq:app:ceq}
1271: \end{equation}
1272: %
1273: In this, $P_s(N_{i,0})$ is the steady state probability distribution
1274: for the starting point $N_i = N_{i,0}$, and $P(N_{i,t}|N_{i,0}; t)$ is
1275: the conditional probability distribution for $N_i = N_{i,t}$ at time
1276: $t$, given that the system started with $N_i = N_{i,0}$ at $t = 0$.
1277: These probability distributions could in principle be found by solving
1278: the chemical master equation.  We now define two kinds of averages
1279: %
1280: \begin{equation}
1281: \begin{array}{l}
1282: \myav{f(N_i)}_{t|0}=
1283: \int d\{N_{i,t}\}\;f(N_{i,t})\;P(N_{i,t}|N_{i,0}; t),\\[6pt]
1284: \myav{f(N_i)}_0=
1285: \int d\{N_{i,0}\}\;f(N_{i,0})\;P_s(N_{i,0}).
1286: \end{array}
1287: \end{equation}
1288: %
1289: In words, the first is the average value of a function of $N_i =
1290: N_{i,t}$ at time $t$, given that the system started with $N_i =
1291: N_{i,0}$ at $t = 0$.  The second is the average value of a function of
1292: $N_i = N_{i,0}$ in steady state conditions.  In terms of these,
1293: Eq.~\eqref{eq:app:ceq} can be written as
1294: %
1295: \begin{equation}
1296: C_{ij}(t)=\myav{{}\,\Delta N_i\,\myav{\Delta N_j}_{t|0}\,}_0.
1297: \label{eq:app:cgen}
1298: \end{equation}
1299: %
1300: This result is completely general \cite{gardinerbook}.  
1301: 
1302: For a linear reaction network, taking the first moment of the chemical
1303: master equation proves an analogue to Eq.~\eqref{eq:dnit}, namely
1304: %
1305: \begin{equation}
1306: \frac{d\myav{\Delta N_i}_{t|0}}{dt} =
1307: {\textstyle\sum_j}K_{ij}\myav{\Delta N_j}_{t|0}.
1308: \end{equation}
1309: %
1310: Combining this with Eq.~\eqref{eq:app:cgen} demonstrates that
1311: $C_{ij}(t)$ obeys Eq.~\eqref{eq:corr} in the main text.
1312: 
1313: \subsection{Chemical Langevin equations}
1314: \label{sec:appl}
1315: %
1316: We prove that the power spectra obtained from the chemical Langevin
1317: equations for a linear reaction network are equivalent to the exact
1318: results obtained from the chemical master equation.  As the first
1319: step, Gillespie has shown that the chemical Langevin equations for a
1320: general network are \cite{GillCLE}
1321: %
1322: \begin{equation}
1323: \frac{d N_i}{dt}={\textstyle\sum_\alpha} \nu_{i\alpha} a_\alpha
1324: +{\textstyle\sum_\alpha} \nu_{i\alpha} a_\alpha^{1/2}\,\Gamma_\alpha.
1325: \label{eq:app:cle}
1326: \end{equation}
1327: %
1328: In this, $\Gamma_\alpha$ are independent unit Gaussian white noise
1329: functions, one for each reaction channel.  Applying this to a linear
1330: reaction network obtains
1331: %
1332: \begin{equation}
1333: \frac{d N_i}{dt}={\textstyle\sum_j} K_{ij} N_j + b_i + \eta_i
1334: \label{eq:app:cle2}
1335: \end{equation}
1336: %
1337: with
1338: %
1339: \begin{equation}
1340: \textstyle
1341: \eta_i=\sumzero k_\alpha^{1/2} \nu_{i\alpha} \Gamma_\alpha
1342: +\sumone 
1343: (k_\alpha \sum_j\epsilon_{j\alpha}N_j)^{1/2}
1344: \nu_{i\alpha} \Gamma_\alpha.
1345: \end{equation}
1346: %
1347: It follows that the $\eta_i$ are Gaussian white noise functions with
1348: the following statistics
1349: %
1350: \begin{equation}
1351: \myav{\eta_i}=0,\quad
1352: \myav{\eta_i(t)\,\eta_j(t')}=H_{ij}\,\delta(t-t').
