This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Slack Reactants: A State-Space Truncation Framework to Estimate Quantitative Behavior of the Chemical Master Equation

Jinsu Kim  Jason Dark  German Eciso   and   Suzanne Sindi Department of Mathematics, University of California, Irvine, [email protected] of Mathematics, University of California, IrvineDepartment of Mathematics, University of California, Irvine, Department of Developmental and Cell Biology, University of California, IrvineDepartment of Applied Mathematics, University of California, Merced
Abstract

State space truncation methods are widely used to approximate solutions of the chemical master equation. While most methods of this kind focus on truncating the state space directly, in this work we propose modifying the underlying chemical reaction network by introducing slack reactants that indirectly truncate the state space. More specifically, slack reactants introduce an expanded chemical reaction network and impose a truncation scheme based on user defined properties, such as mass-conservation. This network structure allows us to prove inheritance of special properties of the original model, such as irreducibility and complex balancing.

We use the network structure imposed by slack reactants to prove the convergence of the stationary distribution and first arrival times. We then provide examples comparing our method with the stationary finite state projection and finite buffer methods. Our slack reactant system appears to be more robust than some competing methods with respect to calculating first arrival times.

1 Introduction

Chemical reaction networks (CRN) are a fundamental tool in the modeling of biological systems, providing a concise representation of known chemical or biological dynamics. A CRN is defined by a family of chemical reactions of the form

i=1nαiXii=1nβiXi,\sum_{i=1}^{n}\alpha_{i}X_{i}\rightarrow\sum_{i=1}^{n}\beta_{i}X_{i},

where αi\alpha_{i} and βi\beta_{i} are the number of species XiX_{i} consumed and produced in this reaction, respectively. The classical approach to modeling this CRN is to consider the concentrations c(t)=(c1(t),c2(t),,cn(t))c(t)=(c_{1}(t),c_{2}(t),\dotsc,c_{n}(t))^{\mkern-2.0mu\top}, where ci(t)c_{i}(t) is the concentration of species XiX_{i} at time tt, and to use a system of nonlinear differential equations to describe the evolution of the concentrations.

Suppose we are interested in studying the typical enzyme-substrate system [30], given by the CRN

S+EDP+E.S+E\rightleftharpoons D\rightarrow P+E. (1)

Given an initial condition where the molecular numbers of both DD and PP are 0, a natural question to ask is how long the system takes to produce the first copy of PP. Similarly, there exists a time in the future where SS and DD are fully depleted, resulting in a chemically-inactive system, and one can ask when this occurs. These are both quantities that the classical deterministic modeling leaves unanswered – by considering only continuous concentrations, there is not a well-defined way to address modeling questions at the level of single molecules as the model assumes that all the reactions simultaneously occur within infinitesimally small time intervals.

Instead, by explicitly considering each individual copy of a molecule, we may formulate a continuous-time Markov chain. This stochastic modeling is especially important when the system consists of low copies of species, in which case the influence of intrinsic noise is magnified [34, 11, 2]. Rather than deterministic concentrations c(t)c(t), we consider the continuous-time Markov chain X(t)=(X1(t),X2(t),,Xn(t))X(t)=(X_{1}(t),X_{2}(t),\dotsc,X_{n}(t))^{\mkern-2.0mu\top} describing the molecular number of each species XiX_{i} at time tt.

The questions regarding the enzyme-substrate system (1), such as the time of the first production of PP, simply correspond to the first passage times of various combinations of states. For a continuous-time Markov process XX, the first passage time to visit a set KK of system states is formally defined as τ=inf{t0:X(t)K}\tau=\inf\{t\geq 0:X(t)\in K\}. One can directly compute E(τ)E(\tau) by using the transition rate matrix AA. With few exceptions (such as CRNs with only zero- or first-order reactions [28]), most chemical reaction networks of any notable complexity will have an intractably large or infinite state-space, i.e. they exhibit the curse of dimensionality. This direct approach can, therefore, suffer from the high dimensionality of the transition rate matrix.

An alternative approach is to estimate the mean first passage time by generating sample trajectories with stochastic simulation algorithms such as the Gillespie algorithm [15]. This overcomes the curse of dimensionality since a single trajectory needs only to keep track of its current population numbers. Nevertheless, there still remain circumstances under which it is more efficient to numerically evaluate the exact solution of the chemical master equation – in particular, when the Markov process rarely visits KK so that significantly long simulations may be required to sample enough trajectories to estimate the mean first passage time [29].

Fortunately, a broad class of state space reduction methods have recently been developed that allow for direct treatment of the transition rate matrix. These methods are based on the truncation-and-augmentation of the state space [35, 36, 32, 16], the finite buffer method [7, 8] and linear programming [25, 26]. A recent review summarizes truncation-based methods [27]. The stationary finite state projection (sFSP)[16], and the finite buffer method [7, 8] are examples of such truncation-based methods, which we will describe in detail below. They all satisfy provable error estimates on the probability distributions when compared with the original distribution. On the other hand, each of these methods has potential limitations for estimating mean first passage times and other quantitative features. For instance, using sFSP depends on the choice of a designated state, which can significantly alter the estimate for first passage times.

In this paper we provide a new algorithm of state space reduction, the slack reactant method, for stochastically modeled CRNs. In this algorithm we generate a new CRN from an existing CRN by adding one or multiple new species, so that the associated stochastic system satisfies mass conservation relations and is confined to finitely many states. In order to ensure equivalent dynamics to the original system, we define a mild form of non-mass action kinetics for the new system. Since the state space reduction is implemented using a fully determined CRN, we can study the CRN using well-known tools of CRN theory such as deficiency zero , Lyapunov functions , etc, as long as they extend to this form of kinetics.

In Section 3.2 we provide an algorithm to produce a slack variable network given a desired set of mass conservation relations. In addition to its theoretical uses, this algorithm allows to implement existing software packages such as CERENA [23], StochDynTools [19] and FEEDME [24] for chemical reaction networks to generate quantities such as the moment dynamics of the associated stochastic processes, using the network structures as inputs.

We employ classical truncation Markov chain approaches to prove convergence theorems for slack networks. For fixed time tt, if a probability density of each slack system under conservation quantity NN converges to its stationary distribution uniformly in NN, then the stationary distribution of the slack system converges to the original stationary distribution as NN tends to infinity. We further prove that under a uniform tail condition of first passage times, the mean first passage time of the original system can be approximated with slack systems confined on a sufficiently large truncated state space. Finally we show that the existence of Lyapunov function for an original system guarantees that all the sufficient conditions for the first passage time convergence are satisfied. We also show that this truncation method is natural in the sense that a slack system admits the same stationary distribution up to a constant multiplication as the stationary distribution of the original system if the original system is complex balanced.

This paper is outlined as follows. In Section 3, we introduce the slack reactant method and include several illustrative examples. In Section 4, we demonstrate that the slack method compares favorably with other state space truncation methods (sFSP, finite buffer method) when calculating mean first passage times. We prove convergence in the mean first passage time, and other properties, in Section 5 In Section 7, we use slack reactants to estimate the mean first passage times for practical CRN models such as a Lotka-Volterra population model and a system of protein synthesis with a slow toggle switch.

2 Stochastic chemical reaction networks

A chemical reaction network (CRN) is a graph that describes the evolution of a biochemical system governed by a number of species (𝒮)(\mathcal{S}) and reactions ()(\mathcal{R}). Each node in the graph represents a possible state of the system and nodes are connected by directed edges when a single reaction transforms one state into another. Each reaction consists of complexes and a reaction rate constant. For example, the reaction representing the transformation of complex ν\nu to complex ν\nu^{\prime} at rate κ\kappa is written as follows:

ν𝜅ν.\nu\xrightarrow{\kappa}\nu^{\prime}. (2)

A complex, such as ν\nu, is defined as a number of each species Si𝒮S_{i}\in\mathcal{S}. That is, ν=(ν1,ν2,,νd)\nu=(\nu_{1},\nu_{2},\dots,\nu_{d}) representing a complex i=1dνiSi\sum_{i=1}^{d}\nu_{i}S_{i}, where νi0\nu_{i}\geq 0 are the stochiometric coefficients indicating how many copies of each species Si𝒮S_{i}\in\mathcal{S} belong in complex ν\nu. The full CRN is thus defined by (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) where 𝒞\mathcal{C} and Λ\Lambda represent the set of complexes and reaction propensities respectively.

When intrinsic noise plays a significant role for system dynamics, we use a continuous-time Markov chain 𝐗(t)=(X1(t),X(2),,Xd(t))\mathbf{X}(t)=(X_{1}(t),X(2),\dots,X_{d}(t)) to model the copy numbers of species SiS_{i} of a reaction network. The stochastic system treats individual reactions as discrete transitions between integer-valued states of the system. The probability density for X(t)X(t) is denoted as

p𝐱0(𝐱,t)=P(X(t)=𝐱|X(0)=𝐱0),p_{\mathbf{x}_{0}}(\mathbf{x},t)=P(X(t)=\mathbf{x}\ |\ X(0)=\mathbf{x}_{0}),

where X(0)X(0) is the initial probability density. We occasionally denote by p(𝐱)p(\mathbf{x}) the probability omitting the initial state 𝐱0\mathbf{x}_{0} when the contexts allows. For each state 𝐱\mathbf{x}, the probability density p(𝐱)p(\mathbf{x}) obeys the chemical master equation (CME), which gives the time-evolution of 𝐩(t)\mathbf{p}(t) with a linear system of ordinary differential equations [14]:

ddt𝐩(t)=𝐩A.\frac{d}{dt}\mathbf{p}^{{\mkern-2.0mu\top}}(t)=\mathbf{p}^{{\mkern-2.0mu\top}}A. (3)

Here, the entry AijA_{ij} (iji\neq j) is the transition rate at which the ii-th state transitions to jj-th state. Letting 𝐱\mathbf{x} and 𝐱\mathbf{x}^{\prime} be the ii-th and jj-th states, the transition rate from 𝐱\mathbf{x} to 𝐱\mathbf{x}^{\prime} is

Aij=νν𝐱+νν=𝐱λνν(𝐱),\displaystyle A_{ij}=\sum_{\begin{subarray}{c}\nu\to\nu^{\prime}\\ \mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu}=\mathbf{x}^{\prime}\end{subarray}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x}),

where λνν\lambda_{\nu\to\nu^{\prime}} is the reaction intensity for a reaction νν\nu\to\nu^{\prime}. The diagonal elements of AA are defined as

Ajj=ijAij.A_{jj}=-\sum_{i\neq j}A_{ij}.

Normally, reaction intensities hold a standard property,

λνν(𝐱)>0if and only ifxiνifor each i.\displaystyle\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})>0\quad\text{if and only if}\quad x_{i}\geq\nu_{i}\quad\text{for each $i$}. (4)

A typical choice for λ\lambda_{\to^{\prime}} is mass-action, which defines for 𝐱=(x1,x2,,xd)\mathbf{x}=(x_{1},x_{2},\dots,x_{d}) and ν=(α1,α2,,αd)\mathbf{\nu}=(\alpha_{1},\alpha_{2},\dots,\alpha_{d}),

λνν=κννk=1dxi(αi)\displaystyle\lambda_{\nu\to\nu^{\prime}}=\kappa_{\nu\to\nu^{\prime}}\prod_{k=1}^{d}x_{i}^{(\alpha_{i})}

where κνν\kappa_{\nu\to\nu^{\prime}} is the rate constant. Here we used the notation m(n)=m(m1)(m2)(mn+1)𝟙{nm},m^{(n)}=m(m-1)(m-2)\cdots(m-n+1)\mathbbm{1}_{\{n\geq m\}}, for positive integers mm and nn.

3 Construct of Slack Networks

In this Section, we introduce the method of slack reactants, which adds new species to an existing CRN so that the state space of the associated stochastic process is truncated to a finite subset of states. This model reduction accurately approximates the original system as the size of truncation is large. We begin with a simple example to demonstrate the main idea of the method.

3.1 Slack reactants for a simple birth-death model

Consider a mass-action 1-dimensional birth-death model,

κ2κ1X.\emptyset\xrightleftharpoons[\kappa_{2}]{\kappa_{1}}X. (5)

For the associated stochastic process XX, the mass-action assumption defines reactions intensities as λX(x)=κ1\lambda_{\emptyset\to X}(x)=\kappa_{1} and λX(x)=κ2x\lambda_{X\to\emptyset}(x)=\kappa_{2}x for each reaction in (5). Note that the count of species XX could be any positive integer value as the birth reaction X\emptyset\to X can occur unlimited times. Therefore the state space for this system consists of infinitely many states. Consider instead the CRN

Yκ2κ1XY\xrightleftharpoons[\kappa_{2}]{\kappa_{1}}X (6)

where we have introduced the slack reactant YY. This new network admits X+Y{X+Y} as a conservative law since for each reaction either one species is degraded by one while the other is produced by one.

Since the purpose of this new network is to approximate the qualitative behavior of the original system (5), we minimize the contribution of the slack reactant YY for modeling the associated stochastic system. Hence, we assign YY a special reaction intensity – instead of λYX(x,y)=κ1y\lambda_{Y\to X}(x,y)=\kappa_{1}y using mass-action, we choose

λYX(x,y)=κ1𝟙{y0},\displaystyle\lambda_{Y\to X}(x,y)=\kappa_{1}\mathbbm{1}_{\{y\geq 0\}}, (7)

and we use the same intensity λXY(x,y)=κ2x\lambda_{X\to Y}(x,y)=\kappa_{2}x for the reaction XYX\to Y. By forcing YY to have “zero-order kinetics,” we ensure that the computed rates remain the same throughout the chemical network except for on the imposed boundary. This choice of reaction intensities not only preserves the conservative law X(t)+Y(t)=X(0)+Y(0)X(t)+Y(t)=X(0)+Y(0) but also prevent the slack reactant YY from having negative counts with the characteristic term 𝟙{y0}\mathbbm{1}_{\{y\geq 0\}}.

Refer to caption
Figure 1: Schematic images of various state-space truncation methods for a CRN with two species: (a) Slack network (this paper), (c) Finite State Projection (FSP) [32], (c) Stationary Finite State Projection (sFSP) [16], (d) Finite Buffer Method [7, 8].

3.2 Algorithm for constructing a network with slack reactants

In general, by using slack reactants, any conservative laws can be deduced in a similar fashion to encode the desired truncation directly in the CRN. We have found this perspective to be advantageous with respect to studying complex CRNs. Rather than thinking about the software implementation of the truncation, it is often easier to design the truncation in terms of slack reactants, then implement the already-truncated system exactly.

We now provide an algorithm to automatically determine the slack reactants required for a specified truncation. It is often the case that a “natural” formulation arises (typically by replacing zero-order complexes with slack reactants) but when that is not the case, one can still find a useful approximation by following our algorithm.

Consider any CRN (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda). Suppose we wish to apply a number of conservation bounds to reduce the state-space of the associated chemical master equation, e.g. many equations of the form

wi,1X1+wi,2X2++wi,dXdNi,w_{i,1}X_{1}+w_{i,2}X_{2}+\dotsc+w_{i,d}X_{d}\leq N_{i}, (8)

for i=1,2,,mi=1,2,\dots,m. Collecting the conservation vectors {𝐰i}i=1m\{\mathbf{w}_{i}\}_{i=1}^{m} into the rows of the m×dm\times d matrix WW and each integer {Ni}\{N_{i}\} into the m×1m\times 1 vector 𝐍\mathbf{N}, we may concisely write

W𝐗𝐍,W\mathbf{X}\preceq\mathbf{N}, (9)

where \preceq denotes an element-wise less-than-or-equal operator and 𝐗\mathbf{X} is the column vector of reactants {Xi}{\{X_{i}\}}.

We augment the d×1d\times 1 vector 𝐗\mathbf{X} with a m×1m\times 1 vector 𝐘\mathbf{Y} of slack reactants to convert the inequality to an equality:

(WI)(𝐗𝐘)=𝐍,\begin{pmatrix}W&I\end{pmatrix}\begin{pmatrix}\mathbf{X}\\ \mathbf{Y}\end{pmatrix}=\mathbf{N}, (10)

where II is the m×mm\times m identity matrix. This matrix multiplication implies mm conservative laws among XiX_{i}’s and slack reactants YiY_{i}’s. In other words, we inject the variables 𝐘=(Y1,Y2,,Ym)\mathbf{Y}=(Y_{1},Y_{2},\dots,Y_{m}) into the existing reactions in such a way as to enforce these augmented conservation laws.

To do so, we need to properly define several CRN quantities. Let SS be the |𝒞|×|||\mathcal{C}|\times|\mathcal{R}| connectivity matrix: if the rr-th reaction establishes an edge from the ii-th complex to the jj-th, then Sir=Sjr=1-S_{ir}=S_{jr}=1 and the other entries in column rr are 0. Let CC be the |𝒮|×|𝒞||\mathcal{S}|\times|\mathcal{C}| complex matrix CC. Each column of CC corresponds to a unique complex and each row corresponds to a reactant. The entry CijC_{ij} is the multiplier of XiX_{i} in complex jj (which may be implicitly 0). The stoichiometry matrix is simply Γ=CS{\Gamma=CS}, whose ii-th column indicates the reaction vector of the ii-th reaction. Note that the CRN admits a conservation law r1X1(t)+r2X2(t)++rdXd(t)=r1X1(0)+r2X2(0)++rdXd(0)r_{1}X_{1}(t)+r_{2}X_{2}(t)+\cdots+r_{d}X_{d}(t)=r_{1}X_{1}(0)+r_{2}X_{2}(0)+\cdots+r_{d}X_{d}(0) if and only if 𝐫Γ=𝟎\mathbf{r}^{\mkern-2.0mu\top}\Gamma=\mathbf{0}, where r=(r1,r2,,rd)r=(r_{1},r_{2},\dots,r_{d})^{\mkern-2.0mu\top}.

Now consider a modified network where we inject Y1,Y2,,YKY_{1},Y_{2},\dots,Y_{K} into the complexes. Let DijD_{ij} denote the (unknown) multiplier of YiY_{i} in the jj-th complex. We require the following to be true:

(WI)(CD)S=0.\begin{pmatrix}W&I\end{pmatrix}\begin{pmatrix}C\\ D\end{pmatrix}S=0. (11)

Then by using the same connectivity matrix SS and a new complex matrix C~=(CD)\widetilde{C}=\begin{pmatrix}C\\ D\end{pmatrix}, we obtain a new CRN that enforces the conservative laws (10). To find DD satisfying (11), we choose DD such that

D=UWCD=U-WC (12)

for an m×|𝒞|m\times|\mathcal{C}| matrix UU with each row of the form ui𝟏=ui(1,1,,1)u_{i}\mathbf{1}^{\top}=u_{i}(1,1,\dots,1)^{\mkern-2.0mu\top} for some positive integer uiu_{i}. Since 𝟏TS=𝟎\mathbf{1}^{T}S=\mathbf{0}, this guarantees (11).

A new network can be generated with the complex matrix C~\widetilde{C} and the connectivity matrix SS. We call this network a regular slack network with species 𝒮~\widetilde{\mathcal{S}} and complexes 𝒞~\widetilde{\mathcal{C}}. For the jj-th reaction νjνj\nu_{j}\to\nu^{\prime}_{j}\in\mathcal{R} corresponding to the jj-th column of the matrix CSCS, we denote by ν~jν~j~\tilde{\nu}_{j}\to\tilde{\nu}^{\prime}_{j}\in\widetilde{\mathcal{R}} the reaction corresponding to the jj-th column of the matrix C~S\widetilde{C}S. That is, ν~jν~j\tilde{\nu}_{j}\to\tilde{\nu}^{\prime}_{j} is the reaction obtained from νjνj\nu_{j}\to\nu^{\prime}_{j} by adding slack reactants.

We model the slack network by X𝐍(t)=(X1𝐍(t),X2𝐍(t),,Xd𝐍(t))X^{\mathbf{N}}(t)=(X^{\mathbf{N}}_{1}(t),X^{\mathbf{N}}_{2}(t),\dots,X^{\mathbf{N}}_{d}(t)) where each of the entries represents the count of species in the new network. Note that the count of each slack reactant YiY_{i} is fully determined by species counts Xi𝐍X^{\mathbf{N}}_{i}’s because of the conservation law(s). As such, we do not explicitly model the YiY_{i}’s.

The intensity function λν~ν~𝐍\lambda^{\mathbf{N}}_{\tilde{\nu}\to\tilde{\nu}} of X𝐍X^{\mathbf{N}} for a reaction ν~ν~\tilde{\nu}\to\tilde{\nu}^{\prime} is defined as

λν~ν~𝐍(𝐱)=λνν(𝐱)i=1m𝟙{yiν~d+1},\displaystyle\lambda^{\mathbf{N}}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})=\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})\prod_{i=1}^{m}\mathbbm{1}_{\{y_{i}\geq\tilde{\nu}_{d+1}\}}, (13)

where yi=Ni(wi1x1+wi1x1++widxd)y_{i}=N_{i}-(w_{i1}x_{1}+w_{i1}x_{1}+\cdots+w_{id}x_{d}), and νν\nu\to\nu^{\prime} is the reaction in \mathcal{R}. Then we denote by (𝒮~,𝒞~,~,Λ𝐍)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{\mathbf{N}}) a new system with slack reactants obtained from the algorithm, where 𝒦N\mathcal{K}_{N} is the collection of kinetic rates {λν~ν~N:ν~ν~~}\{\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}:\tilde{\nu}\to\tilde{\nu}^{\prime}\in\widetilde{\mathcal{R}}\}. We refer this system with slack reactants to a slack system.

Here we highlight that the connectivity matrix SS and the complex matrix CC of the original network are preserved for a new network with slack reactants. Thus the original network and the new network obtained from the algorithm have the same connectivity. This identical connectivity allows the qualitative behavior of the original network, which solely depends on SS and CC, to be inherited to the new network with slack reactants. We particularly exploit the inheritance of accessibility and the inheritance of Poissonian stationary distribution in Section 6.

Remark 3.1.

A single conservation relation, such as 𝐰X+y=N\mathbf{w}\cdot X+y=N with a non-negative vector 𝐰\mathbf{w}, is sufficient to truncate the given state space into a finite state space. Hence, in this manuscript, we mainly consider a slack network that is obtained with a single conservation vector 𝐰\mathbf{w} and a single slack reactant YY.

Remark 3.2.

Although we primarily think about bounding our species counts from above, we could also bound species counts from below by choosing negative integers for wi,jw_{i,j}.

Example 3.1.

We illustrate our slack algorithm with an example CRN consisting of two species and five reactions indicated by edge labels on the following network:

\emptysetAABB\small{1}⃝\small{2}⃝ \small{3}⃝\small{4}⃝\small{5}⃝ (14)

We enumerate the complexes in the order of ,A\emptyset,A and BB. We order the reactions according to their labels on the network. Thus the connectivity matrix SS and complex matrix CC are defined as follows:

S=[110011111000111],S=\begin{bmatrix}-1&1&0&0&-1\\ 1&-1&-1&1&0\\ 0&0&1&-1&1\\ \end{bmatrix},
C=[010001].C=\begin{bmatrix}0&1&0\\ 0&0&1\end{bmatrix}.

Suppose we set a conservation bound A+BNA+B\leq N for some N>0N>0. Then the matrix A=[11]A=\begin{bmatrix}1&1\end{bmatrix} and by (12), we have D=[uuu][011]D=\begin{bmatrix}u&u&u\end{bmatrix}-\begin{bmatrix}0&1&1\end{bmatrix}.

When u=1u=1, the network with slack reactant YY is

YAB,\leavevmode\hbox to61.5pt{\vbox to62.53pt{\pgfpicture\makeatletter\hbox{\hskip 8.10278pt\lower-8.09985pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{{}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{ }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.01389pt}{42.10793pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$Y$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{ }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{41.77458pt}{42.10793pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$A$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{ }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.0434pt}{-3.41666pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$B$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \par\par\par\par{{}}{}{{}}{}{{}}{}{{}}{}{{}}{}\hbox{\hbox{\hbox{\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{8.24097pt}{46.53662pt}\pgfsys@curveto{19.57262pt}{47.91693pt}{26.18425pt}{47.91052pt}{37.05673pt}{46.56493pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.99243}{-0.12282}{0.12282}{0.99243}{37.05672pt}{46.56493pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.87808pt}{51.09892pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{37.51323pt}{44.54106pt}\pgfsys@curveto{26.18425pt}{43.13896pt}{19.57262pt}{43.13255pt}{8.69757pt}{44.45724pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-0.99266}{0.12093}{-0.12093}{-0.99266}{8.69759pt}{44.45723pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.87808pt}{39.95055pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{40.55536pt}{39.16425pt}\pgfsys@curveto{28.95477pt}{24.29826pt}{21.3928pt}{16.72781pt}{6.90268pt}{5.3933pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-0.78764}{-0.61612}{0.61612}{-0.78764}{6.90268pt}{5.39331pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ }}{ } {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{28.3003pt}{17.38603pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{5.1099pt}{6.54037pt}\pgfsys@curveto{16.72781pt}{21.3928pt}{24.29826pt}{28.95477pt}{38.8016pt}{40.27238pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.78838}{0.6152}{-0.6152}{0.78838}{38.8016pt}{40.27237pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ }}{ } {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{17.38603pt}{28.3003pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}} {{{{{}}{}{}{}{}{{}}}}}{}{{{{{}}{}{}{}{}{{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}{}{}{}{}{{}}\pgfsys@moveto{0.0pt}{37.22194pt}\pgfsys@lineto{0.0pt}{8.75986pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{0.0pt}{8.75986pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.533pt}{22.76091pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}, (15)

where we have the conservation relation A+B+Y=NA+B+Y=N, where N=A(0)+B(0)+Y(0)N=A(0)+B(0)+Y(0).

When u=2u=2, the network with slack reactant YY is

2YA+YB+Y,\leavevmode\hbox to102.32pt{\vbox to79.56pt{\pgfpicture\makeatletter\hbox{\hskip 17.17456pt\lower-17.17456pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{{}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{ }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-6.5139pt}{42.10793pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$2Y$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{ }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{54.41191pt}{42.52458pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$A+Y$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{ }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.16837pt}{-3.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$B+Y$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \par\par\par\par{{}}{}{{}}{}{{}}{}{{}}{}{{}}{}\hbox{\hbox{\hbox{\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{10.21373pt}{46.77884pt}\pgfsys@curveto{26.15625pt}{49.06192pt}{35.3298pt}{49.24586pt}{50.8935pt}{47.65112pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.9948}{-0.10193}{0.10193}{0.9948}{50.8935pt}{47.65114pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{30.75284pt}{52.19629pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{51.35109pt}{43.44524pt}\pgfsys@curveto{35.3298pt}{41.80363pt}{26.15625pt}{41.98755pt}{10.66907pt}{44.20544pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-0.9899}{0.14175}{-0.14175}{-0.9899}{10.66908pt}{44.20544pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{30.75284pt}{38.85316pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{55.34895pt}{34.4pt}\pgfsys@curveto{41.14984pt}{22.17142pt}{32.25903pt}{16.23415pt}{15.93411pt}{8.01085pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-0.8931}{-0.44987}{0.44987}{-0.8931}{15.93413pt}{8.01086pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ }}{ } {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{39.92035pt}{16.14455pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{{}}{{{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{{{}}{{}}{{}}{{}}{{}}}{{{{}}{}{}{}{}{{}}}} }{{}{}}{{}} {}{}{}{{{}}{{}}{{}}} {{{}}{{}}{{}}} {}{{}}{}{{}}{}{{}}{}{}{}{}{}{}{}{{}}{}{{}}{}{}{}{}{}{{}}{}{}{}{}{{}}\pgfsys@moveto{13.17438pt}{11.32785pt}\pgfsys@curveto{27.40805pt}{23.547pt}{36.27519pt}{29.44803pt}{52.63101pt}{37.65445pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.8938}{0.44846}{-0.44846}{0.8938}{52.63101pt}{37.65443pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ }}{ } {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{28.62524pt}{29.5547pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}} {{{{{}}{}{}{}{}{{}}}}}{}{{{{{}}{}{}{}{}{{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}{}{}{}{}{{}}\pgfsys@moveto{0.0pt}{35.23438pt}\pgfsys@lineto{0.0pt}{17.8346pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{0.0pt}{17.8346pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.533pt}{26.3045pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}, (16)

where we have the same conservation relation A+B+Y=NA+B+Y=N.

Here we explain why network (15) is preferred to (16) because it is less intrusive. Let X=(XA,XB,XY)X=(X_{A},X_{B},X_{Y}) and X=(XA,XB,XY)X^{\prime}=(X^{\prime}_{A},X^{\prime}_{B},X^{\prime}_{Y}) be the stochastic process associated with networks (15) and (16) respectively. Suppose the initial state of both XX and XX^{\prime} is (0,0,N)(0,0,N). XX can reach the state (0,N,0)(0,N,0) by transitioning NN times with the reaction YBY\to B. This state corresponds to the state (0,N)(0,N) in the original network (14). On the other hand, XX^{\prime} cannot reach the state (0,N,0)(0,N,0). This is because the states (0,0,N)(0,0,N) and (0,N1,1)(0,N-1,1) are the only states from which XX^{\prime} jumps to (0,0,N)(0,0,N). However, no reaction in (16) can be fired at the states since no species YY presents at those states.

Consequently, one state, that is accessible in the original network (14), is lost in the system associated with network (16). However, it can be shown that the stochastic process associated with network (15) preserves all the states of the original network. This occurs mainly because the matrix DD for network (15) is sparser than the matrix DD for network (16). We discuss how to minimize the effect of slack reactants in Section 3.3.

3.3 Optimal Slack CRNs for effective approximation of the original systems

The algorithm we introduced to construct a network with slack reactants provides is valid and unique up to any user-defined conservation bounds (8), and the outcome is the matrix DD, shown in (12), that indicates the stoichiometric coefficient of slack reactants at each complex.

As we showed in Example 3.1, to minimize the ‘intrusiveness’ of a slack network, we can simplify a slack network by setting as many Dij=0{D_{ij}=0} as possible. To do that, we choose the entries of 𝐮\mathbf{u} such that uiu_{i} is the maximum entry of the ii-th row of ACAC. We further optimize the effect of the slack reactants by removing the ‘redundant’ stoichiometric coefficient of slack reactants. For example, for a CRN,

A2A,\displaystyle\emptyset\rightleftharpoons A\to 2A, (17)

the algorithm generates the following new CRN with a single slack reactant YY

2YA+Y2A.\displaystyle 2Y\rightleftharpoons A+Y\to 2A. (18)

However, by breaking up the connectivity, we can also generate another network

YA,A+Y2A.\displaystyle Y\rightleftharpoons A,\quad A+Y\to 2A. (19)

The network in (18) is more intrusive than the network in (19) in the sense of accessibility. At any state where Y=0Y=0, the system associated with (18) will remain unchanged because no reaction can take place. However, the reaction AYA\to Y in (19) can always occur despite Y=0Y=0. Hence (19) preserves the accessibility of the original system associated with (17) as any state for AA is accessible from any other state in the original reaction system (17). We refer such a system with slack reactants generated by canceling redundant slack reactants to an optimized slack system. In Section 6, we explore the accessibility of an optimized reaction network with slack reactants in a more general setting.

Finally, we can make a network with slack reactants admit a better approximation of a given CRN by choosing an optimized conservation relation in (8). First, we assume that only a single conservation law and a single slack reactant are added to a given CRN. For the purpose of state space truncation onto finitely many states, a single conservation law is enough as all species could be bounded by NN as shown in (8). Let this single conservation law be

w1X1+w2X2+wdXd+Y=N.w_{1}X_{1}+w_{2}X_{2}+\cdots w_{d}X_{d}+Y=N.

Then the matrix WW in (9) is a vector (w1,w2,,wd)(w_{1},w_{2},\dots,w_{d})^{\mkern-2.0mu\top}, and we denote this by 𝐰\mathbf{w}. By the definition of the intensities (13) for a network with slack reactants, some reactions are turned off when Y=0Y=0, i.e. w1X1++wdXd=Nw_{1}X_{1}+\dots+w_{d}X_{d}=N. Geometrically, a reaction outgoing from the hyperplane w1X1++wdXd=Nw_{1}X_{1}+\dots+w_{d}X_{d}=N is turned off (Figure 2).

Hence we optimize the estimation with slack reactants by minimizing such intrusiveness of turned off reactions. To do that, we choose 𝐯\mathbf{v} which minimizes the number of the reactions in {νν:(νν)𝐰>0}\{\nu\to\nu^{\prime}\in\mathcal{R}:(\nu^{\prime}-\nu)\cdot\mathbf{w}>0\}.

X1X_{1}X2X_{2}O(N,0)(N,0)(0,N)(0,N)(νν)𝐰0(\nu^{\prime}-\nu)\cdot\mathbf{w}\leq 0(νν)𝐰>0(\nu^{\prime}-\nu)\cdot\mathbf{w}>0w1X1+w2X2=N\hskip 85.35826ptw_{1}X_{1}+w_{2}X_{2}=N
Figure 2: The dotted arrows correspond to the reaction vectors that are turned off (i.e. the associated reaction intensity is zero) at the bound of the state space for a slack system as described in Section 3.3.