1353: \end{equation}
1354: %
1355: The correlation matrix $H_{ij}$ is \emph{identical} to the matrix that
1356: features in the computation of the variance-covariance matrix, given
1357: by Eq.~\eqref{eq:app:h} above, which can be interpreted in this
1358: context as a fluctuation-dissipation theorem.
1359: 
1360: Working in terms of the deviation from steady state, and taking the
1361: Fourier transform of the chemical Langevin equations, obtains
1362: %
1363: \begin{equation}
1364: i\omega\,\Delta  N_i = 
1365: {\textstyle\sum_j} K_{ij} \Delta  N_j +  \eta_i.
1366: \label{eq:app:cle3}
1367: \end{equation}
1368: %
1369: We conclude
1370: %
1371: \begin{equation}
1372: P_i(\omega) = \myav{|\Delta N_i|^2} = 
1373: {\textstyle\sum_{jk}}B_{ij} H_{jk} B_{ki}^\dag
1374: \label{eq:app:picle}
1375: \end{equation}
1376: %
1377: where $B_{ij}(i\omega)$ is the inverse of $i\omega\delta_{ij}-K_{ij}$
1378: and $B_{ij}^\dag(i\omega) = B_{ji}(-i\omega)$ is the adjoint.  Our
1379: task is to prove that Eq.~\eqref{eq:app:picle} is equivalent to
1380: Eqs.~\eqref{eq:lcorr} and \eqref{eq:l2p} in the main text.
1381: 
1382: The problem is made easier if we rewrite everything in abstract matrix
1383: notation.  The result we have just obtained can be written as
1384: %
1385: \begin{equation}
1386: \mymat{P}_1=\mymat{B}\cdot \mymat{H}\cdot\mymat{B}^\dag
1387: \label{eq:app:bhb}
1388: \end{equation}
1389: %
1390: with
1391: %
1392: \begin{equation}
1393: \mymat{B}=(i\omega\mymat{I}-\mymat{K})^{-1},\quad
1394: \mymat{B}^\dag=(-i\omega\mymat{I}-\mymat{K}^\mathrm{T})^{-1}
1395: \label{eq:app:b}
1396: \end{equation}
1397: %
1398: ($\mymat{K}^\mathrm{T}$ is the transpose of $\mymat{K}$).  The
1399: diagonal elements of $\mymat{P}_1$ are the power spectra from the
1400: chemical Langevin equation route.
1401: 
1402: We now similarly rewrite the results from the analysis of the chemical
1403: master equation.  Eqs.~\eqref{eq:app:vcm}, \eqref{eq:lcorr} and
1404: \eqref{eq:l2p} can be written respectively as
1405: %
1406: \begin{equation}
1407: \begin{array}{l}
1408: \mymat{K}\cdot\mymat{S}+\mymat{S}\cdot\mymat{K}^\mathrm{T}+\mymat{H}=0,\\[6pt]
1409: i\omega \mymat{C}-\mymat{S}=\mymat{K}\cdot\mymat{C},
1410: \quad\text{or}\quad\mymat{C} = \mymat{B}\cdot\mymat{S},\\[6pt]
1411: \mymat{P}_2=\mymat{C}+\mymat{C}^\dag
1412: =\mymat{B}\cdot\mymat{S} + \mymat{S}\cdot\mymat{B}^\dag
1413: \end{array}
1414: \label{eq:app:p2}
1415: \end{equation}
1416: %
1417: (note that $\mymat{S}$ is real and symmetric).  The diagonal elements
1418: of $\mymat{P}_2$ are the power spectra from the chemical master
1419: equation route.
1420: 
1421: Eliminate $\mymat{H}$ between the first of Eqs.~\eqref{eq:app:p2} and
1422: Eq.~\eqref{eq:app:bhb} to get
1423: %
1424: \begin{equation}
1425: \mymat{P}_1=-(\mymat{B}\cdot\mymat{S}
1426: \cdot\mymat{K}^\mathrm{T}\!\cdot\mymat{B}^\dag
1427: +\mymat{B}\cdot\mymat{K}\cdot\mymat{S}\cdot\mymat{B}^\dag).