4 Comparison to Other Truncation-Based Methods

In this Section we demonstrate that our method can potentially resolve limitations in calculating mean first passage times observed in other methods of state space truncation, namely sFSP and the finite-buffer method. Both methods require the user to make decisions about the state-space truncation that may introduce variability in the results. While all methods will converge to the true result as the size of the state space increases, we show our method is less dependent on user defined quantities. This minimizes additional decision-making on the part of the user that can lead to suboptimal results, especially in a context where the solution of the original system is not known.

4.1 Comparison to the sFSP method

A well known state truncation algorithm is known as the Finite State Projection (FSP) method [32]. For a given continuous-time Markov chain, the associated FSP model is restricted to a finite state space. If the process escapes this truncated space, the state is sent to a designated absorbing state (see Figure 1B). For a fixed time tt, the probability density function of the original system can be approximated by using the associated FSP model with sufficiently many states. The long-term dynamics of the original system, however, is not well approximated because the probability flow of the FSP model leaks to the designated absorbing state in the long run.

To fix this limitation of FSP, Gupta et al. proposed the stationary Finite State Projection method (sFSP)[16]. This method also projects the original state space onto a finite state space as the FSP method intended to. But sFSP does not create a designated absorbing state, as all outgoing transitions from the projected finite state space are merged to a single state xx^{*} ‘inside’ the finite state space (Figure 1C). The sFSP has been frequently used to estimate the long-term distribution of discrete stochastic models. However, if the size of the truncated state space is not sufficiently large, this method could fail to provide accurate estimation for the first passage time. To demonstrate this case, we consider the following simple 2-dimensional model. In the network shown in Figure 3A, two X1X_{1} proteins are dimerized into protein X2X_{2} while X1X_{1} is being produced at relatively high rate. The state space of the original model is the full 2-dimensional positive integer grid. We estimate the time until the system reaches one of the two states indicated in red in Figure 3C, and we use alternative methods to do this.

Refer to caption
Figure 3: Comparison between an sFSP model and a slack network for the same reaction system. A. Original 2-dimensional system and its slack network. B. Mean first passage time as a function of the truncated state size, comparing sFSP and the slack reactant method. C. The state space truncation corresponding to the sFSP method and the slack method for this model. Target states for the first passage time are indicated in red. D. Mean of 100 sample trajectories obtained by Gillespie simulations for the original, the slack system and the sFSP system. The red lines indicate the mean of the trajectories that have touched the target states within [0,500], while the black lines are the mean of the trajectories that did not touch the target space during this time. E. Probability denstiy heat maps of the stochastic processes modeled under the original, the slack system and the sFSP model, respectively, at t=500t=500.

For the sFSP, we project the original state space onto the rectangle by restricting X1NX_{1}\leq N and X2NX_{2}\leq N for some N>0N>0, and we fix the origin (0,0)(0,0) as the designated state xx^{*} (Figure 3C). If the process associated with the sFSP model escapes the rectangle, it transports to the designated state immediately. On the other hand, we also consider a slack network shown in Figure 3B, where we introduce the conservation law X1+2X2+Y=NX_{1}+2X_{2}+Y=N for some N>0N>0. Let τ=inf{t0:X1=1 and X2{1,2}}\tau=\inf\{t\geq 0:X_{1}=1\text{ and }X_{2}\in\{1,2\}\} be the first passage time we want to estimate.

By using the inverse matrix of the truncated transition matrix for each method, as detailed in [9], we obtain the mean of τ\tau by using different values of NN. We also obtain an ‘almost’ true mean of τ\tau by using 10510^{5} Gillespie simulations of the original process. As shown in Figure 3B, the slack network model provides a more accurate mean first passage time estimation for the size of truncation in between 100 and 400 if the designated return state of sFSP is (0,0)(0,0).

The inaccurate estimate from the sFSP is due to the choice of a return state. The sFSP model escapes the confined space often because the production rate of X1X_{1} is relatively high. When it returns back to the origin, it is more likely to visit the target states {X1=1 and X2{1,2}}\{X_{1}=1\text{ and }X_{2}\in\{1,2\}\} than the original process.

Figure 3 shows that the mean first passage time of this system using sFSP depends significantly on the location of the chosen designated state. One of the two states is a particularly poor choice for sFSP, but it illustrates the idea that without previous knowledge of the system it can be difficult to know which states will perform well.

We display the behavior of individual solutions of the original model, the slack network, and the sFSP model in Figure 3D. The trajectory plots show that within the time interval [0,500][0,500], almost half of the 100 samples from both the original model and the slack network model stay far away from the target states, while all the 100 sample trajectories from the sFSP model stay close to the target states. We also illustrate this point in Figure 3E with heat maps of the three models at t=500t=500. Note that only in the case of the sFSP, the probability densities are concentrated at the target states.

4.2 Comparison to the finite buffer method

The finite buffer method was proposed to estimate the stationary probability landscape with state space truncation and a novel state space enumeration algorithm [7, 8]. For a given stochastic model associated with a CRN, the finite buffer method sets inequalities among species such as (8), so-called buffer capacities. Then at each state 𝐱\mathbf{x}, the transition rate of a reaction νν\nu\to\nu^{\prime} is set to be zero if at least one of the inequality does not hold at 𝐱+νν\mathbf{x}+\nu^{\prime}-\nu. We note the algorithm, described in Section 3.2, for generating a slack network uses the same inequalities. Thus, the finite buffer method and the slack reactant method truncate the state space in the same way. We have shown, in Section 3.3 this type of truncation can create ‘additional’ absorbing states. These additional absorbing states change the accessibility between the states which means the mean first passage times cannot be accurately estimated. However, the regular slack systems preserve the network structure of the original network. Hence we are able to prove, as we already noted, that regular slack networks inherit the accessibility of the original network as long as the original network is ‘weakly reversible’ as we will define below.

We demonstrate this disparity between the the finite buffer and slack methods with the following network. Consider the mass-action system (14) with a fixed initial state 𝐱0=(a0,b0)\mathbf{x}_{0}=(a_{0},b_{0}). We are interested in estimating the mean first passage time to a target state 𝐱T=(10,10)\mathbf{x}_{T}=(10,10). Note that the state space is irreducible (i.e., every state is accessible from any other state) as the network consists of unions of cycles. This condition, the union of cycles, is precisely what is meant by weakly reversible.[13, 5] Thus, the original stochastic system has no absorbing state and is accessible to the target state 𝐱T\mathbf{x}_{T}.

To use the finite buffer method on this network, we set 2XA+XBN2X_{A}+X_{B}\leq N as the buffer capacity, where X=(XA,XB)X=(X_{A},X_{B}) is the associated stochastic process. (Here we choose N>30N>30 so the state space contains the target state 𝐱T\mathbf{x}_{T}.) Hence when XX satisfies 2XA+XB=N2X_{A}+X_{B}=N, the reactions A\emptyset\to A, BAB\to A and B\emptyset\to B cannot be fired as 2XA+XB2X_{A}+X_{B} exceeds the buffer capacity. We now demonstrate the system has a new absorbing state. By first using reaction ABA\to B, to deplete all AA, and then B\emptyset\to B, every state can reach the state (0,N)(0,N) in finite time with positive probability. The state (0,N)(0,N) is absorbing state because no other reactions can occur. Reactions AA\to\emptyset and ABA\to B require at least one AA species, and any other reactions lead to states exceeding the buffer capacity. Therefore, the finite buffer method has introduced a new absorbing state not present in the original model so that it is not accessible to 𝐱T\mathbf{x}_{T} with positive probability.

Now, we show the explicit network structure of our slack network formulation will preserve the accessibility of the original system. We consider the same inequality 2XA+XBN2X_{A}+X_{B}\leq N as above with N>30N>30. We generate the slack network by using the algorithm shown in Section 3.2,

2Y2YAAB+YB+Y (20)

Note that the associated stochastic process X=(XA,XB,Y)X=(X_{A},X_{B},Y) admits the conservation relation 2XA+XB+Y=N2X_{A}+X_{B}+Y=N implying that 2XA+XBN2X_{A}+X_{B}\leq N. The state (0,N,0)(0,N,0) can not be reached as the only state that is accessible to (0,N,0)(0,N,0) is (0,N1,2)(0,N-1,2), but it violates the conservation law.

As we highlighted in Section 3.3, slack networks preserve the connectivity of the original network (14), hence the network (20) is also weakly reversible. Thus the state space of the stochastic process associated with the slack network is irreducible by Corollary 6.1. This implies that there is no absorbing state and the system is accessible to 𝐱T\mathbf{x}_{T}, unlike the stochastic process associated with the finite buffer relation.

5 Convergence Theorems for slack networks

In this section we establish theoretical results on the convergence of properties of a slack network to the original network. (Proofs of the theorems below are provided in Appendix B.) Many of these results rely on theorems from Hart and Tweedie[18] who studied when the probability density function of a truncated Markov process converges to that of the original Markov process. We employ the same idea of their proof to show the convergence of a slack network to the original network.

By assuming “uniforming mixing,” we show the convergence of the stationary distribution of the slack system to the stationary distribution of the original system as the conservation quantity NN grows. Furthermore, we show the convergence of mean first passage times for the slack network to the true mean first passage times. In particular, all these conditions hold when there is a Lyapunov function for the original system.

In this Section, assume that a given CRN (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) is well defined for all time tt and let (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) be an associated slack network obtained with a single conservative quantity 𝐰XN\mathbf{w}\cdot X\leq N. We denote by XX and XNX_{N} the associated stochastic processes the original CRN and the slack network respectively. We fix the initial state for both systems, i.e., X(0)=XN(0)=𝐱0X(0)=X_{N}(0)=\mathbf{x}_{0} for some 𝐱0\mathbf{x}_{0} and for each NN. (This means we can only consider slack systems where NN is large enough so that 𝐰𝐱(0)<N\mathbf{w}\cdot\mathbf{x}(0)<N.) Assume that both the original and slack systems are irreducible, and denote by 𝕊\mathbb{S} and 𝕊N\mathbb{S}_{N} the state spaces for each respectively. (In Section 6, we prove accessibility properties that the slack system can inherit from the original system.)

Notice that every state in 𝕊N\mathbb{S}_{N} satisfies our conservation inequality. That is, for every 𝐱𝕊N\mathbf{x}\in\mathbb{S}_{N} we have 𝐰𝐱N\mathbf{w}\cdot\mathbf{x}\leq N. It is possible that 𝕊N=𝕊N+1\mathbb{S}_{N}=\mathbb{S}_{N+1} for some NN. For simplicity, we assume that the truncated state space is always enlarged with respect to the conservative quantity NN, that is 𝕊N𝕊N+1\mathbb{S}_{N}\subset\mathbb{S}_{N+1} for each NN. (For the general case, we could simply consider a subsequence NkN_{k} such that Nk<Nk+1N_{k}<N_{k+1} and 𝕊Nk𝕊Nk+1\mathbb{S}_{N_{k}}\subset\mathbb{S}_{N_{k+1}} for each kk.)

As defined in Section 2, λννΛ\lambda_{\nu\to\nu^{\prime}}\in\Lambda is the intensity of a reaction νν\nu\to\nu^{\prime} for the associated stochastic system XX. We also denote by λν~ν~NΛN\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}\in\Lambda^{N} the intensity of a reaction ν~ν~\tilde{\nu}\to\tilde{\nu}^{\prime} for the associated stochastic system XNX_{N}. Finally we let p(𝐱,t)p(\mathbf{x},t) and pN(𝐱,t)p_{N}(\mathbf{x},t) be the probability density function of XX and XNX_{N}, respectively. We begin with the convergence of the probability density functions of the slack network to the original network with increasing NN.

Theorem 5.1.

For any 𝐱𝕊N\mathbf{x}\in\mathbb{S}_{N} and t0t\geq 0, we have

limN|p(𝐱,t)pN(𝐱,t)|=0.\displaystyle\lim_{N\to\infty}|p(\mathbf{x},t)-p_{N}(\mathbf{x},t)|=0.

A Markov process defined on a finite state space admits a stationary distribution. Hence XNX_{N} admits a stationary distribution πN\pi_{N}. If the slack system satisfies the condition of “uniform mixing”, that is the convergence rate of pN(𝐱,t)πN(𝐱)1\|p_{N}(\mathbf{x},t)-\pi_{N}(\mathbf{x})\|_{1} is uniform in NN, then we have the following result:

Theorem 5.2.

Suppose XX admits a stationary distribution π\pi. Suppose further that there exists a positive function h(t)h(t), which is independent of NN, such that pN(,t)πN1h(t)\|p_{N}(\cdot,t)-\pi_{N}\|_{1}\leq h(t) and limth(t)=0\displaystyle\lim\limits_{t\to\infty}h(t)=0. Then

limNππN1=0.\displaystyle\lim_{N\to\infty}\|\pi-\pi_{N}\|_{1}=0.

We now consider convergence of the mean first passage time of XNX_{N}. Recall, we assumed that both stochastic processes have the same initial state X(0)=XN(0)=𝐱0X(0)=X_{N}(0)=\mathbf{x}_{0} and both state spaces 𝕊\mathbb{S} and 𝕊N\mathbb{S}_{N} are irreducible. Hence, for any K𝕊K\subseteq\mathbb{S}, each state in 𝕊N\mathbb{S}_{N} is accessible to KK for sufficiently large NN.

Theorem 5.3.

For a subset KK of the state space of XX, let τ\tau and τN\tau_{N} be the first passage times to KK for XX and to K𝕊NK\cap\mathbb{S}_{N} for XNX_{N}, respectively. Assume the following conditions,

  1. 1.

    XX admits a stationary distribution π\pi, and

  2. 2.

    limNππN1=0\displaystyle\lim\limits_{N\to\infty}\|\pi-\pi_{N}\|_{1}=0.

Then for any t0t\geq 0,

limN|P(τ>t)P(τN>t)|=0.\displaystyle\lim_{N\to\infty}|P(\tau>t)-P(\tau_{N}>t)|=0.

If we further assume that

  1. 3.

    E(τ)<E(\tau)<\infty, and

  2. 4.

    there exists g(t)g(t) such that P(τN>t)g(t)P(\tau_{N}>t)\leq g(t) for all NN and 0g(t)𝑑t<\int_{0}^{\infty}g(t)dt<\infty,

then

limN|E(τ)E(τN)|=0.\displaystyle\lim_{N\to\infty}|E(\tau)-E(\tau_{N})|=0.
Remark 5.1.

To obtain convergence of higher moments of the first passages time, we need only replace conditions E(τK)<E(\tau_{K})<\infty and 0g(t)𝑑t<\int_{0}^{\infty}g(t)dt<\infty with

E(τKn)<and0g(t1n)𝑑t<,\displaystyle E(\tau^{n}_{K})<\infty\quad\text{and}\quad\int_{0}^{\infty}g\left(t^{\frac{1}{n}}\right)dt<\infty, (21)

respectively.

We now show that if a Lyapunov function exists for the original system, the conditions in Theorem 5.2 and Theorem 5.3 hold. The Lyapunov function approach has been proposed by Meyn and Tweedie [31] and it has been used to study long-term dynamics of Markov processes [6, 16, 4, 17] especially exponential ergodicity. Gupta et al [16] uses a linear Lyapunov function to show that the stationary distribution of an sFSP model converges to a stationary distribution of the original stochastic model and use the Lyapunov function explicitly compute the convergence rate. In particular, we show Lyapunov functions exist for the examples we consider in Section 7.

Theorem 5.4.

Suppose there exists a function VV and positive constants CC and DD such that for all 𝐱\mathbf{x},

  1. 1.

    V(𝐱)1V(\mathbf{x})\geq 1 for all 𝐱𝕊\mathbf{x}\in\mathbb{S},

  2. 2.

    VV is an increasing function in the sense that

    V(𝐱N+1)V(𝐱N)V(\mathbf{x}_{N+1})\geq V(\mathbf{x}_{N})

    for each 𝐱N+1𝕊N+1𝕊N\mathbf{x}_{N+1}\in\mathbb{S}_{N+1}\setminus\mathbb{S}_{N} and 𝐱N𝕊N\mathbf{x}_{N}\in\mathbb{S}_{N}, and

  3. 3.

    ννλνν(𝐱)(V(𝐱+νν)V(𝐱))\displaystyle\sum\limits_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-V(\mathbf{x}))

    CV(𝐱)+D.\leq-CV(\mathbf{x})+D.

Then the conditions in Theorem 5.3 hold.

Remark 5.2.

The conditions (21) hold if a Lyapunov function satisfying the conditions in Theorem 5.4 exists. Thus, the convergence of the higher moments of the first passage time also follows.

6 Inheritance of slack networks

As we showed in Section 4.2, not all state space truncations preserve accessibility of states in the original system. (For the example in Section 4.2, the truncation created a new absorbing state.) Thus, it is desirable to obtain reduced models that are guaranteed to maintain the accessibility of the original system to predetermined target states. In this Section, we show that under mild conditions, both a regular slack system and an optimized slack system preserve the original accessibility. The proofs of the theorems introduced in this section are in Appendix C. The key to these results is the condition of weak reversibility.

Definition 6.1.

A reaction network is weakly reversible if each connected component of the network is strongly connected. That is, if there is a path of reactions from a complex ν\nu to ν\nu^{\prime}, then there is a path of reactions ν\nu^{\prime} to ν\nu.

We note that the weakly reversible condition applies to the network graph of the CRN. The network graph consists of complexes (nodes) and reactions (edges). It is a sufficient condition for irreducibility of the associated mass-action stochastic process. Indeed, the sufficiency of weak reversibility holds even under general kinetics as long as condition (4) is satisfied.[33] Hence irreducibility of a regular slack network follows since it preserves weak reversibility of the original network and the kinetics modeling the regular slack system satisfies (4).

Corollary 6.1.

Let (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) be a weakly reversible CRN with intensity functions Λ={λyy}\Lambda=\{\lambda_{y\to y^{\prime}}\} satisfying (4). Then the state space of the associated stochastic process with a regular slack network (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) is a union of closed communication classes for any NN.

In case the original network is not weakly reversible, we can still guarantee that optimized slack systems have the same accessibility as the original system provided all species have a degradation reaction (SiS_{i}\to\emptyset).

Theorem 6.1.

Let (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) be a reaction network such that {Si:Si𝒮}\{S_{i}\to\emptyset:S_{i}\in\mathcal{S}\}\subset\mathcal{R}. Suppose that the stochastic process XX associated with (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) and beginning at the point 𝐱0\mathbf{x}_{0} is irreducible. Let XNX_{N} be the stochastic process associated with an optimized slack network (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) such that XN(0)=𝐱0X_{N}(0)=\mathbf{x}_{0} for every NN large enough. Then, for any subset KK of the state space of XX, there exists N0N_{0} such that XNX_{N} reaches KK almost surely for NN0N\geq N_{0}.

This theorem follows from the fact that a slack system only differs from the original system when its runs out of slack reactants. However, in an optimized slack system, degradation reactions are allowable with no slack reactants. Hence, our proof of Theorem 6.1 relies on the presence of all degradation reactions.

A slack network may also inherit its stationary distribution from the original reaction system. When the original system admits a stationary distribution of a product form of Poissons under the complex balance condition, a slack system inherits the same form of the stationary distribution as well. A reaction system is complex balanced if the associated deterministic mass-action system admits a steady state cc^{*}, such that

ν𝒞ννfν(𝐜)=ν𝒞ννfν(𝐜),\displaystyle\sum_{\begin{subarray}{c}\nu\in\mathcal{C}\\ \nu\to\nu^{\prime}\in\mathcal{R}\end{subarray}}f_{\nu}(\mathbf{c}^{*})=\sum_{\begin{subarray}{c}\nu^{\prime}\in\mathcal{C}\\ \nu\to\nu^{\prime}\in\mathcal{R}\end{subarray}}f_{\nu^{\prime}}(\mathbf{c}^{*}),

where fν(x)=x1ν1xdνdf_{\nu}(x)=x_{1}^{\nu_{1}}\cdots x_{d}^{\nu_{d}} is the deterministic mass-action rate [12]. If a reaction system is complex balanced, then its associated stochastic mass-action system admits a stationary distribution corresponding to a product of Poisson distributions centered at the complex balance steady state.[3]. The following lemma show that the complex balancing of the original network is inherited by a regular slack network.

Lemma 6.1.

Suppose that (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) is a reaction network whose mass-action deterministic model admits a complex balanced steady state cc^{*}. Then any regular slack network (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) with slack reactants Y1,,YmY_{1},\cdots,Y_{m} also admits a complex balanced steady state at c~=(c,1,1,,1)\tilde{c}=(c^{*},1,1,\dots,1)^{\top}.

Remark 6.1.

Note that a regular slack network also preserves the deficiency of the original network. Deficiency δ\delta of a reaction network is an index such that

δ=ns,\displaystyle\delta=n-\ell-s,

where nn is the number of the complexes, \ell is the number of connected components and ss is the rank of the stoichiometry matrix of the reaction network. Deficiency characterizes the connectivity of the network structure, and surprisingly it can also determine the long-term behavior of the system dynamics regardless of the system parameters.[12, 20, 3] A regular slack network and original network have the same number of complexes nn, and the same connectivity matrix SS, which implies they have the same number of connected components \ell. Furthermore, using the notation from Section 3.2, the stoichiometry matrices are Γ=CS\Gamma=CS for the original network and

Γ~=(CUAC)S=(CSWCS)\widetilde{\Gamma}=\begin{pmatrix}C\\ U-AC\end{pmatrix}S=\begin{pmatrix}CS\\ -WCS\end{pmatrix}

for a slack network, which means that they have the same rank ss. Together, these imply that the original network and its regular slack network have the same deficiency.

Since the complex balancing is inherited with the same steady state values for XiX_{i}, we have the following stochastic analog of inheritance of the Poissonian stationary distribution for regular slack systems.

Theorem 6.2.

Let XX be the stochastic mass-action system associated with a complex balanced (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) with an initial condition X(0)=𝐱0X(0)=\mathbf{x}_{0}. Let XNX_{N} be the stochastic system associated with a regular slack system (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) with XN(0)=𝐱0X_{N}(0)=\mathbf{x}_{0}. Then, for the state space 𝕊N\mathbb{S}_{N} of XNX_{N}, there exists a constant MN>0M_{N}>0 such that

πN(𝐱)=MNπ(𝐱)for 𝐱𝕊N,\displaystyle\pi_{N}(\mathbf{x})=M_{N}\pi(\mathbf{x})\quad\text{for $\mathbf{x}\in\mathbb{S}_{N}$,}

where π\pi and πN\pi_{N} are the stationary distributions of XX and XNX_{N}, respectively.

We demonstrate Lemma 6.1 and Theorem 6.2 with a simple example.

Example 6.1.

Consider two networks,

X210,\displaystyle X\xrightleftharpoons[2]{1}0, (22)
X21Y.\displaystyle X\xrightleftharpoons[2]{1}Y. (23)

Let XX and XNX_{N} be systems (22) and (23), respectively, where NN is the conservation quantity XN(t)NX_{N}(t)\leq N. Under mass-action kinetics, the complex balance steady state of (22) is c=2c^{*}=2. Under mass-action kinetics, c~=(2,1)\tilde{c}=(2,1) is a complex balance steady state of (23).

Now let π\pi and πN\pi_{N} be the stationary distribution of XX and XNX_{N}, respectively. By Theorem 6.4, Anderson et al [3], π\pi is a product form of Poissons such that

π(x)=e2cxx!for each state x.\displaystyle\pi(x)=e^{-2}\frac{{c^{*}}^{x}}{x!}\quad\text{for each state $x$}.

By plugging π\pi into the chemical master equation 3 of XNX_{N} and showing that for each 𝐱\mathbf{x}

λXYN(𝐱+1)π(𝐱+1)+λYXN(𝐱1)π(𝐱1)\displaystyle\lambda^{N}_{X\to Y}(\mathbf{x}+1)\pi(\mathbf{x}+1)+\lambda^{N}_{Y\to X}(\mathbf{x}-1)\pi(\mathbf{x}-1)
=(x+1)𝟙{Nx10}e22x+1(x+1)!+2𝟙{Nx+11}e22x1(x1)!\displaystyle=(x+1)\mathbbm{1}_{\{N-x-1\geq 0\}}e^{-2}\frac{2^{x+1}}{(x+1)!}+2\mathbbm{1}_{\{N-x+1\geq 1\}}e^{-2}\frac{2^{x-1}}{(x-1)!}
=x𝟙{Nx0}e22xx!+2𝟙{Nx1}e22xx!\displaystyle=x\mathbbm{1}_{\{N-x\geq 0\}}e^{-2}\frac{2^{x}}{x!}+2\mathbbm{1}_{\{N-x\geq 1\}}e^{-2}\frac{2^{x}}{x!}
=λXYN(𝐱)π(𝐱)+λYXN(𝐱)π(𝐱)\displaystyle=\lambda^{N}_{X\to Y}(\mathbf{x})\pi(\mathbf{x})+\lambda^{N}_{Y\to X}(\mathbf{x})\pi(\mathbf{x})

we can verify that π\pi is a stationary solution of the chemical master equation 3 of XNX_{N}. Since the state space of XNX_{N} is {x0:xN}\{x\in\mathbb{Z}_{\geq 0}:x\leq N\}, we choose a constant MNM_{N} such that

0xNMNπ(x)=1.\displaystyle\sum_{0\leq x\leq N}M_{N}\pi(x)=1.

Then πN=MNπ\pi_{N}=M_{N}\pi is the stationary distribution of XNX_{N}.

Refer to caption
Figure 4: Calculating Mean Time to Extinction of a Lotka-Volterra Model with Migration Using Slack Reactants. A. The reaction network for the Lotka-Volterra model with migration (left). The Lotka-Volterra model with a slack reactant YY (right). The parameters are κ1=0.1,κ2=0.1,κ3=0.2,κ4=0.6,κ5=0.2\kappa_{1}=0.1,\kappa_{2}=0.1,\kappa_{3}=0.2,\kappa_{4}=0.6,\kappa_{5}=0.2 and κ6=0.2\kappa_{6}=0.2. B. Convergence of the mean time to extinction of the slack network (blue) to the true network (solid red). C. Heatmaps of the probability density at time 10001000 of the original model and the slack system with various truncation size.

7 Applications of Slack Networks

In this Section, we demonstrate the utility of slack reactants in computing mean first-passage times for two biological examples. For both examples, we compute the mean first passage time via the matrix inversion approach as shown in Appendix A.

7.1 A Lotka-Volterra model with migration

Consider a Lotka-Volterra model with migration shown in Figure 4A. In this model, species AA is the prey, and species BB is the predator. Clearly, the state space this model is infinite (A,B)(A,B) such that A0A\geq 0, B0B\geq 0. We will use slack reactants to determine the expected time to extinction of either species. More specifically, let K={(A,B):A=0 or B=0}K=\{(A,B):A=0\text{ or }B=0\}. We will calculate the mean first arrival time to KK from an initial condition (A(0),B(0))(A(0),B(0)). (In our simulations in Figure 4, we chose (A(0),B(0))=(3,3)(A(0),B(0))=(3,3).)

To generate our slack network, we choose a conservative bound w(A,B)Nw\cdot(A,B)^{\top}\leq N with w=(1,1)w=(1,1). As we discussed in Section 3.3, this ww minimizes the intrusiveness of slack reactants because the number of reactions νν\nu\to\nu^{\prime} such that (νν)w>0(\nu^{\prime}-\nu)\cdot w>0 is minimized. By using the algorithm introduced in Section 3.2, we generate a regular slack network (LABEL:eq:regular_slack) with a slack reactant YY,

B+Yκ2κ12Yκ4κ3A+Yκ52A,A+Bκ62B.\displaystyle\begin{split}&B+Y\xrightleftharpoons[\kappa_{2}]{\kappa_{1}}2Y\xrightleftharpoons[\kappa_{4}]{\kappa_{3}}A+Y\xrightarrow{\kappa_{5}}2A,\\ &A+B\xrightarrow{\kappa_{6}}2B.\end{split} (24)

As the slack reactant YY in reactions B+Y2YA+YB+Y\rightleftharpoons 2Y\rightleftharpoons A+Y can be canceled, we further generate the optimized slack network shown in Figure 4A. We let A(0)+B(0)+Y(0)=NA(0)+B(0)+Y(0)=N, which is the conservation quantity of the new network.

Let τ\tau be the first passage time from our initial condition to KK. First, we examine the accessibility of the set KK. Because our reaction network contains creation and destruction of all species (i.e., BAB\rightleftharpoons\emptyset\rightleftharpoons A) the original model is irreducible and any state is accessible to KK. Furthermore, Theorem 6.1 guarantees that the stochastic model associated with the optimized slack network is also accessible to KK from any state.

Next, by showing there exists a Lyapunov function satisfying the condition of Theorem 5.4 for the original model, we are guaranteed the first passages times from our slack network will converge to the true first passage times. (See Appendix E.1 for more detail.) Therefore, as the plot shows in Figure 4B, the mean first extinction time of the slack network converges to that of the original model, as NN increases. The mean first passage time of the original model was obtained by averaging 10910^{9} sample trajectories. These trajectories were computed serially on a commodity machine and took 4.64.6 hours to run. In contrast, the mean first passage times of the slack systems were computed analytically on the same computer and took at most 13 seconds. Figure 4 also shows that using only 10310^{3} samples is misleading as the simulation average has not yet converged to the true mean first passage time. Finally, as expected from Theorem 5.1 the probability density of the slack network converges to that of the original network (see Figure 4C).

7.2 Protein synthesis with slow toggle switch

We now consider a protein synthesis model with a toggle switch, see Figure 5A. Protein species XX and ZZ may be created, but only when their respective genes DXD^{X} or DZD^{Z} are in the active (unoccupied) state, D0XD^{X}_{0} and D0ZD^{Z}_{0}. Each protein acts as a repressor for the other by binding at the promoter of the opposite gene and forcing it into the inactive (occupied) state (D1XD^{X}_{1} and D1ZD^{Z}_{1}). In this system we consider only one copy of each gene, so that, D0X+D1X=D0Z+D1Z=1D^{X}_{0}+D^{X}_{1}=D^{Z}_{0}+D^{Z}_{1}=1 for all time. Thus, we focus primarily on the state space of protein numbers only (X,Z)(X,Z).

The deterministic form of such systems are often referred to as a “bi-stable switch” as it is characterized by steady states (X,0)(X^{*},0) and (XX “on” and ZZ “off”) and (0,Z)(0,Z^{*}) (XX “off” and ZZ “on”). This stochastic form of toggle switch has been shown to exhibit a multi-modal long-term probability density due to switches between these two deterministic states due to rapid significant changes in the numbers of proteins XX and ZZ by synthesis or degradation (depending on the state of promoters) [1]. Figure 5C shows that the associated stochastic system admits a tri-modal long-term probability density. Thus the system transitions from one mode to other modes and, for the kinetic parameters chosen in Figure 5, rarely leaves the region R={(X,Z)|0X30 or 0Z30}R=\{(X,Z)|0\leq X\leq 30\text{ or }0\leq Z\leq 30\}. Significant departures from a stable region of a genetic switch may be associated with irregular and diseased cell fates. As such, the first passage time of this system outside of RR may indicate the appearance of an unexpected phenotype in a population. Because this event is rare, estimating first passage times with direct stochastic simulations, such as with the Gillespie algorithm [15], will be complicated by the long time taken to exit the region.

As in the previous example, slack systems provide a valuable tool for direct calculation of mean first passage times. In this example, we consider the time a trajectory beginning at state (X,Z,D0X,D0Z)=(0,0,1,1)(X,Z,D^{X}_{0},D^{Z}_{0})=(0,0,1,1) enters the target set K={(X,Z)|X>30 and Z>30}=RcK=\{(X,Z)|X>30\text{ and }Z>30\}=R^{c} and compute τ\tau, the first passage time to KK.

Since the species corresponding to the status of promoters (D0X,D1X,D0ZD^{X}_{0},D^{X}_{1},D^{Z}_{0} and D1ZD^{Z}_{1}) are already bounded, we use the conservative bound X+ZNX+Z\leq N to generate a regular slack network in Figure 5B with the algorithm introduced in Section 3.2. The original toggle switch model is irreducible (because of the degradation X0X\to 0, Z0Z\to 0 and protein synthesis Z+D0XD1XZ+D^{X}_{0}\leftarrow D^{X}_{1}, X+D0ZD1ZX+D^{Z}_{0}\leftarrow D^{Z}_{1} reactions). Moreover by Theorem 6.1, the degradation reactions guarantee that the slack system is also accessible to KK from any state.

As shown in Figure 5D, the mean first passage time of the slack system appears to be converting to approximately 3.171×1093.171\times 10^{9}. To prove that the limit of the slack system is actually the original mean first passage time, we construct a Lyapunov functions satisfying the conditions of Theorem 5.4. See Appendix E.2 for more details about the construction of the Lyapunov function.