1428: \label{eq:app:p1a}
1429: \end{equation}
1430: %
1431: It follows from Eq.~\eqref{eq:app:b} that
1432: %
1433: \begin{equation}
1434: \mymat{B}\cdot\mymat{K}=i\omega\mymat{B}-\mymat{I},\quad
1435: \mymat{K}^\mathrm{T}\!\cdot\mymat{B}^\dag
1436: =-i\omega\mymat{B}^\dag-\mymat{I}.
1437: \end{equation}
1438: %
1439: On substituting these into Eq.~\eqref{eq:app:p1a}, there is a
1440: cancellation of terms, and one finds
1441: %
1442: \begin{equation}
1443: \mymat{P}_1=\mymat{B}\cdot\mymat{S}+\mymat{S}\cdot\mymat{B}^\dag.
1444: \end{equation}
1445: %
1446: Comparison with the last of Eqs.~\eqref{eq:app:p2} shows that
1447: $\mymat{P}_1=\mymat{P}_2$.  This contains our desired result, and
1448: proves that the power spectra obtained by the chemical Langevin
1449: equation route are the same as those obtained by the (exact) chemical
1450: master equation route.  Actually, we have a slightly stronger result,
1451: since it follows that $\myav{|\Delta N_X|^2} =
1452: \mymat{x}^\mathrm{T}\!\cdot \mymat{P}_1\cdot\mymat{x} =
1453: \mymat{x}^\mathrm{T}\!\cdot \mymat{P}_2\cdot\mymat{x}$, for any linear
1454: combination $N_X=\sum_i x_i N_I$.
1455: 
1456: Swain has presented an analysis of the chemical Langevin equations for,
1457: effectively, a general linear reaction network \cite{Swain}.  He
1458: obtains general results for the correlation functions and the
1459: variance-covariance matrix in terms of the eigenvalues of $K_{ij}$.
1460: The present section can be read as proving that the
1461: variance-covariance matrix can be found by a simpler route, by solving
1462: Eq.~\eqref{eq:app:vcm} in terms of the noise correlation matrix.  For
1463: the correlation functions though, our route involves an inverse
1464: Laplace transform which in general requires that a polynomial be
1465: factorised, and is equivalent to solving the secular equation for
1466: $K_{ij}$.  So our methods do not provide any additional simplification
1467: compared to Swain for the correlation functions.  As we have
1468: emphasised in the main text though, our results demonstrate that the
1469: power spectra are much easier to calculate than the correlation
1470: functions.
1471: 
1472: %\bibliographystyle{biophysj}
1473: %\bibliography{noise}
1474: \begin{thebibliography}{28}
1475: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1476: \expandafter\ifx\csname bibnamefont\endcsname\relax
1477:   \def\bibnamefont#1{#1}\fi
1478: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1479:   \def\bibfnamefont#1{#1}\fi
1480: \expandafter\ifx\csname citenamefont\endcsname\relax
1481:   \def\citenamefont#1{#1}\fi
1482: \expandafter\ifx\csname url\endcsname\relax
1483:   \def\url#1{\texttt{#1}}\fi
1484: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1485: \providecommand{\bibinfo}[2]{#2}
1486: \providecommand{\eprint}[2][]{\url{#2}}
1487: 
1488: \bibitem[{\citenamefont{Rao et~al.}(2002)\citenamefont{Rao, Wolf, and
1489:   Arkin}}]{RWA}
1490: \bibinfo{author}{\bibfnamefont{C.~V.} \bibnamefont{Rao}},
1491:   \bibinfo{author}{\bibfnamefont{D.~M.} \bibnamefont{Wolf}}, \bibnamefont{and}
1492:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Arkin}},
1493:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{420}},
1494:   \bibinfo{pages}{231} (\bibinfo{year}{2002}).