Refer to caption
Figure 5: Calculating Mean First Passage Time of a Slow Toggle Switch Using Slack Reactants. A. Protein synthesis with slow toggle switch. B. The toggle switch model with a slack reactant YY. Parameters are α=0.1,α=10,κ=2000\alpha-=0.1,\alpha=10,\kappa=2000 and κ=100\kappa-=100. C. Multimodal long-term probability distribution of the original model. D. Convergence of the first passage time of the slack system with various truncation sizes.

8 Discussion & Conclusions

We propose a new state space truncation method for stochastic reaction networks (see Section 3). In contrast to other methods, such as FSP, sFSP and the finite buffer method, we truncate the state space indirectly by expanding the original chemical reaction network to include slack reactants. The truncation is imposed through user defined conservation relations among species. The explicit network structure of slack reactants facilitates proofs of convergence (Section 5) and allows the use of existing software packages to study the slack network itself [23]. Indeed, any user-defined choices for conservation laws, conservation amounts, and stochiometric structure, can be used to construct a slack network with our algorithm. We provide guidelines for optimal user choices that can increase the similarity between the slack system and the original model (see Section 3.3).

Slack systems can be used to estimate the dynamical behavior of the original stochastic model. In Section 4, we used a simple example to show that the slack method can lead to a better approximation for the mean first passage time than the sFSP method and the finite buffer method. In particular, in Section 5 we provide theorems that show the slack system approximates the probability density and the mean first passage time of the original system. Because slack networks preserve network properties, such as weak reversibility, the slack system is also likely to have the same accessibility to a target state as the original model (see Section 6). In particular, we note that weak reversibility guarantees our slack truncation does not introduce absorbing states.

In Section 6, we show that this truncation method is natural in the sense that the stationary distributions of the original and slack system are identical up to multiplication by a constant when the original system is complex balanced. Finally, in Section 7, we use slack networks to calculate first passage times for two biological examples. We expect that this new theoretical framework for state space truncation will be useful in the study of other biologically motivated stochastic chemical reaction systems.

ACKNOWLEDGMENTS

This work is partially supported by NSF grants DMS1616233, DMS1763272, BCS1344279, Simons Foundation grant 594598 and the by the Joint DMS/NIGMS Initiative to Support Research at the Interface of the Biological and Mathematical Sciences (R01-GM126548).

Data Availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Appendix A First passage time for Markov processes with finitely many states

The Markov chain associated with a slack network has always a finite state space. There are many different methods to analytically derive the mean first passage time of a Markov chain with a finite state space [22, 21, 37, 9]. In this paper, we use the method of Laplace transform which is also used in [37].

For a continuous time Markov process, let 𝕊\mathbb{S} be the finite state space and let QQ be the transition rate matrix, i.e. QijQ_{ij} is the transition rate from state ii to state jj if iji\neq j and Qii=jiQijQ_{ii}=-\sum_{j\neq i}Q_{ij}.

For a subset K={i1,i2,,ik}𝕊K=\{i_{1},i_{2},\dots,i_{k}\}\subset\mathbb{S}, we define a truncated transition matrix QKQ_{K} that is obtained by removing the iji_{j} th row and column from QQ for j=1,2,,kj=1,2,\dots,k. Then the mean first passage time to set KK starting from the ii-th state is the ii-th entry of QK1𝟏-Q_{K}^{-1}\mathbf{1}, where 𝟏\mathbf{1} is a column vector with each entry 11.

Appendix B Proofs of Convergence Theorems

In this section, we provide the proofs of the theorems introduced in Section 5. We use the same notations and the same assumptions as we used in Section 5.

Proof of Theorem 5.1: We employ the main idea shown in the proof of Theorem 2.1 in Hart and Tweedie [18]. Let a state 𝐱\mathbf{x} and time tt be fixed. We consider large enough NN so that 𝐱𝕊N1\mathbf{x}\in\mathbb{S}_{N-1}.

We use an FSP model on 𝕊N1\mathbb{S}_{N-1} of the original system XX with the designated absorbing state 𝐱𝕊N1c\mathbf{x}^{*}\in\mathbb{S}_{N-1}^{c}. Let pN1FSPp^{\text{FSP}}_{N-1} be the probability density function of this FSP model.

Let TNT_{N} be the first time for XNX_{N} to hit 𝕊N𝕊N1\mathbb{S}_{N}\setminus\mathbb{S}_{N-1}. We generate a coupling of XNX_{N} and the FSP model restricted on 𝕊N1\mathbb{S}_{N-1} as they move together by TNT_{N} and they move independently after TNT_{N}. Then pN1FSP(𝐱,t)=P(XN(t)=𝐱,TN>t)p^{\text{FSP}}_{N-1}(\mathbf{x},t)=P(X_{N}(t)=\mathbf{x},T_{N}>t) for 𝐱𝕊N1\mathbf{x}\in\mathbb{S}_{N-1} because the FSP model has stayed in 𝕊N1\mathbb{S}_{N-1} if and only if XNX_{N} has never touched 𝕊N𝕊N1\mathbb{S}_{N}\setminus\mathbb{S}_{N-1}. Thus

pN(𝐱,t)\displaystyle p_{N}(\mathbf{x},t) =P(XN(t)=𝐱,t<TN)+P(XN(t)=𝐱,tTN)\displaystyle=P(X_{N}(t)=\mathbf{x},t<T_{N})+P(X_{N}(t)=\mathbf{x},t\geq T_{N})
=pN1FSP(𝐱,t)+P(XN(t)=𝐱,TNt)\displaystyle=p^{\text{FSP}}_{N-1}(\mathbf{x},t)+P(X_{N}(t)=\mathbf{x},T_{N}\leq t) (25)
pN1FSP(𝐱,t).\displaystyle\geq p^{\text{FSP}}_{N-1}(\mathbf{x},t).

Furthermore

P(XN(t)=𝐱,tTN)P(tTN)\displaystyle P(X_{N}(t)=\mathbf{x},t\geq T_{N})\leq P(t\geq T_{N})
=pN1FSP(𝐱,t)=1pN1FSP(𝕊N1,t),\displaystyle=p^{\text{FSP}}_{N-1}(\mathbf{x}^{*},t)=1-p^{\text{FSP}}_{N-1}(\mathbb{S}_{N-1},t),

where we used the fact that after TNT_{N}, the FSP process is absorbed at 𝐱\mathbf{x}^{*}. Thus

pN1FSP(𝐱,t)pN(𝐱,t)pN1FSP(𝐱,t)+1pN1FSP(𝕊N1,t).\displaystyle p^{\text{FSP}}_{N-1}(\mathbf{x},t)\leq p_{N}(\mathbf{x},t)\leq p^{\text{FSP}}_{N-1}(\mathbf{x},t)+1-p^{\text{FSP}}_{N-1}(\mathbb{S}_{N-1},t). (26)

Note that

pN1FSP(𝐱,t)=P(X(t)=𝐱,t<TN)p(𝐱,t).\displaystyle p^{\text{FSP}}_{N-1}(\mathbf{x},t)=P(X(t)=\mathbf{x},t<T_{N})\leq p(\mathbf{x},t).

Since TNT_{N} increases to \infty almost surely, as NN increases, pN1FSP(𝐱,t)p^{\text{FSP}}_{N-1}(\mathbf{x},t) monotonically increases in NN and converges to p(𝐱,t)p(\mathbf{x},t), as NN\to\infty for each 𝐱𝕊N\mathbf{x}\in\mathbb{S}_{N}. Then by using the monotone convergence theorem, the term

pN1FSP(𝕊N1,t)=x𝕊pN1FSP(x,t)𝟙{x𝕊N1}p^{\text{FSP}}_{N-1}(\mathbb{S}_{N-1},t)=\sum_{x\in\mathbb{S}}p^{\text{FSP}}_{N-1}(x,t)\mathbbm{1}_{\{x\in\mathbb{S}_{N-1}\}}

converges to p(𝕊,t)=1p(\mathbb{S},t)=1. Therefore, by taking limN\displaystyle\lim\limits_{N\to\infty} in both sides of (26) the result follows. \square

Proof of Theorem 5.2: This proof is a slight generalization of the proof of Theorem 3.3 in Hart and Tweedie [18]. Since the convergence of pN(,t)πN1\|p_{N}(\cdot,t)-\pi_{N}\|_{1} is independent of NN, for any ϵ>0\epsilon>0, we choose sufficiently large t0t_{0} such that

p(,t0)π(t0)1ϵandpN(,t0)πN(t0)1ϵ\displaystyle\|p(\cdot,t_{0})-\pi(t_{0})\|_{1}\leq\epsilon\quad\text{and}\quad\|p_{N}(\cdot,t_{0})-\pi_{N}(t_{0})\|_{1}\leq\epsilon

for all NN. Then by using the triangle inequalities

ππN1\displaystyle\|\pi-\pi_{N}\|_{1} p(,t0)π1+pN(,t0)πN1\displaystyle\leq\|p(\cdot,t_{0})-\pi\|_{1}+\|p_{N}(\cdot,t_{0})-\pi_{N}\|_{1}
+p(,t0)pN(,t0)1\displaystyle\ \ +\|p(\cdot,t_{0})-p_{N}(\cdot,t_{0})\|_{1}
p(,t0)π1+pN(,t0)πN1\displaystyle\leq\|p(\cdot,t_{0})-\pi\|_{1}+\|p_{N}(\cdot,t_{0})-\pi_{N}\|_{1}
+p(,t0)pN1FSP(,t0)1\displaystyle\ \ +\|p(\cdot,t_{0})-p^{\text{FSP}}_{N-1}(\cdot,t_{0})\|_{1}
+pN1FSP(,t0)pN(,t0)1\displaystyle\ \ +\|p^{\text{FSP}}_{N-1}(\cdot,t_{0})-p_{N}(\cdot,t_{0})\|_{1}
2ϵ+𝐱𝕊|p(𝐱,t0)pN1FSP(𝐱,t0)|+x𝕊|pN1FSP(𝐱,t0)pN(𝐱,t0)|.\displaystyle\begin{split}&\leq 2\epsilon+\sum_{\mathbf{x}\in\mathbb{S}}|p(\mathbf{x},t_{0})-p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})|\\ &\ \ +\sum_{x\in\mathbb{S}}|p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})-p_{N}(\mathbf{x},t_{0})|.\end{split} (27)

Note that as we mentioned in the proof of Theorem 5.1, we have monotone convergence of pN1FSP(𝐱,t0)p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0}) to p(𝐱,t0)p(\mathbf{x},t_{0}) for each 𝐱𝕊N1\mathbf{x}\in\mathbb{S}_{N-1}, as NN\to\infty. Hence by the monotone convergence theorem, the first summation in (LABEL:eq:conv_of_mfpt_final) goes to zero, as NN\to\infty. Note further that from (25) we have

|pN1FSP(𝐱,t0)pN(𝐱,t0)|=P(XN(t)=𝐱,TNt0).|p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})-p_{N}(\mathbf{x},t_{0})|=P(X_{N}(t)=\mathbf{x},T_{N}\leq t_{0}).

Hence the second summation in (LABEL:eq:conv_of_mfpt_final) satisfies

x𝕊|pN1FSP(𝐱,t0)pN(𝐱,t0)|\displaystyle\sum_{x\in\mathbb{S}}|p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})-p_{N}(\mathbf{x},t_{0})| x𝕊N1|pN1FSP(𝐱,t0)pN(𝐱,t0)|\displaystyle\leq\sum_{x\in\mathbb{S}_{N-1}}|p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})-p_{N}(\mathbf{x},t_{0})|
+pN1FSP(x,t0)\displaystyle\ \ +p^{\text{FSP}}_{N-1}(x^{*},t_{0})
+P(XN(t)𝕊N𝕊N1).\displaystyle\ \ +P(X_{N}(t)\in\mathbb{S}_{N}\setminus\mathbb{S}_{N-1}).

Note that by (25)

x𝕊N1|pN1FSP(𝐱,t0)pN(𝐱,t0)|\displaystyle\sum_{x\in\mathbb{S}_{N-1}}|p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})-p_{N}(\mathbf{x},t_{0})|
=x𝕊N1P(XN(t0)=𝐱,t0TN)=P(t0TN)\displaystyle=\sum_{x\in\mathbb{S}_{N-1}}P(X_{N}(t_{0})=\mathbf{x},t_{0}\geq T_{N})=P(t_{0}\geq T_{N})

Furthermore pN1FSP(x,t0)=P(t0TN)p^{\text{FSP}}_{N-1}(x^{*},t_{0})=P(t_{0}\geq T_{N}) and P(XN(t)𝕊N𝕊N1)=P(TN<t0)P(X_{N}(t)\in\mathbb{S}_{N}\setminus\mathbb{S}_{N-1})=P(T_{N}<t_{0}). Hence x𝕊|pN1FSP(𝐱,t0)pN(𝐱,t0)|0\sum_{x\in\mathbb{S}}|p^{\text{FSP}}_{N-1}(\mathbf{x},t_{0})-p_{N}(\mathbf{x},t_{0})|\to 0, as NN\to\infty because TNT_{N}\to\infty almost surely, as NN\to\infty.

Consequently, we have

limNππN12ϵ.\displaystyle\lim_{N\to\infty}\|\pi-\pi_{N}\|_{1}\leq 2\epsilon.

Since we choose an arbitrary ϵ\epsilon, the completes the proof. \square

In order to prove Theorem 5.3, we consider an ‘absorbing’ Markov process associated with XX and XNX_{N}. Let X¯\bar{X} and X¯N\bar{X}_{N} be Markov processes such that

X¯(t)={X(t) if t<τ,X(τ) if tτ.\displaystyle\bar{X}(t)=\begin{cases}X(t)&\text{ if $t<\tau$,}\\ X(\tau)&\text{ if $t\geq\tau$}.\end{cases}
X¯N(t)={XN(t) if t<τN,XN(τN) if tτN.\displaystyle\bar{X}_{N}(t)=\begin{cases}X_{N}(t)&\text{ if $t<\tau_{N}$,}\\ X_{N}(\tau_{N})&\text{ if $t\geq\tau_{N}$}.\end{cases}

That is, X¯\bar{X} and X¯N\bar{X}_{N} are coupled process to XX and XNX_{N}, respectively. Furthermore they are absorbed to KK once the coupled process (XX for X¯\bar{X} and XNX_{N} for X¯N\bar{X}_{N}) visits the set KK. These coupled process have the following relation.

Lemma B.1.

Let t0t\geq 0 be fixed and ϵ>0\epsilon>0 be arbitrary. Suppose XX admits a stationary distribution π\pi. Suppose further that limNππN1=0\displaystyle\lim\limits_{N\to\infty}\|\pi-\pi_{N}\|_{1}=0. Then there exists a finite subset K¯\bar{K} such that P(X¯(t)K¯c)<ϵP(\bar{X}(t)\in\bar{K}^{c})<\epsilon and P(X¯N(t)K¯c)<ϵP(\bar{X}_{N}(t)\in\bar{K}^{c})<\epsilon for sufficiently large NN.

Proof.

Since the probabilities of X¯\bar{X} and X¯N\bar{X}_{N} are leaking to KK, we have for each tt and for each 𝐱Kc\mathbf{x}\in K^{c}

P(X¯(t)=𝐱)P(X(t)=𝐱),andP(X¯N(t)=𝐱)<P(XN(t)=𝐱).\displaystyle\begin{split}&P(\bar{X}(t)=\mathbf{x})\leq P(X(t)=\mathbf{x}),\quad\text{and}\\ &P(\bar{X}_{N}(t)=\mathbf{x})<P(X_{N}(t)=\mathbf{x}).\end{split} (28)

This can be formally proved as for each xKcx\in K^{c}

P(X(t)=𝐱)\displaystyle P(X(t)=\mathbf{x}) =P(X(t)=𝐱,τ>t)+P(X(t)=𝐱,τt)\displaystyle=P(X(t)=\mathbf{x},\tau>t)+P(X(t)=\mathbf{x},\tau\leq t)
=P(X¯(t)=𝐱)+P(X(t)=𝐱,τKt)\displaystyle=P(\bar{X}(t)=\mathbf{x})+P(X(t)=\mathbf{x},\tau^{K}\leq t)
P(X¯(t)=𝐱).\displaystyle\geq P(\bar{X}(t)=\mathbf{x}).

In the same way, we can prove that P(X¯N(t)=𝐱)<P(XN(t)=𝐱)P(\bar{X}_{N}(t)=\mathbf{x})<P(X_{N}(t)=\mathbf{x}) for each 𝐱Kc\mathbf{x}\in K^{c}.

Let ϵ>0\epsilon^{\prime}>0 be arbitrary. Then limNππN1=0\displaystyle\lim\limits_{N\to\infty}\|\pi-\pi_{N}\|_{1}=0 implies that for any subset UU, we have

πN(U)π(U)+ϵandπ(U)πN(U)+ϵ,\displaystyle\pi_{N}(U)\leq\pi(U)+\epsilon^{\prime}\quad\text{and}\quad\pi(U)\leq\pi_{N}(U)+\epsilon^{\prime}, (29)

for sufficiently large NN. We use this property and (LABEL:eq:two_probs_bounded) combined with the ‘monotonicity’ of the chemical master equation to show the result.

First of all, note that we are assuming X(0)=XN(0)=𝐱0X(0)=X_{N}(0)=\mathbf{x}_{0}. Hence there is a constant γ1>0\gamma_{1}>0 such that

P(X(0)=𝐱)=𝟙{𝐱=𝐱0}γπ(𝐱)P(X(0)=\mathbf{x})=\mathbbm{1}_{\{\mathbf{x}=\mathbf{x}_{0}\}}\leq\gamma\pi(\mathbf{x})

for all 𝐱\mathbf{x}. Moreover by choosing sufficiently small ϵ\epsilon^{\prime} in (29), we have the minimum minNπN(𝐱0)π(𝐱0)ϵ>0\min_{N}\pi_{N}(\mathbf{x}_{0})\geq\pi(\mathbf{x}_{0})-\epsilon^{\prime}>0. Thus there exists γ2\gamma_{2} such that

P(XN(0)=𝐱)=𝟙{𝐱=𝐱0}γ2πN(𝐱),P(X_{N}(0)=\mathbf{x})=\mathbbm{1}_{\{\mathbf{x}=\mathbf{x}_{0}\}}\leq\gamma_{2}\pi_{N}(\mathbf{x}),

for all sufficiently large NN and for each 𝐱\mathbf{x}. Note that the chemical master equation is a system of ordinary differential equations, and P(X(t)=)P(X(t)=\cdot) and π\pi are the solution to the system with the initial conditions P(X(0)=)P(X(0)=\cdot) and π\pi. Hence the monotonicity of a system of ordinary differential equations, it follows that

P(X(t)=x)γ1π(x)for any x and t.\displaystyle P(X(t)=x)\leq\gamma_{1}\pi(x)\quad\text{for any $x$ and $t$}. (30)

In the same way, for any NN it follows that

P(XN(t)=x)γ2πN(x)for any x and t.\displaystyle P(X_{N}(t)=x)\leq\gamma_{2}\pi_{N}(x)\quad\text{for any $x$ and $t$}. (31)

The detailed proof about the monotonicity of the chemical master equation is shown in Lemma 3.2 of Enciso and Kim. [10]

Secondly, there exists a finite set K¯\bar{K} such that π(K¯c)<ϵ\pi(\bar{K}^{c})<\epsilon^{\prime} because π\pi is a probability distribution, Furthermore by (29), πN(K¯c)π(K¯c)+ϵ<2ϵ\pi_{N}(\bar{K}^{c})\leq\pi(\bar{K}^{c})+\epsilon^{\prime}<2\epsilon^{\prime} for any sufficiently large NN.

Finally we choose ϵ=ϵ2γ\epsilon^{\prime}=\frac{\epsilon}{2\gamma}, where γ=max{γ1,γ2}\gamma=\max\{\gamma_{1},\gamma_{2}\}. Then by summing up (LABEL:eq:two_probs_bounded), (30) and (31) over xK¯cx\in\bar{K}^{c}, the result follows. ∎

We prove Theorem 5.3 by using the ‘absorbing‘ Markov processes X¯\bar{X} and X¯N\bar{X}_{N} coupled to XX and XNX_{N}, respectively.

Proof of Theorem 5.3: We first break the term P(τ>t)P(τN>t)P(\tau>t)-P(\tau_{N}>t) to show that

limN|P(τ>t)P(τN>t)|=0.\lim_{N\to\infty}|P(\tau>t)-P(\tau_{N}>t)|=0.

Let ϵ>0\epsilon>0 be an arbitrarily small number. Let K¯\bar{K} be a finite set we found in Lemma B.1. Then by using triangular inequalities,

|P(τ>t)P(τN>t)|=|P(X¯(t)Kc)P(X¯N(t)Kc)|𝐱KcK¯|P(X¯(t)=𝐱)P(X¯N(t)=𝐱)|+𝐱KcK¯c|P(X¯(t)=𝐱)P(X¯N(t)=𝐱)|𝐱KcK¯|P(X¯(t)=𝐱)P(X¯N(t)=𝐱)|+P(X¯(t)K¯c)+P(X¯N(t)K¯c)𝐱KcK¯|P(X¯(t)=𝐱)P(X¯N(t)=𝐱)|+2ϵ.\displaystyle\begin{split}&|P(\tau>t)-P(\tau_{N}>t)|\\ &=|P(\overline{X}(t)\in K^{c})-P(\overline{X}_{N}(t)\in K^{c})|\\ &\leq\sum_{\mathbf{x}\in K^{c}\cap\bar{K}}|P(\overline{X}(t)=\mathbf{x})-P(\overline{X}_{N}(t)=\mathbf{x})|\\ &+\sum_{\mathbf{x}\in K^{c}\cap\bar{K}^{c}}|P(\overline{X}(t)=\mathbf{x})-P(\overline{X}_{N}(t)=\mathbf{x})|\\ &\leq\sum_{\mathbf{x}\in K^{c}\cap\bar{K}}|P(\overline{X}(t)=\mathbf{x})-P(\overline{X}_{N}(t)=\mathbf{x})|\\ &\ \ +P(\overline{X}(t)\in\bar{K}^{c})+P(\overline{X}_{N}(t)\in\bar{K}^{c})\\ &\leq\sum_{\mathbf{x}\in K^{c}\cap\bar{K}}|P(\overline{X}(t)=\mathbf{x})-P(\overline{X}_{N}(t)=\mathbf{x})|+2\epsilon.\end{split} (32)

Note that by the same proof of Theorem 5.1, we have the convergence

limN|P(X¯(t)=𝐱)P(X¯N(t)=𝐱)|=0for each 𝐱𝕊.\displaystyle\lim_{N\to\infty}|P(\overline{X}(t)=\mathbf{x})-P(\overline{X}_{N}(t)=\mathbf{x})|=0\quad\text{for each $\mathbf{x}\in\mathbb{S}$}.

Since the summation 𝐱KcK¯\sum_{\mathbf{x}\in K^{c}\cap\bar{K}} is finite, we have that by taking NN\to\infty in (32)

limN|P(τK>t)P(τNK>t)|=0+2ϵ.\displaystyle\lim_{N\to\infty}|P(\tau^{K}>t)-P(\tau^{K}_{N}>t)|=0+2\epsilon. (33)

Now, to show the convergence of the mean first passage times, note that

|E(τ)E(τN)|0|P(τ>t)P(τN>t)|𝑑t,\displaystyle|E(\tau)-E(\tau_{N})|\leq\int_{0}^{\infty}|P(\tau>t)-P(\tau_{N}>t)|dt,

where the integrand is bounded by P(τK>t)+g(t)P(\tau_{K}>t)+g(t).

Condition (iii) in Theorem 5.3 implies that E(τk)<E(\tau_{k})<\infty and 0g(t)𝑑t<\int_{0}^{\infty}g(t)dt<\infty. Hence the dominant convergence theorem and (33) imply that

limN|E(τ)E(τN)|=0.\displaystyle\lim_{N\to\infty}|E(\tau)-E(\tau_{N})|=0.

\square

The existence of a special Lyapunov function ensures that the conditions in Theorem 5.3 hold. In order to use the absorbing Markov processes, as we did in the previous proof, we define a Lyapunov function for X¯\overline{X} based on a given Lyapunov function VV. Let λ¯νν\bar{\lambda}_{\nu\to\nu^{\prime}} and λ¯ννN\bar{\lambda}^{N}_{\nu\to\nu^{\prime}} denote the intensity of a reaction νν\nu\to\nu^{\prime} for X¯\overline{X} and X¯N\overline{X}_{N}, respectively. For a given function VV such that V(𝐱)1V(\mathbf{x})\geq 1 for any 𝐱𝕊\mathbf{x}\in\mathbb{S}, we define V¯\overline{V} such that

V¯(𝐱)={V(𝐱)if 𝐱Kc1if 𝐱K,\displaystyle\overline{V}(\mathbf{x})=\begin{cases}V(\mathbf{x})\quad&\text{if $\mathbf{x}\in K^{c}$}\\ 1\quad&\text{if $\mathbf{x}\in K$},\\ \end{cases}

so that V(𝐱)V(𝐱)V(\mathbf{x})\geq V(\mathbf{x}) for any 𝐱\mathbf{x}.

Note that λ¯νν(x)=λνν(x)\bar{\lambda}_{\nu\to\nu^{\prime}}(x)=\lambda_{\nu\to\nu^{\prime}}(x) if xKcx\in K^{c}. Hence for each 𝐱Kc\mathbf{x}\in K^{c},

ννλ¯νν(𝐱)(V¯(𝐱+νν)V¯(𝐱))ννλνν(𝐱)(V(𝐱+νν)V(𝐱))\displaystyle\begin{split}&\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\bar{\lambda}_{\nu\to\nu^{\prime}}(\mathbf{x})(\overline{V}(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-\overline{V}(\mathbf{x}))\\ &\leq\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-V(\mathbf{x}))\end{split} (34)

because V(𝐱)=V¯(𝐱)V(\mathbf{x})=\overline{V}(\mathbf{x}) and V(x+νν)V¯(x+νν)V(x+\nu^{\prime}-\nu)\geq\overline{V}(x+\nu^{\prime}-\nu) for every reaction x+ννx+\nu^{\prime}-\nu. Moreover, Since λ¯νν(x)=0λνν(x)\bar{\lambda}_{\nu\to\nu^{\prime}}(x)=0\leq\lambda_{\nu\to\nu^{\prime}}(x) if xKx\in K by the definition of X¯\overline{X} for each reaction νν\nu\to\nu^{\prime}. Hence, for each xKx\in K

ννλ¯νν(𝐱)(V¯(𝐱+νν)V¯(𝐱))=0.\displaystyle\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\bar{\lambda}_{\nu\to\nu^{\prime}}(\mathbf{x})(\overline{V}(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-\overline{V}(\mathbf{x}))=0. (35)

From (LABEL:eq:_relation_between_v_and_v_bar) and (35), therefore, we conclude that for any xx

ννλ¯νν(𝐱)(V¯(𝐱+νν)V¯(𝐱)){CV(𝐱)+Dif 𝐱Kc,0if 𝐱KCV¯(x)+D,\displaystyle\begin{split}&\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\bar{\lambda}_{\nu\to\nu^{\prime}}(\mathbf{x})(\overline{V}(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-\overline{V}(\mathbf{x}))\\ &\leq\begin{cases}-CV(\mathbf{x})+D\quad&\text{if $\mathbf{x}\in K^{c}$},\\ 0&\text{if $\mathbf{x}\in K$}\end{cases}\\ &\leq-C^{\prime}\bar{V}(x)+D^{\prime},\end{split} (36)

where C=CC^{\prime}=C and D=max{C+1,D}D^{\prime}=\max\{C+1,D\}.

Proof of Theorem 5.4: In this proof, we show that all the conditions in Theorem 5.3 are met by using the given function VV. First, we show that XX admits a stationary distribution. This follows straightforwardly by Theorem 3.2 in Hard and Tweedie [18] because condition 3 in Theorem 5.4 means that VV is a Lyapunov function for XX.

The condition basically means that every outward reaction νν\nu\to\nu^{\prime} (i.e. 𝐱+νν𝕊Nc\mathbf{x}+\nu^{\prime}-\nu\in\mathbb{S}^{c}_{N} and 𝐱𝕊N\mathbf{x}\in\mathbb{S}_{N}) in \mathcal{R} gives a non-negative drift as V(x+νν)V(x)0V(x+\nu^{\prime}-\nu)-V(x)\geq 0. We use condition 2 in Theorem 5.4 to show that VV is also a Lyapunov function for XNX_{N} for any N.

We denote by λν~ν~N\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}} the intensity of a reaction ν~ν~\tilde{\nu}\to\tilde{\nu}^{\prime} in (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}). Suppose that a reaction ν~ν~~\tilde{\nu}\to\tilde{\nu}^{\prime}\in\widetilde{\mathcal{R}} is obtained from νν\nu\to\nu^{\prime}\in\mathcal{R} by adding a slack reactant. That is, q(ν~ν~)=ννq(\tilde{\nu}^{\prime}-\tilde{\nu})=\nu^{\prime}-\nu, where qq is a projection function such that q(x1,,xd,xd+1,,xd+r)=(x1,,xd)q(x_{1},\dots,x_{d},x_{d}+1,\dots,x_{d+r})=(x_{1},\dots,x_{d}). Then, by the definition (13) of the intensity in a slack network, we have λν~ν~N(𝐱)=0\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})=0 when 𝐱+q(ν~ν~)𝕊N\mathbf{x}+q(\tilde{\nu}^{\prime}-\tilde{\nu})\not\in\mathbb{S}_{N}, because the XNX_{N} is confined in 𝕊N\mathbb{S}_{N}. Furthermore, by condition 2 in Theorem 5.4, V(x+νν)V(x)V(x+\mathbf{\nu}^{\prime}-\mathbf{\nu})\geq V(x) when 𝐱+q(ν~ν~)𝕊N\mathbf{x}+q(\tilde{\nu}^{\prime}-\tilde{\nu})\not\in\mathbb{S}_{N} because this means that x+νν𝕊Mx+\mathbf{\nu}^{\prime}-\mathbf{\nu}\in\mathbb{S}_{M} for some M>NM>N. This implies if x+νν𝕊Nx+\mathbf{\nu}^{\prime}-\mathbf{\nu}\not\in\mathbb{S}_{N}

0=λν~ν~N(𝐱)(V(x+νν)V(x))λνν(𝐱)(V(x+νν)V(x)).\displaystyle\begin{split}0&=\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})(V(x+\mathbf{\nu}^{\prime}-\mathbf{\nu})-V(x))\\ &\leq\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(x+\mathbf{\nu}^{\prime}-\mathbf{\nu})-V(x)).\end{split} (37)

In case 𝐱+q(ν~ν~)𝕊N𝕊\mathbf{x}+q(\tilde{\nu}^{\prime}-\tilde{\nu})\in\mathbb{S}_{N}\subseteq\mathbb{S}, we have that λν~ν~N(𝐱)=λνν(𝐱)\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})=\lambda_{\nu\to\nu^{\prime}}(\mathbf{x}) by the definition (13). Hence by condition 3 and (37) we have for any 𝐱\mathbf{x}

ν~ν~λν~ν~N(𝐱)(V(𝐱+νν)V(𝐱))ννλνν(𝐱)(V(𝐱+νν)V(𝐱))CV(x)+D.\displaystyle\begin{split}&\sum_{\tilde{\nu}\to\tilde{\nu}^{\prime}\in\mathcal{R}}\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})(V(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-V(\mathbf{x}))\\ &\leq\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\mathbf{\nu}^{\prime}-\mathbf{\nu})-V(\mathbf{x}))\\ &\leq-CV(x)+D.\end{split} (38)

Thus VV is also a Lyapunov function for XNX_{N}. Hence Theorem 6.1 (the Foster-Lyapunov criterion for exponential ergodicity) in Meyn and Tweedie [31] implies that for each NN there exist β>0\beta>0 and η>0\eta>0, which are only dependent of CC and DD, such that

pN(,t)πN1βV(𝐱0)eηt.\displaystyle\|p_{N}(\cdot,t)-\pi_{N}\|_{1}\leq\beta V(\mathbf{x}_{0})e^{-\eta t}.

This guarantees that the condition in Theorem 5.2 holds with h(t)=βV(𝐱0)eηth(t)=\beta V(\mathbf{x}_{0})e^{-\eta t}. Hence we have limNππN1=0\displaystyle\lim\limits_{N\to\infty}\|\pi-\pi_{N}\|_{1}=0 by Theorem 5.2.