1495: 
1496: \bibitem[{\citenamefont{Arkin et~al.}(1998)\citenamefont{Arkin, Ross, and
1497:   McAdams}}]{ARMcA}
1498: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Arkin}},
1499:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ross}}, \bibnamefont{and}
1500:   \bibinfo{author}{\bibfnamefont{H.~H.} \bibnamefont{McAdams}},
1501:   \bibinfo{journal}{Genetics} \textbf{\bibinfo{volume}{149}},
1502:   \bibinfo{pages}{1633} (\bibinfo{year}{1998}).
1503: 
1504: \bibitem[{\citenamefont{Elowitz and Leibler}(2000)}]{Elowitz}
1505: \bibinfo{author}{\bibfnamefont{M.~B.} \bibnamefont{Elowitz}} \bibnamefont{and}
1506:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Leibler}},
1507:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{403}},
1508:   \bibinfo{pages}{335} (\bibinfo{year}{2000}).
1509: 
1510: \bibitem[{\citenamefont{Paulsson et~al.}(2000)\citenamefont{Paulsson, Berg, and
1511:   Ehrenberg}}]{PBE}
1512: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Paulsson}},
1513:   \bibinfo{author}{\bibfnamefont{O.~G.} \bibnamefont{Berg}}, \bibnamefont{and}
1514:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ehrenberg}},
1515:   \bibinfo{journal}{Proc. Natl. Acad. Sci.} \textbf{\bibinfo{volume}{97}},
1516:   \bibinfo{pages}{7148} (\bibinfo{year}{2000}).
1517: 
1518: \bibitem[{\citenamefont{{Santill\'an} and Mackey}(2004{\natexlab{a}})}]{SanMAc}
1519: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Santill\'an}}}
1520:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~C.}
1521:   \bibnamefont{Mackey}}, \bibinfo{journal}{Biophys. J.}
1522:   \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{1282}
1523:   (\bibinfo{year}{2004}{\natexlab{a}}).
1524: 
1525: \bibitem[{\citenamefont{Berg}(1978)}]{Berg}
1526: \bibinfo{author}{\bibfnamefont{O.~G.} \bibnamefont{Berg}}, \bibinfo{journal}{J.
1527:   theor. Biol.} \textbf{\bibinfo{volume}{71}}, \bibinfo{pages}{587}
1528:   (\bibinfo{year}{1978}).
1529: 
1530: \bibitem[{\citenamefont{McAdams and Arkin}(1997)}]{McAA}
1531: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{McAdams}} \bibnamefont{and}
1532:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Arkin}},
1533:   \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{94}},
1534:   \bibinfo{pages}{814} (\bibinfo{year}{1997}).
1535: 
1536: \bibitem[{\citenamefont{Thattai and van Oudenaarden}(2001)}]{TvO}
1537: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Thattai}} \bibnamefont{and}
1538:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{van Oudenaarden}},
1539:   \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{98}},
1540:   \bibinfo{pages}{8614} (\bibinfo{year}{2001}).
1541: 
1542: \bibitem[{\citenamefont{Gardiner}(2004)}]{gardinerbook}
1543: \bibinfo{author}{\bibfnamefont{C.~W.} \bibnamefont{Gardiner}},
1544:   \emph{\bibinfo{title}{Handbook of Stochastic Methods}}
1545:   (\bibinfo{publisher}{Springer}, \bibinfo{address}{Berlin},
1546:   \bibinfo{year}{2004}).
1547: 
1548: \bibitem[{\citenamefont{Gillespie}(2000)}]{GillCLE}
1549: \bibinfo{author}{\bibfnamefont{D.~T.} \bibnamefont{Gillespie}},
1550:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{113}},
1551:   \bibinfo{pages}{297} (\bibinfo{year}{2000}).
1552: 
1553: \bibitem[{\citenamefont{{Santill\'an} and
1554:   Mackey}(2004{\natexlab{b}})}]{SanMac2}
1555: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Santill\'an}}}
1556:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~C.}
1557:   \bibnamefont{Mackey}}, \bibinfo{journal}{Biophys. J.}
1558:   \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{75}
1559:   (\bibinfo{year}{2004}{\natexlab{b}}).
1560: 
1561: \bibitem[{\citenamefont{Samoilov et~al.}(2002)\citenamefont{Samoilov, Arkin,
1562:   and Ross}}]{SAR}
1563: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Samoilov}},
1564:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Arkin}}, \bibnamefont{and}
1565:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ross}}, \bibinfo{journal}{J.
1566:   Phys. Chem. A} \textbf{\bibinfo{volume}{106}}, \bibinfo{pages}{10205}
1567:   (\bibinfo{year}{2002}).
1568: 
1569: \bibitem[{\citenamefont{Paulsson}(2004)}]{Paulsson}
1570: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Paulsson}},
1571:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{427}},
1572:   \bibinfo{pages}{415} (\bibinfo{year}{2004}).
1573: 
1574: \bibitem[{\citenamefont{Shibata and Fujimoto}(2005)}]{Shibata}
1575: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Shibata}} \bibnamefont{and}
1576:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Fujimoto}},
1577:   \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{102}},
1578:   \bibinfo{pages}{331} (\bibinfo{year}{2005}).
1579: 
1580: \bibitem[{\citenamefont{{T\u{a}nase-Nicola}
1581:   et~al.}(2005)\citenamefont{{T\u{a}nase-Nicola}, Warren, and ten
1582:   Wolde}}]{TNWtW}
1583: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{{T\u{a}nase-Nicola}}},
1584:   \bibinfo{author}{\bibfnamefont{P.~B.} \bibnamefont{Warren}},
1585:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~R.} \bibnamefont{ten
1586:   Wolde}} (\bibinfo{year}{2005}), \bibinfo{note}{see q-bio/0508027}.
1587: 
1588: \bibitem[{\citenamefont{Kepler and Elston}(2001)}]{Kepler01}
1589: \bibinfo{author}{\bibfnamefont{T.~B.} \bibnamefont{Kepler}} \bibnamefont{and}
1590:   \bibinfo{author}{\bibfnamefont{T.~C.} \bibnamefont{Elston}},
1591:   \bibinfo{journal}{Biophys. J.} \textbf{\bibinfo{volume}{81}},
1592:   \bibinfo{pages}{3116} (\bibinfo{year}{2001}).
1593: 
1594: \bibitem[{\citenamefont{Haseltine and Rawlings}(2002)}]{Haseltine02}
1595: \bibinfo{author}{\bibfnamefont{E.~L.} \bibnamefont{Haseltine}}
1596:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~B.}
1597:   \bibnamefont{Rawlings}}, \bibinfo{journal}{J. Chem. Phys.}
1598:   \textbf{\bibinfo{volume}{117}}, \bibinfo{pages}{6959} (\bibinfo{year}{2002}).
1599: 
1600: \bibitem[{\citenamefont{Shibata}(2003{\natexlab{a}})}]{Shibata03a}
1601: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Shibata}}, \bibinfo{journal}{J.
1602:   Chem. Phys.} \textbf{\bibinfo{volume}{119}}, \bibinfo{pages}{6629}
1603:   (\bibinfo{year}{2003}{\natexlab{a}}).
1604: 
1605: \bibitem[{\citenamefont{Shibata}(2003{\natexlab{b}})}]{Shibata03}
1606: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Shibata}},
1607:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{67}},
1608:   \bibinfo{pages}{061906} (\bibinfo{year}{2003}{\natexlab{b}}).
1609: 
1610: \bibitem[{\citenamefont{Bundschuh et~al.}(2003)\citenamefont{Bundschuh, Hayot,
1611:   and Jayaprakash}}]{Bundschuh}
1612: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bundschuh}},
1613:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Hayot}}, \bibnamefont{and}
1614:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Jayaprakash}},
1615:   \bibinfo{journal}{Biophys. J.} \textbf{\bibinfo{volume}{84}},
1616:   \bibinfo{pages}{1606} (\bibinfo{year}{2003}).