Now to show that the first passage time τ\tau has the finite mean, we apply the Foster-Lyapunov criterion to X¯\bar{X}. Since (LABEL:eq:Lyapunov_condition_of_bar_x) meets the conditions of Theorem 6.1 in Meyn and Tweedie [31], the probability of X¯\bar{X} converges in time to its stationary distribution exponentially fast. That is, for any subset UU, there exist β¯>0\bar{\beta}>0 and η¯>0\bar{\eta}>0 such that

P(X¯(t)U)π¯(U)βV(𝐱0)eηt,\displaystyle\|P(\bar{X}(t)\in U)-\bar{\pi}(U)\|\leq\beta V(\mathbf{x}_{0})e^{-\eta t},

where π¯\bar{\pi} is the stationary distribution of X¯\overline{X}. This, in turn, implies that

E(τK)=0P(τ>t)𝑑t=0P(X¯(t)Kc)𝑑t0|P(X¯(t)Kc)π¯(Kc)|+π¯(Kc)dt0βV(𝐱0)eηt𝑑t<,\displaystyle\begin{split}E(\tau_{K})&=\int_{0}^{\infty}P(\tau>t)dt\\ &=\int_{0}^{\infty}P(\bar{X}(t)\in K^{c})dt\\ &\leq\int_{0}^{\infty}|P(\bar{X}(t)\in K^{c})-\bar{\pi}(K^{c})|+\bar{\pi}(K^{c})dt\\ &\leq\int_{0}^{\infty}\beta V(\mathbf{x}_{0})e^{-\eta t}dt<\infty,\end{split} (39)

where the second inequality follows as π¯(KC)=0\bar{\pi}(K^{C})=0, which is because X¯\bar{X} is eventually absorbed in KK as the original process XX is irreducible and closed.

Finally we show that for any NN, there exists g(t)g(t) such that P(τN<t)<g(t)P(\tau_{N}<t)<g(t) and 0g(t)𝑑t<\int_{0}^{\infty}g(t)dt<\infty for the first passage time τK,N\tau_{K,N}. By the same reasoning we used to derive (LABEL:eq:Lyapunov_condition_of_bar_x), we also derive by using (LABEL:eq:relation2_V_and_VN) that

ν~ν~λ¯ν~ν~N(𝐱,y)(V¯(𝐱+νν)V¯(𝐱))CV¯(𝐱)+D.\displaystyle\begin{split}&\sum_{\tilde{\nu}\to\tilde{\nu}^{\prime}\in\mathcal{R}}\bar{\lambda}^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x},y)(\overline{V}(\mathbf{x}+\nu^{\prime}-\nu)-\overline{V}(\mathbf{x}))\\ &\leq-C^{\prime}\overline{V}(\mathbf{x})+D^{\prime}.\end{split} (40)

Hence in the same way as used for (39), we have the exponential ergodicty of X¯N\overline{X}_{N} by Theorem 6.1 in Meyn and Tweedie [31], and then we derive that

P(τN>t)\displaystyle P(\tau_{N}>t) =P(X¯N(t)Kc)\displaystyle=P(\overline{X}_{N}(t)\in K^{c})
|P(X¯(t)Kc)π¯N(Kc)|+π¯N(Kc)\displaystyle\leq|P(\overline{X}(t)\in K^{c})-\bar{\pi}_{N}(K^{c})|+\bar{\pi}_{N}(K^{c})
βV(𝐱0)eηt,\displaystyle\leq\beta V(\mathbf{x}_{0})e^{-\eta t},

where π¯N\bar{\pi}_{N} is the stationary distribution of X¯N\overline{X}_{N} which also has zero probability in KCK^{C} as X¯N\overline{X}_{N} is eventually absorbed in KK. Since β\beta and η\eta only depend on CC^{\prime} and DD^{\prime}, we let g(t)=βV(𝐱0)eηtg(t)=\beta V(\mathbf{x}_{0})e^{-\eta t} that satisfies that P(τK,N>t)g(t)P(\tau_{K,N}>t)\leq g(t) for any NN and 0g(t)𝑑t<\int_{0}^{\infty}g(t)dt<\infty. \square

Appendix C Proofs of Accessibility Theorems

In this section, we prove Theorem 6.1. We begin with a necessary lemma. In the following lemma, for a given reaction system (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda), we generate a slack system (S~,𝒞~,~,ΛN)(\widetilde{S},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) admitting a single slack reactant YY. We denote by 𝐰\mathbf{w} the vector for which the slack system admits a the conservation bound 𝐰xN\mathbf{w}\cdot x\leq N for each state xx of the slack system. We also let cc be the maximum stochiometry coefficient of the slack reactant YY in the slack network. Finally note that 𝒞~\widetilde{\mathcal{C}} and 𝒞\mathcal{C} has the one-to-one correspondence as every complex ν~𝒞~\tilde{\nu}\in\widetilde{\mathcal{C}} is obtained by adding the slack reactant YY to a complex ν𝒞\nu\in\mathcal{C} with the stochiometric coefficient uνu_{\nu}. That is, ν~=(ν,u)\tilde{\nu}=(\nu,u) for some ν𝒞\nu\in\mathcal{C} and uνu_{\nu}. Hence as the proof of Theorem 5.4, we let qq be a one-to-one projection function such that q(ν~)=νq(\tilde{\nu})=\nu where ν~=(ν,u)\tilde{\nu}=(\nu,u) for some ν𝒞\nu\in\mathcal{C} and uνu_{\nu}.

Lemma C.1.

For a reaction system (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda), let (S~,𝒞~,~,ΛN)(\widetilde{S},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}) be a slack system with a single slack reactant YY and a conservation vector 𝐰\mathbf{w}. Let cc be the maximum stochiometric coefficient of the slack reactant YY in (S~,𝒞~,~,ΛN)(\widetilde{S},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}). Then if cN(𝐰𝐱)c\leq N-(\mathbf{w}\cdot\mathbf{x}) for a state 𝐱\mathbf{x}, then

λν~ν~N(x)>0if only ifλνν(x)>0,\displaystyle\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(x)>0\quad\text{if only if}\quad\lambda_{\nu\to\nu^{\prime}}(x)>0,

where ν~ν~~\tilde{\nu}\to\tilde{\nu}^{\prime}\in\widetilde{\mathcal{R}} and νν\nu\to\nu^{\prime}\in\mathcal{R} are reactions such that q(ν~)=νq(\tilde{\nu})=\nu and q(ν~)=νq(\tilde{\nu})=\nu^{\prime}.

Proof.

Let ν~𝒞~\tilde{\nu}\in\widetilde{\mathcal{C}} such that ν~=(ν,uν)\tilde{\nu}=(\nu,u_{\nu}) with the stoichiometric coefficient uνu_{\nu} of YY. By the definition of λν~ν~N\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}} shown in (13),

λν~ν~N(x)>0if only ifλνν(x)>0,\displaystyle\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(x)>0\quad\text{if only if}\quad\lambda_{\nu\to\nu^{\prime}}(x)>0,

so long as y=N𝐰𝐱uνy=N-\mathbf{w}\cdot\mathbf{x}\geq u_{\nu}. Since c=maxν𝒞{uν}c=\max_{\nu\in\mathcal{C}}\{u_{\nu}\}, the result follows. ∎

Proof of Theorem 6.1: We denote by 𝕊\mathbb{S} the irreducible and closed state space of XX such that X(0)=𝐱0X(0)=\mathbf{x}_{0}. We also denote by 𝕊N\mathbb{S}_{N} the communication class of XNX_{N} such that x0𝕊Nx_{0}\in\mathbb{S}_{N} for each NN. Then 𝕊N𝕊\mathbb{S}_{N}\subseteq\mathbb{S}, and the irreducibility of 𝕊\mathbb{S} guarantees that i=1𝕊i=𝕊\cup_{i=1}^{\infty}\mathbb{S}_{i}=\mathbb{S}. Therefore there exists N1N_{1} such that 𝕊N,𝐱0K\mathbb{S}_{N,\mathbf{x}_{0}}\cap K\neq\emptyset if NN1N\geq N_{1}. This implies that it is enough to show that 𝕊N,𝐱0\mathbb{S}_{N,\mathbf{x}_{0}} is closed for NN large enough because 𝕊N,𝐱0\mathbb{S}_{N,\mathbf{x}_{0}} is a finite subset. To prove this, we claim that there exists N0N_{0} such that if NN0N\geq N_{0}, then 𝕊N\mathbb{S}_{N} admits no out-going flow.

To prove the claim by contradiction, we assume that for a sequence NkN_{k} such that limkNk=\lim_{k\to\infty}N_{k}=\infty, there is an out-going flow from 𝕊Nk\mathbb{S}_{N_{k}} to a state 𝐱k𝕊Nkc\mathbf{x}_{k}\in\mathbb{S}_{N_{k}}^{c}. We find a sequence of reactions with which XNkX_{N_{k}} returns to 𝕊Nk\mathbb{S}_{N_{k}} from 𝐱k\mathbf{x}_{k} for all large enough kk.

First of all, note {XiY}~\{X_{i}\to Y\}\subset\widetilde{\mathcal{R}} because each reaction XiX_{i}\to\emptyset\in\mathcal{R} is converted to XiYX_{i}\to Y in any optimized slack network by its definition of an optimized slack network (See Section 3.3). Hence for each kk, by firing the reactions {XiY}\{X_{i}\to Y\}, XNkX_{N_{k}} can reach to 𝟎=(0,0,,0)\mathbf{0}=(0,0,\dots,0) from 𝐱k\mathbf{x}_{k}.

The next aim is to show that for any fixed kk large enough, there exist a sequence of reactions {ν~1ν~1,ν~nν~n}~\{\tilde{\nu}_{1}\to\tilde{\nu}^{\prime}_{1},\dots\tilde{\nu}_{n}\to\tilde{\nu}^{\prime}_{n}\}\subset\widetilde{\mathcal{R}} such that we have i) λν~iν~iNk(𝐱k+j=1i1(ν~jν~j))>0\lambda^{N_{k}}_{\tilde{\nu}_{i}\to\tilde{\nu}^{\prime}_{i}}(\mathbf{x}_{k}+\sum_{j=1}^{i-1}(\tilde{\nu}^{\prime}_{j}-\tilde{\nu}_{j}))>0 for each ii, and ii) 𝐱k+j=1n(q(ν~j)q(ν~j))𝕊Nk\mathbf{x}_{k}+\sum_{j=1}^{n}(q(\tilde{\nu}^{\prime}_{j})-q(\tilde{\nu}_{j}))\in\mathbb{S}_{N_{k}}. This means that XNkX_{N_{k}} can reach to 𝕊Nk\mathbb{S}_{N_{k}} from 𝐱k\mathbf{x}_{k} along the reactions ν~iν~i\tilde{\nu}_{i}\to\tilde{\nu}_{i} in that order. For this aim, we make a use of the irreducibility of 𝕊\mathbb{S} since 𝕊\mathbb{S} so that there exists a sequence of reactions {ν1ν1,,νnνn}\{\nu_{1}\to\nu^{\prime}_{1},\cdots,\nu_{n}\to\nu^{\prime}_{n}\}\subset\mathcal{R} such that i) λνiνi(𝟎+j=1i1(νjνj))>0\lambda_{\nu_{i}\to\nu^{\prime}_{i}}(\mathbf{0}+\sum_{j=1}^{i-1}(\nu^{\prime}_{j}-\nu_{j}))>0 for each ii and ii) 𝟎+j=1n(νjνj)𝕊N1\mathbf{0}+\sum_{j=1}^{n}(\nu^{\prime}_{j}-\nu_{j})\in\mathbb{S}_{N_{1}}. By making N0N1N_{0}\geq N_{1} large enough, we have if NkN0N_{k}\geq N_{0}, then

𝐰(𝟎+j=1i(νjνj))<Ncfor each i.\displaystyle\mathbf{w}\cdot\left(\mathbf{0}+\sum_{j=1}^{i}(\nu^{\prime}_{j}-\nu_{j})\right)<N-c\quad\text{for each $i$.}

Hence by Lemma C.1, for each reaction ν~iν~i\tilde{\nu}_{i}\to\tilde{\nu}^{\prime}_{i} such that q(ν~i)=νiq(\tilde{\nu}_{i})=\nu_{i} and q(ν~i)=νiq(\tilde{\nu}^{\prime}_{i})=\nu^{\prime}_{i} we have

λν~iν~iN(𝟎+j=1i1(q(ν~j)q(ν~j)))>0for each i.\displaystyle\lambda^{N}_{\tilde{\nu}_{i}\to\tilde{\nu}^{\prime}_{i}}\left(\mathbf{0}+\sum_{j=1}^{i-1}(q(\tilde{\nu}^{\prime}_{j})-q(\tilde{\nu}_{j}))\right)>0\quad\text{for each $i$}. (41)

Finally since q(ν~i)=νiq(\tilde{\nu}_{i})=\nu_{i} and q(ν~i)=νiq(\tilde{\nu}^{\prime}_{i})=\nu^{\prime}_{i} for each ii, we have

𝟎+j=1i1(q(ν~j)q(ν~j))𝕊N0𝕊N.\displaystyle\mathbf{0}+\sum_{j=1}^{i-1}(q(\tilde{\nu}^{\prime}_{j})-q(\tilde{\nu}_{j}))\in\mathbb{S}_{N_{0}}\subseteq\mathbb{S}_{N}. (42)

Hence (41) and (42) imply that if NkN0N_{k}\geq N_{0}, then XNkX_{N_{k}} can re-enter to 𝕊Nk\mathbb{S}_{N_{k}} from 𝟎\mathbf{0} with positive probability.

Consequently, we constructed a sequence of reactions along which XNkX_{N_{k}} can re-enter 𝕊Nk\mathbb{S}_{N_{k}} from 𝐱k\mathbf{x}_{k} for any kk such that NkN0N_{k}\geq N_{0}. Since each 𝕊Nk\mathbb{S}_{N_{k}} is a communication class, 𝐱k𝕊Nk\mathbf{x}_{k}\in\mathbb{S}_{N_{k}} and ,in turn, this contradictions to the assumption that 𝐱k𝕊Nk\mathbf{x}_{k}\in\mathbb{S}_{N_{k}}. Hence the claim holds so that 𝕊N\mathbb{S}_{N} is closed for any NN large enough. \square

Appendix D Proofs for Lemma 6.1 and Theorem 6.2

Proof of Lemma 6.1: In the deterministic model 𝐱~(t)=(x~1(t),,x~d(t),y1(t),,ym(t))\tilde{\mathbf{x}}(t)=(\tilde{x}_{1}(t),\dots,\tilde{x}_{d}(t),y_{1}(t),\dots,y_{m}(t))^{\top} associated with (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}), if we fix yi(t)1y_{i}(t)\equiv 1 for all i=1,2,,mi=1,2,\dots,m, then x~(t)\tilde{x}(t) follows the same ODE system as the deterministic model x(t)x(t) associated with the original network (𝒮,𝒞,,Λ)(\mathcal{S},\mathcal{C},\mathcal{R},\Lambda) does. Therefore c~=(c,1,1,,1)\tilde{c}=(c^{*},1,1,\dots,1)^{\top} is a complex balanced steady state for x~(t)\tilde{x}(t). It has been shown that the existence of a single positive complex balanced steady state implies that all other positive steady states are complex balanced [12], therefore completing the proof.

\square

Proof of Theorem 6.2: Let 𝐰i\mathbf{w}_{i} and NiN_{i} be the conservative vector and the conservative quantity of the slack network (𝒮~,𝒞~,~,ΛN)(\widetilde{\mathcal{S}},\widetilde{\mathcal{C}},\widetilde{\mathcal{R}},\Lambda^{N}), respectively. Then π\pi is a stationary solution of the chemical master equation (3) of XX,

ννλνν(𝐱ν+ν)π(𝐱ν+ν)\displaystyle\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x}-\nu^{\prime}+\nu)\pi(\mathbf{x}-\nu^{\prime}+\nu)
=ννλνν(𝐱)π(𝐱).\displaystyle=\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})\pi(\mathbf{x}).

Especially, as shown in Theorem 6.4, Anderson et al[3], π\pi satisfies the stochastic complex balance for XX: for each complex ν\nu^{*}\in\mathcal{R} and a state 𝐱\mathbf{x},

ννν=νλνν(𝐱ν+ν)π(𝐱ν+ν)=ννν=νλνν(𝐱)π(𝐱).\displaystyle\begin{split}&\sum_{\begin{subarray}{c}\nu\to\nu^{\prime}\in\mathcal{R}\\ \nu^{\prime}=\nu^{*}\end{subarray}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x}-\nu^{\prime}+\nu)\pi(\mathbf{x}-\nu^{\prime}+\nu)\\ &=\sum_{\begin{subarray}{c}\nu\to\nu^{\prime}\in\mathcal{R}\\ \nu=\nu^{*}\end{subarray}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})\pi(\mathbf{x}).\end{split} (43)

Let qq be the one-to-one projection function qq that we defined for the proof of Theorem 6.1. Note that each reaction ν~ν~~\tilde{\nu}\to\tilde{\nu}^{\prime}\in\widetilde{\mathcal{R}} is defined as it satisfies the conservative law

𝐰i(νν)+ν~d+iνd+i=0,\displaystyle\mathbf{w}_{i}\cdot(\nu^{\prime}-\nu)+\tilde{\nu}^{\prime}_{d+i}-\nu_{d+i}=0, (44)

where q(ν~)=νq(\tilde{\nu})=\nu and q(ν~)=νq(\tilde{\nu}^{\prime})=\nu^{\prime}. Then π\pi also satisfies the stochastic complex balance for XNX_{N} because (LABEL:eq:sto_cx_bal) and (44) imply that for each complex ν~~\tilde{\nu}^{*}\in\widetilde{\mathcal{R}} and a state 𝐱\mathbf{x},

ν~ν~;~ν~=ν~λν~ν~N(𝐱q(ν)+q(ν))π(𝐱q(ν)+q(ν))\displaystyle\sum_{\begin{subarray}{c}\tilde{\nu}\to\tilde{\nu}^{\prime};\in\widetilde{\mathcal{R}}\\ \tilde{\nu}^{\prime}=\tilde{\nu}^{*}\end{subarray}}\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x}-q(\nu^{\prime})+q(\nu))\pi(\mathbf{x}-q(\nu^{\prime})+q(\nu))
=ννν~=ν~λν~ν~(𝐱ν+ν)π(𝐱ν+ν)i=1r𝟙{Ni𝐰i(𝐱ν+ν)ν~d+i}\displaystyle=\sum_{\begin{subarray}{c}\nu\to\nu^{\prime}\in\mathcal{R}\\ \tilde{\nu}^{\prime}=\tilde{\nu}^{*}\end{subarray}}\lambda_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x}-\nu^{\prime}+\nu)\pi(\mathbf{x}-\nu^{\prime}+\nu)\prod_{i=1}^{r}\mathbbm{1}_{\{N_{i}-\mathbf{w}_{i}\cdot(\mathbf{x}-\nu^{\prime}+\nu)\geq\tilde{\nu}_{d+i}\}}
=ννν~=ν~λν~ν~(𝐱ν+ν)π(𝐱ν+ν)i=1r𝟙{Ni𝐰i𝐱ν~d+i}\displaystyle=\sum_{\begin{subarray}{c}\nu\to\nu^{\prime}\in\mathcal{R}\\ \tilde{\nu}^{\prime}=\tilde{\nu}^{*}\end{subarray}}\lambda_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x}-\nu^{\prime}+\nu)\pi(\mathbf{x}-\nu^{\prime}+\nu)\prod_{i=1}^{r}\mathbbm{1}_{\{N_{i}-\mathbf{w}_{i}\cdot\mathbf{x}\geq\tilde{\nu}^{*}_{d+i}\}}
=ννν=νλνν(𝐱)π(𝐱)𝟙{Ni𝐰i𝐱ν~d+i}\displaystyle=\sum_{\begin{subarray}{c}\nu\to\nu^{\prime}\in\mathcal{R}\\ \nu=\nu^{*}\end{subarray}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})\pi(\mathbf{x})\mathbbm{1}_{\{N_{i}-\mathbf{w}_{i}\cdot\mathbf{x}\geq\tilde{\nu}^{*}_{d+i}\}}
=ν~ν~~ν~=ν~λν~ν~N(𝐱)π(𝐱).\displaystyle=\sum_{\begin{subarray}{c}\tilde{\nu}\to\tilde{\nu}^{\prime}\in\widetilde{\mathcal{R}}\\ \tilde{\nu}=\tilde{\nu}^{*}\end{subarray}}\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})\pi(\mathbf{x}).

Then by summing up for each ν~~\tilde{\nu}^{*}\in\widetilde{\mathcal{R}},

ν~ν~;~λν~ν~N(𝐱q(ν)+q(ν))π(𝐱q(ν)+q(ν))\displaystyle\sum_{\tilde{\nu}\to\tilde{\nu}^{\prime};\in\widetilde{\mathcal{R}}}\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x}-q(\nu^{\prime})+q(\nu))\pi(\mathbf{x}-q(\nu^{\prime})+q(\nu))
=ν~ν~~λν~ν~N(𝐱)π(𝐱),\displaystyle=\sum_{\tilde{\nu}\to\tilde{\nu}^{\prime}\in\widetilde{\mathcal{R}}}\lambda^{N}_{\tilde{\nu}\to\tilde{\nu}^{\prime}}(\mathbf{x})\pi(\mathbf{x}),

and, in turn, πN\pi_{N} is a stationary solution of the chemical master equation of XNX_{N}. Since the state space of XX and XNX_{N} differ, MNM_{N} is a constant such that the sum of MNπ(𝐱)M_{N}\pi(\mathbf{x}) over the state space of XNX_{N} is one. Then πN=MNπ\pi_{N}=M_{N}\pi. \hfill\square

Appendix E Lypunov Functions for Example Systems

E.1 Lotka-Volterra with Migration

We construct a Lyapunov function satisfying the conditions of Theorem 5.4 for the Lotka-Volterra model with migration (Figure 4A). First, for both the original model and the slack network, we assume A(0)=a0A(0)=a_{0}, B(0)=b0B(0)=b_{0} that are not depend on NN.

Let V(𝐱)=ew𝐱V(\mathbf{x})=e^{w\cdot\mathbf{x}} with w=(1,1)w=(1,1)^{\top}. The condition in Theorem 5.4 obviously holds. Note that the slack network of Figure 4A admits a conservation law A+B+Y=NA+B+Y=N and the state space of the system 𝕊N\mathbb{S}_{N} is {(a,b)|a+bN}\{(a,b)\ |\ a+b\leq N\}. Then condition 2 in Theorem 5.4 also follows by the definition of VV.

Now, to show condition 3 in Theorem 5.4, we first note that for 𝐱=(a,b)\mathbf{x}=(a,b),

ννλνν(𝐱)(V(𝐱+νν)V(𝐱))=ew𝐱ννλνν(𝐱)(ew(νν)1)=V(𝐱)(κ1b(e11)+κ2(e1)+κ3(e11)+κ4a(e11)+κ5a(e1)).\displaystyle\begin{split}&\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x}))\\ &=e^{w\cdot\mathbf{x}}\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(e^{w\cdot(\nu^{\prime}-\nu)}-1)\\ &=V(\mathbf{x})\Big{(}\kappa_{1}b(e^{-1}-1)+\kappa_{2}(e-1)+\kappa_{3}(e^{-1}-1)\\ &\ \ +\kappa_{4}a(e^{-1}-1)+\kappa_{5}a(e-1)\Big{)}.\end{split} (45)

Since we assumed κ4κ5=3\frac{\kappa_{4}}{\kappa_{5}}=3 (see caption of Figure 4), each coefficient of aa and bb is negative. Thus, there exists M>0M>0 such that for each 𝐱{(a,b)|a>M or bM}\mathbf{x}\in\{(a,b)\ |\ a>M\text{ or }b\geq M\}, the left-hand side in (LABEL:eq:LV_detail1) satisfies

ννλνν(𝐱)(V(𝐱+νν)V(𝐱))\displaystyle\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x}))
CV(𝐱)\displaystyle\leq-CV(\mathbf{x})

for some C>0C>0. Letting

D=(C+1)maxaM and bMννλνν(𝐱)(V(𝐱+νν)V(𝐱)),D=(C+1)\displaystyle\max_{a\leq M\text{ and }b\leq M}\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x})),

condition 3 in Theorem 5.4 holds with CC and DD.

E.2 Protein Synthesis with Slow Toggle Switch

We also make a use of the Lyapunov approach shown in Theorem 5.4 to prove that the first passage time of the slack system in Figure 5B converges to the original first passage time, as the truncation size NN goes infinity. Let XN=(X,Z,D0X,D1X,D0Z,D1Z)X_{N}=(X,Z,D^{X}_{0},D^{X}_{1},D^{Z}_{0},D^{Z}_{1}) be the stochastic system associated with the slack system. Recall that the slack network admits a conservation relation wXNw\cdot X\leq N with w=(1,1,0,0,0,0)w=(1,1,0,0,0,0). Hence we define a Lyapunov function V(𝐱)=ex+yV(\mathbf{x})=e^{x+y} where 𝐱=(x,z,dx0,dx1,dz0,dz1)\mathbf{x}=(x,z,dx_{0},dx_{1},dz_{0},dz_{1}). By the definition of VV, it is obvious that conditions 1 and 2 in Theorem 5.4 hold. So we show that condition 3 in Theorem holds.

For a reaction ννI:={Z+D0XD1X,X0,X+D0ZD1Z,Z0}\nu\to\nu^{\prime}\in\mathcal{R}_{I}:=\{Z+D^{X}_{0}\to D^{X}_{1},X\to 0,X+D^{Z}_{0}\to D^{Z}_{1},Z\to 0\}, it is clear that the term V(𝐱+νν)V(𝐱)V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x}) is negative. For each reaction νν\nu\to\nu^{\prime} in Ic\mathcal{R}^{c}_{I}, it is also clear that the term V(𝐱+νν)V(𝐱)V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x}) is positive. However, the reaction intensity for a reaction in I\mathcal{R}_{I} is linear in either xx and zz, while the reaction intensity for a reaction in Ic\mathcal{R}^{c}_{I} is constant. Therefore there exists a constant M>0M>0 such that if x>Mx>M or z>Mz>M, then

ννλνν(𝐱)(V(𝐱+νν)V(𝐱))\displaystyle\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x}))
CV(𝐱)\displaystyle\leq-CV(\mathbf{x})

for some C>0C>0. Hence, letting

D=(C+1)maxxM and zMννλνν(𝐱)(V(𝐱+νν)V(𝐱)),D=(C+1)\displaystyle\max_{x\leq M\text{ and }z\leq M}\sum_{\nu\to\nu^{\prime}\in\mathcal{R}}\lambda_{\nu\to\nu^{\prime}}(\mathbf{x})(V(\mathbf{x}+\nu^{\prime}-\nu)-V(\mathbf{x})),

condition 3 in Theorem 5.4 holds with CC and DD.

References

  • [1] M Ali Al-Radhawi, Domitilla Del Vecchio, and Eduardo D Sontag. Multi-modality in gene regulatory networks with slow promoter kinetics. PLoS computational biology, 15(2):e1006784, 2019.
  • [2] David F Anderson and Daniele Cappelletti. Discrepancies between extinction events and boundary equilibria in reaction networks. Journal of mathematical biology, 79(4):1253–1277, 2019.
  • [3] David F. Anderson, Gheorghe Craciun, and Thomas G. Kurtz. Product-form stationary distributions for deficiency zero chemical reaction networks. Bulletin of Mathematical Biology, 72(8):1947–1970, Nov 2010.
  • [4] David F Anderson and Jinsu Kim. Some network conditions for positive recurrence of stochastically modeled reaction networks. SIAM Journal on Applied Mathematics, 78(5):2692–2713, 2018.
  • [5] Elias August and Mauricio Barahona. Solutions of weakly reversible chemical reaction networks are bounded and persistent. IFAC Proceedings Volumes, 43(6):42–47, 2010.
  • [6] Corentin Briat, Ankit Gupta, and Mustafa Khammash. Antithetic integral feedback ensures robust perfect adaptation in noisy biomolecular networks. Cell systems, 2(1):15–26, 2016.
  • [7] Youfang Cao and Jie Liang. Optimal enumeration of state space of finitely buffered stochastic molecular networks and exact computation of steady state landscape probability. BMC Systems Biology, 2(1):30, 2008.
  • [8] Youfang Cao, Anna Terebus, and Jie Liang. Accurate chemical master equation solution using multi-finite buffers. Multiscale Modeling & Simulation, 14(2):923–963, 2016.
  • [9] Tom Chou and Maria R D’Orsogna. First passage problems in biology. In First-passage phenomena and their applications, pages 306–345. World Scientific, 2014.
  • [10] German Enciso and Jinsu Kim. Constant order multiscale reduction for stochastic reaction networks. arXiv preprint arXiv:1909.11916, 2019.
  • [11] German Enciso and Jinsu Kim. Embracing noise in chemical reaction networks. Bulletin of mathematical biology, 81(5):1261–1267, 2019.
  • [12] Martin Feinberg. Complex balancing in general kinetic systems. Archive for rational mechanics and analysis, 49(3):187–194, 1972.
  • [13] Martin Feinberg. Chemical reaction network structure and the stability of complex isothermal reactors—i. the deficiency zero and deficiency one theorems. Chemical engineering science, 42(10):2229–2268, 1987.
  • [14] Daniel T Gillespie. A rigorous derivation of the chemical master equation. Physica A: Statistical Mechanics and its Applications, 188(1-3):404–425, 1992.
  • [15] Daniel T Gillespie. Stochastic simulation of chemical kinetics. Annu. Rev. Phys. Chem., 58:35–55, 2007.
  • [16] Ankit Gupta, Jan Mikelson, and Mustafa Khammash. A finite state projection algorithm for the stationary solution of the chemical master equation. The Journal of chemical physics, 147(15):154101, 2017.
  • [17] Mads Christian Hansen, Wiuf Carsten, et al. Existence of a unique quasi-stationary distribution in stochastic reaction networks. Electronic Journal of Probability, 25, 2020.
  • [18] Andrew G Hart and Richard L Tweedie. Convergence of invariant measures of truncation approximations to markov processes. 2012.
  • [19] João Pedro Hespanha. StochDynTools — a MATLAB toolbox to compute moment dynamics for stochastic networks of bio-chemical reactions. Available at http://www.ece.ucsb.edu/~hespanha, May 2007.
  • [20] Fritz Horn. Necessary and sufficient conditions for complex balancing in chemical kinetics. Archive for Rational Mechanics and Analysis, 49(3):172–186, 1972.
  • [21] Jeffrey J Hunter. Generalized inverses and their application to applied probability problems. Technical report, VIRGINIA POLYTECHNIC INST AND STATE UNIV BLACKSBURG DEPT OF INDUSTRIAL …, 1980.
  • [22] Jeffrey J Hunter. The computation of the mean first passage times for markov chains. Linear Algebra and its Applications, 549:100–122, 2018.
  • [23] Atefeh Kazeroonian, Fabian Fröhlich, Andreas Raue, Fabian J Theis, and Jan Hasenauer. Cerena: Chemical reaction network analyzer—a toolbox for the simulation and analysis of stochastic chemical kinetics. PloS one, 11(1), 2016.
  • [24] Jae Kyoung Kim and Eduardo Sontag. FEEDME — a MATLAB codes to calculate stationary moments of feed forward network and complex balanced network. Available at http://www.ece.ucsb.edu/~hespanha, May 2007.
  • [25] Juan Kuntz, Philipp Thomas, Guy-Bart Stan, and Mauricio Barahona. Approximations of countably-infinite linear programs over bounded measure spaces. arXiv preprint arXiv:1810.03658, 2018.
  • [26] Juan Kuntz, Philipp Thomas, Guy-Bart Stan, and Mauricio Barahona. Bounding the stationary distributions of the chemical master equation via mathematical programming. The Journal of chemical physics, 151(3):034109, 2019.
  • [27] Juan Kuntz, Philipp Thomas, Guy-Bart Stan, and Mauricio Barahona. Stationary distributions of continuous-time markov chains: a review of theory and truncation-based approximations. arXiv preprint arXiv:1909.05794, 2019.
  • [28] Fernando López-Caamal and Tatiana T Marquez-Lago. Exact probability distributions of selected species in stochastic chemical reaction networks. Bulletin of mathematical biology, 76(9):2334–2361, 2014.
  • [29] Shev MacNamara, Kevin Burrage, and Roger B Sidje. Multiscale modeling of chemical kinetics via the master equation. Multiscale Modeling & Simulation, 6(4):1146–1168, 2008.
  • [30] Leonor Menten and MI Michaelis. Die kinetik der invertinwirkung. Biochem Z, 49:333–369, 1913.
  • [31] Sean P Meyn and Richard L Tweedie. Stability of markovian processes iii: Foster–lyapunov criteria for continuous-time processes. Advances in Applied Probability, 25(3):518–548, 1993.
  • [32] Brian Munsky and Mustafa Khammash. The finite state projection algorithm for the solution of the chemical master equation. The Journal of chemical physics, 124(4):044104, 2006.
  • [33] Loïc Paulevé, Gheorghe Craciun, and Heinz Koeppl. Dynamical properties of discrete reaction networks. Journal of mathematical biology, 69(1):55–72, 2014.
  • [34] Johan Paulsson. Summing up the noise in gene networks. Nature, 427(6973):415–418, 2004.
  • [35] E Seneta. Finite approximations to infinite non-negative matrices. In Mathematical Proceedings of the Cambridge Philosophical Society, volume 63, pages 983–992. Cambridge University Press, 1967.
  • [36] Richard L Tweedie. Truncation procedures for non-negative matrices. Journal of Applied Probability, 8(2):311–320, 1971.
  • [37] Romain Yvinec, Maria R d’Orsogna, and Tom Chou. First passage times in homogeneous nucleation and self-assembly. The Journal of chemical physics, 137(24):244107, 2012.