1617: 
1618: \bibitem[{\citenamefont{Rao and Arkin}(2003)}]{Rao03}
1619: \bibinfo{author}{\bibfnamefont{C.~V.} \bibnamefont{Rao}} \bibnamefont{and}
1620:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Arkin}},
1621:   \bibinfo{journal}{J. Chem. Phys} \textbf{\bibinfo{volume}{118}},
1622:   \bibinfo{pages}{4999} (\bibinfo{year}{2003}).
1623: 
1624: \bibitem[{\citenamefont{Roussel and Zhu}(2004)}]{Roussel04}
1625: \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Roussel}} \bibnamefont{and}
1626:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Zhu}}, \bibinfo{journal}{J.
1627:   Chem. Phys.} \textbf{\bibinfo{volume}{121}}, \bibinfo{pages}{8716}
1628:   (\bibinfo{year}{2004}).
1629: 
1630: \bibitem[{\citenamefont{Chatterjee et~al.}(2005)\citenamefont{Chatterjee,
1631:   Vlachos, and Katsoulakis}}]{Chatterjee05}
1632: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Chatterjee}},
1633:   \bibinfo{author}{\bibfnamefont{D.~G.} \bibnamefont{Vlachos}},
1634:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~A.}
1635:   \bibnamefont{Katsoulakis}}, \bibinfo{journal}{J. Chem. Phys}
1636:   \textbf{\bibinfo{volume}{122}}, \bibinfo{pages}{024112}
1637:   (\bibinfo{year}{2005}).
1638: 
1639: \bibitem[{\citenamefont{Cao et~al.}(2005)\citenamefont{Cao, Gillespie, and
1640:   Petzold}}]{Cao05}
1641: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Cao}},
1642:   \bibinfo{author}{\bibfnamefont{D.~T.} \bibnamefont{Gillespie}},
1643:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~R.}
1644:   \bibnamefont{Petzold}}, \bibinfo{journal}{J. Chem. Phys.}
1645:   \textbf{\bibinfo{volume}{122}}, \bibinfo{pages}{014116}
1646:   (\bibinfo{year}{2005}).
1647: 
1648: \bibitem[{\citenamefont{Straube et~al.}(2005)\citenamefont{Straube, Flockerzi,
1649:   {M\"uller}, and Hauser}}]{Straube05}
1650: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Straube}},
1651:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Flockerzi}},
1652:   \bibinfo{author}{\bibfnamefont{S.~C.} \bibnamefont{{M\"uller}}},
1653:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~J.~B.}
1654:   \bibnamefont{Hauser}}, \bibinfo{journal}{J. Phys. Chem. A}
1655:   \textbf{\bibinfo{volume}{109}}, \bibinfo{pages}{441} (\bibinfo{year}{2005}).
1656: 
1657: \bibitem[{\citenamefont{Gillespie}(1977)}]{GillKMC}
1658: \bibinfo{author}{\bibfnamefont{D.~T.} \bibnamefont{Gillespie}},
1659:   \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{81}},
1660:   \bibinfo{pages}{2340} (\bibinfo{year}{1977}).
1661: 
1662: \bibitem[{\citenamefont{van Kampen}(2001)}]{kampenbook}
1663: \bibinfo{author}{\bibfnamefont{N.~G.} \bibnamefont{van Kampen}},
1664:   \emph{\bibinfo{title}{Stochastic Processes in Physics and Chemistry.}}
1665:   (\bibinfo{publisher}{Elsevier}, \bibinfo{address}{Amsterdam},
1666:   \bibinfo{year}{2001}).
1667: 
1668: \bibitem[{\citenamefont{Swain}(2004)}]{Swain}
1669: \bibinfo{author}{\bibfnamefont{P.~S.} \bibnamefont{Swain}},
1670:   \bibinfo{journal}{J. Mol. Biol.} \textbf{\bibinfo{volume}{344}},
1671:   \bibinfo{pages}{965} (\bibinfo{year}{2004}).
1672: 
1673: \end{thebibliography}
1674: 
1675: \end{document}
1676: 
1677: %%% END DOCUMENT %%%
1678: