Processing math: 100%
Research article Special Issues

Fast reactions with non-interacting species in stochastic reaction networks

  • We consider stochastic reaction networks modeled by continuous-time Markov chains. Such reaction networks often contain many reactions, potentially occurring at different time scales, and have unknown parameters (kinetic rates, total amounts). This makes their analysis complex. We examine stochastic reaction networks with non-interacting species that often appear in examples of interest (e.g. in the two-substrate Michaelis Menten mechanism). Non-interacting species typically appear as intermediate (or transient) chemical complexes that are depleted at a fast rate. We embed the Markov process of the reaction network into a one-parameter family under a two time-scale approach, such that molecules of non-interacting species are degraded fast. We derive simplified reaction networks where the non-interacting species are eliminated and that approximate the scaled Markov process in the limit as the parameter becomes small. Then, we derive sufficient conditions for such reductions based on the reaction network structure for both homogeneous and time-varying stochastic settings, and study examples and properties of the reduction.

    Citation: Linard Hoessly, Carsten Wiuf. Fast reactions with non-interacting species in stochastic reaction networks[J]. Mathematical Biosciences and Engineering, 2022, 19(3): 2720-2749. doi: 10.3934/mbe.2022124

    Related Papers:

    [1] Daniele Cappelletti, Badal Joshi . Transition graph decomposition for complex balanced reaction networks with non-mass-action kinetics. Mathematical Biosciences and Engineering, 2022, 19(8): 7649-7668. doi: 10.3934/mbe.2022359
    [2] Qiqi Deng, Aimin Chen, Huahai Qiu, Tianshou Zhou . Analysis of a non-Markov transcription model with nuclear RNA export and RNA nuclear retention. Mathematical Biosciences and Engineering, 2022, 19(8): 8426-8451. doi: 10.3934/mbe.2022392
    [3] Zhenquan Zhang, Junhao Liang, Zihao Wang, Jiajun Zhang, Tianshou Zhou . Modeling stochastic gene expression: From Markov to non-Markov models. Mathematical Biosciences and Engineering, 2020, 17(5): 5304-5325. doi: 10.3934/mbe.2020287
    [4] David F. Anderson, Tung D. Nguyen . Results on stochastic reaction networks with non-mass action kinetics. Mathematical Biosciences and Engineering, 2019, 16(4): 2118-2140. doi: 10.3934/mbe.2019103
    [5] Giorgos Minas, David A Rand . Parameter sensitivity analysis for biochemical reaction networks. Mathematical Biosciences and Engineering, 2019, 16(5): 3965-3987. doi: 10.3934/mbe.2019196
    [6] David F. Anderson, Jinsu Kim . Mixing times for two classes of stochastically modeled reaction networks. Mathematical Biosciences and Engineering, 2023, 20(3): 4690-4713. doi: 10.3934/mbe.2023217
    [7] Maya Mincheva, Gheorghe Craciun . Graph-theoretic conditions for zero-eigenvalue Turing instability in general chemical reaction networks. Mathematical Biosciences and Engineering, 2013, 10(4): 1207-1226. doi: 10.3934/mbe.2013.10.1207
    [8] Linda J. S. Allen, Vrushali A. Bokil . Stochastic models for competing species with a shared pathogen. Mathematical Biosciences and Engineering, 2012, 9(3): 461-485. doi: 10.3934/mbe.2012.9.461
    [9] H.Thomas Banks, Shuhua Hu . Nonlinear stochastic Markov processes and modeling uncertainty in populations. Mathematical Biosciences and Engineering, 2012, 9(1): 1-25. doi: 10.3934/mbe.2012.9.1
    [10] ZongWang, Qimin Zhang, Xining Li . Markovian switching for near-optimal control of a stochastic SIV epidemic model. Mathematical Biosciences and Engineering, 2019, 16(3): 1348-1375. doi: 10.3934/mbe.2019066
  • We consider stochastic reaction networks modeled by continuous-time Markov chains. Such reaction networks often contain many reactions, potentially occurring at different time scales, and have unknown parameters (kinetic rates, total amounts). This makes their analysis complex. We examine stochastic reaction networks with non-interacting species that often appear in examples of interest (e.g. in the two-substrate Michaelis Menten mechanism). Non-interacting species typically appear as intermediate (or transient) chemical complexes that are depleted at a fast rate. We embed the Markov process of the reaction network into a one-parameter family under a two time-scale approach, such that molecules of non-interacting species are degraded fast. We derive simplified reaction networks where the non-interacting species are eliminated and that approximate the scaled Markov process in the limit as the parameter becomes small. Then, we derive sufficient conditions for such reductions based on the reaction network structure for both homogeneous and time-varying stochastic settings, and study examples and properties of the reduction.



    Reaction network theory offers a quantitative framework for biochemistry, systems biology, and cellular biology, by enabling the modeling of biological systems. Deterministic models have been the main focus area with contributions going back more than hundred years [1]. However, the increasing interest in living systems at the cellular level motivates the use of stochastic models to describe the variation and noise found in systems with low molecule numbers. Typically, stochastic models are continuous-time Markov chains (CTMCs) on the state space Zn0, where a state represents the vector of molecule numbers of the n species in the system.

    We are here concerned with the transient behavior of such CTMCs, in contrast to the stationary behavior (the existence of stationary distributions). In practice, variations in molecule numbers and reaction rates might yield phenomena that evolve on different time-scales, enabling simplifications [2]. In particular, we are interested in systems with two time-scales, also known as slow-fast systems, where a set of reactions are fast (in a relative sense) compared to the remaining (slow) reactions. The objective is to approximate, in a mathematically rigorous way, the dynamics of the original CTMC, with another CTMC in smaller dimension (with fewer species). This other CTMC should ideally be interpreted as a reaction network, obtained by reduction of the original reaction network. We will give conditions for when this can be done.

    In the deterministic setting, the heuristic quasi-steady state approximation (QSSA) [3] and singular perturbation theory in the sense of Tikhonov-Fenichel [4,5] have been the main means to derive lower dimensional reduced models, where [5] is the classic reference of Tikhonov, while [4] is from Fenichel. (See [6] or [7] and references therein for the relationship between the two approaches and when the QSSA is valid, and [8] for comparison with stochastic QSSA.) A motivation for our work is the situation for which reactions involving so-called non-interacting species [9] in the reactant has reaction rate constants scaled by 1/ϵ, where ϵ>0 is a small number [10]. We will consider a similar situation for stochastic reaction networks and point out differences between the two settings.

    Many stochastic studies consider physical or heuristic-based derivations to extract reduced reaction networks, for example, SDEs or hybrid models, or ad-hoc reductions [11,12,13]. [11] comments on general deterministic approaches, [12] studies heuristics for stochastic single scale reductions, while [13] reviews practical aspects of deterministic and stochastic reductions. One stream of research eliminates species using heuristic projection arguments [14,15], where [14] treats some reversible cases and [15] irreversible examples. Rigorous simplifications and reductions often follow scaling limits of Markov processes in a multi-scale setting [16]. These might be applied to concrete examples, if certain conditions are satisfied and the different scaling parameters are balanced (in a specific sense), as done for specific examples of reaction networks [17], more generally [18], or for spatial reaction networks [19]. Scaling laws for a special class of reaction networks with intermediate species (a special type of non-interacting species) and their explicit reduction are given in [20]. We extends this work to reaction networks with non-interacting species on two time-scales.

    These two time-scales separate the transition intensities of the CTMC into a fast and a slow component such that the Q-matrix has the following form

    Qϵ=1ϵ˜Q+ˆQ.

    The transition intensities of the reactions in the network are divided into two kinds:

    ● Fast reactions with scaled transition intensity λϵyy(x):=1ϵλyy(x).

    ● Slow reactions with unscaled transition intensity λϵyy(x):=λyy(x).

    The fast reactions are determined by non-interacting species. Our work is inspired by previous work on non-interacting species in deterministic systems that focussed on algebraic reductions [9], applications to biological examples [21], the connection to singular perturbation [10], and intermediate species in stochastic models [20]. The technical part is based on singularly perturbed Markov chains [22] and watched Markov chains [23]. In order to derive the limiting dynamics systematically, we associate a graph to the reaction network that captures the dynamics of the fast reactions. This enables the definition of the reduced reaction network and its dynamics. We give limit theorems for the approximation of the original CTMC to the reduced CTMC on compact time intervals. Furthermore, we study the case of time-heterogeneous CTMCs, and show that the same reductions work.

    As an example, consider a mass-action reaction network as follows

    S1κ1U1+S21ϵκ2S3,U11ϵκ3S4,

    with two fast reactions determined by the presence of the non-interacting species U1, that degrades fast. The reduced reaction network is given as

    S1S3,S1S2+S4,

    with transition intensities κ1κ2zS1(zS2+1)κ2(zS2+1)+κ3 for the first reaction and κ1κ3zS1κ2(zS2+1)+κ3 for the second. After creation of a U1 molecule (and a S2 molecule), then it might be degraded either by consumption of the S2 molecule, resulting in the net reaction S1S3, or without consumption of the S2 molecule, resulting in the net reaction S1S2+S4. In both cases, the transition intensities reflect that the presence of S1 is required for the reactions to take place.

    We next outline the content, where in § 2 preliminaries on graph theory and reaction networks are covered. In § 3, we introduce non-interacting species and introduce a graph that is used to define the reduced reaction network by elimination of the non-interacting species. In § 4, we study transient approximations for stochastic reaction networks with non-interacting species via the previously introduced reduction. In § 5, we give realistic examples, study sufficient conditions for reductions and compare stationary properties of the reduction with the original reaction network. Finally in § 6, we discuss the results, approach, and elaborate on the relation to the literature. In the Appendices § A, § B, § C, we give proofs as well as brief introductions to the theory on singularly perturbed CTMCs and watched CTMCs.

    Let Rp be the real p-dimensional space, and Rp>0 (Rp0) the subset of elements of Rp with strictly positive (non-negative) entries in all components. A vector yRp is written as (y1,,yp), where yi is the i-th component. For vectors y1,,ykRp, max(y1,,yk) denotes the component-wise maximum, and y1y2 denotes component-wise inequality. The inner product between y1 and y2 is denoted y1,y2. The cardinality of a set A is denoted |A|.

    G=(V,E) a directed graph consists of the set of vertices V and edge set E. A directed subgraph of G=(V,E) is a directed graph G=(V,E) with VV,EE with E on V. A walk is a directed path Vi1Vi2Vil1Vil (potentially listed as the corresponding sequence of edges).

    A multi-digraph G is a directed graph where multiple edges between the same vertices are allowed. In particular, a multi-digraph comes with two functions

    s:EVt:EV.

    where the functions s,t are the source and target function, respectively. Both self-edges (edges e with t(e)=s(e)) and parallel edges are possible.

    A RN on a finite set S is a digraph N=(C,R), where S is a finite set of species S={S1,,Sn}, C a potentially infinite set of complexes and R a potentially infinite set of reactions R={r1,r2}. Complexes are non-negative linear combinations of species, y=ni=1yiSi, identified with vectors y=(y1,,yn) in Zn0. Reactions are directed edges between complexes, written as ri:yiyi or, generically, as r:yy, potentially omitting ri,r. A reaction is said to consume the reactant y and create the product y. An RN is said to be finite if R is finite, and otherwise it is infinite.

    We diverge in two ways from the standard definition of RNs: trivial reactions r:yy (self-loops) are allowed, and the numbers of complexes and reactions are allowed to be infinite. Both extensions are useful when dealing with reduced RNs. From a dynamical point of view, trivial reactions might always be ignored as the dynamics is the same with and without them. Realistic model of bursty gene expression with infinitely many reactions have been proposed in the literature [24,25], where [24] focusses on corresponding theory for extensions of density-dependent Markov processes and [25] heuristically investigate such models. However, our motivation is not to accommodate such examples, but rather to ensure that an RN obtained by reduction of a finite RN is also an RN, finite or infinite. The construction of the reduced RN also holds even if the original RN is infinite.

    Definition 2.1. (ⅰ) Two species interact if they both appear in a complex of C.

    (ⅱ) A subset US is non-interacting if it contains no pair of interacting species, and the stoichiometric coefficients of the species in U in all complexes are either 0 or 1. The species of U are said to be non-interacting, and the species in SU are said to be core species.

    Example 2.2. Consider a two-substrate Michaelis Menten mechanism [26,Section 3.1.2]:

    r1:E+AEA,r2:EAE+A,r3:EA+BEAB,
    r4:EABEA+B,r5:EABE+P+Q,

    where E is an enzyme catalyzing the conversion of two substrates A,B into two other substrates P,Q by means of transient (or intermediate) steps; here EA and EAB are known as intermediate complexes formed by binding of the molecules in the reactants.

    This RN has species set S={E,A,B,EA,EAB,P,Q} and complex set C={E+A,EA,EA+B,EAB,E+P+Q}.

    The sets U1={EA,EAB}, U2={EA,P} and U3={EAB} are non-interacting, with corresponding core species O1=SU1={E,A,B,P,Q}, O2=SU2={E,A,B,EAB,Q}, and O3=SU3={E,A,EA,B,P,Q}. In particular, a species might be a core or a non-interacting species, depending on the choice of U.

    In this paper, it is convenient to work with the directed stoichiometric subspace of N=(C,R), defined as

    T={r:yyRαr(yy)|αr>0,rR,whereRRisfinite}Rn

    (note, this definition allows the RN to be infinite). For vRn, the set (v+T)Rn0 defines a directed stoichiometric compatibility class of N. The RN is said to be conservative, respectively, sub-conservative, if there exists a positive vector cRn>0, such that c,yy=0, respectively, c,yy0 for any reaction r:yyR. A sub-conservative RN has compact directed stoichiometric compatibility classes (but not stoichiometric compatibility classes). For a weakly reversible RN, the directed and the undirected stoichiometric subspaces are the same.

    An SRN is an RN together with a CTMC X(t),t0, on Zn0, modeling the number of molecules of each species over time. A reaction r:yy fires with transition intensity λr(x), in which case the state jumps from X(t)=x to x+yy [27]. The Markov process with transition intensities λr:Zn0R0, rR, has Q-matrix

    Q(x,x+ξ):=r:yyR:yy=ξλr(x).

    For (stochastic) mass-action kinetics, the transition intensity for r:yy is

    λr(x)=κr(x)!(xy)!1{x:xy}(x),xZn0,

    where x!:=ni=1xi!, and κr is a positive reaction rate constant [27]. If there are infinitely many reactions, we assume

    rRλr(x)<,forallxZn0, (2.1)

    such that the corresponding CTMC is well-defined in the sense that it has no instantaneous jumps [28].

    When the reactions are indexed, R={r1,r2,}, we occasionally write λi and κi, i=1,2,, for convenience. The following assumption holds in particular for stochastic mass-action kinetics.

    Assumption 1. For all reactions, r:yyR, the transition intensity λr:Zn0R0 satisfies the following

    λr(x)>0xy.

    Under Assumption 1, (X(t))t0, stays in Zn0, if X(0)=xZn0. In particular, (X(t))t0, is confined to the directed stoichiometric compatibility class (x+T)Zn0. Assumption 1 is fundamental and enforces the reactions to be compatible with the transition intensities of the CTMC.

    In the following, we focus on RNs and SRNs with non-interacting species. To fix notation, let US be a non-interacting subset of species. For simplicity, we let U={U1,,Um} and O=SU={S1,,Sp} (p=nm), such that S=OU={S1,,Sp,U1,,Um}. We let

    ρO:RnRp,ρU:RnRm

    be the projections onto the first p coordinates and last m coordinates of Rn, respectively. Consequently, we denote a state by x=(z,u)Zp0×Zm0=Zn0.

    Define the following sets of reactions

    RU={r:yyRρU(y)0},RU={r:yyRρU(y)0},

    such that RURU are the reactions that involve species in U.

    We consider a subset of fast (a terminology to be motivated below) reactions FRU with the following structural property.

    Definition 3.1. The reactions in F are proper w.r.t. U, that is, any non-interacting species is part of a sequence of reactions in (RURU)F of the form

    ri0:yi0yi0,ri1:yi1yi1,ril:yilyil,

    with i0,,il{1,,k}, ρU(yi0)=ρU(yil)=0 and ρU(yij)=ρU(yij1)0 for j=1,,l. Such a sequence of reactions is called a fast chain.

    Hence, any molecule of a non-interacting species can be degraded through a sequence of fast reactions (provided sufficient molecule numbers of core species).

    Example 3.2. Recall Example 2.2. The set U1={EA,EAB} is a non-interacting set and F=RU1 is a set of fast proper reactions. For example, the following is a fast chain:

    r1:E+AEA,r3:EA+BEAB,r5:EABE+P+Q.

    Also, U2={EA,P} is a non-interacting set of species, but F=RU2 is not a set of fast proper reactions, as there are no fast reactions with P in the reactant.

    We introduce a labeled multi-digraph to capture the conversion and creation of non-interacting species through fast chains. This graph is similar to the one introduced in [21]. We later use this graph to define a reduced RN and a reduced SRN by elimination of non-interacting species.

    Definition 3.3. Let N=(C,R) be an RN on S, US a set of non-interacting species, and FRU a set of fast proper reactions. Let GU,F=(VU,F,EU,F) be the labeled multi-digraph with vertex set

    VU,F:={in,out}{UiUiU},

    and edge set

    EU,F:={inrUir:yyRURU,ρU(y)=Ui}{UirUj,r:yyFsuchthatρU(y)=Ui,ρU(yγ)=Uj}{Uirout,r:yyFsuchthatρU(y)=Ui,ρU(y)=0}.

    Furthermore, let L:EU,F{1,} be the function that maps γEU,F to the corresponding index of the reaction label, that is, if γEU,F has label ri, then L(γ)=i.

    The construction of the graph GU,F also makes sense in the case of an infinite RN. The vertex set is finite, but the number of edges between any two vertices might be infinite.

    Example 3.4. Consider Example 2.2 with U=U1={EA,EAB} and F=RU. Then GU,F is

    There are infinitely many walks from in to out, as one might take arbitrary many 'rounds' in the loop before exiting.

    By definition, any finite sequence of reactions that corresponds to the reaction labels of a walk in GU,F with start vertex in and end vertex out is a fast chain. Define the set of such walks by

    WU,F:={(γ1,,γl)(EU,F)ll2,(γ1,,γl)isawalkinGU,Fwiths(γ1)=in, t(γl)=out}.

    Note that WU,F might be infinite, even for finite RNs, as in Example 3.4 above.

    We will consider the situation in which the reactions of F occur fast compared to the remaining reactions RF, in the sense that the transition intensities of the reactions in F are scaled by 1/ϵ for small ϵ. Thus, we consider a fast-slow dynamic regime. In that case, it is natural to expect that whenever a non-interacting molecule is created, then it will be degraded almost instantaneously through fast reactions, before any other non-fast reaction occurs. Such sequences of reactions (creation and degradation) are encoded in the walks of WU,F. To understand the fast dynamics, it is therefore important to understand the net gain of core species in the walks and their probabilities of occurring.

    In preparation for this, consider a walk Γ=(γ1,γ2,,γl)WU,F, and denote

    wi:=yL(γi)+i1j=1yL(γj)yL(γj),i=1,,l,

    where L(γj) is the reaction index of the edge γj. Define

    r(Γ):=max(w1,,wl).

    Note that r(Γ) depends on the order of the elements of Γ.

    Lemma 3.5. Let Γ=(γ1,γ2,,γl)WU,F. Then the following holds:

    r(Γ)0, and p(Γ):=r(Γ)+li=1yL(γi)yL(γi)0

    pU(r(Γ))=0, pU(p(Γ))=0.

    Proof. We give complete proofs here for convenience, but note that the results also follow from [29].The second item follows by definition. As we take the maximum coordinate-wise and w10, then r(Γ) has non-negative coordinate in each species. To see that p(Γ) has non-negative coordinate in each species, we note that by definition yL(γl)+l1i=1yL(γi)yL(γi)r(Γ). Adding li=1yL(γi)yL(γi) to both sides, we get 0yL(γl)p(Γ), and the result follows.

    By Lemma 3.5, we might consider r(Γ) and p(Γ) as elements of Zp0, and as reactant and product, respectively, of the reduced reaction r(Γ)p(Γ), obtained by contraction along Γ. Lemma 3.7 below justifies this view, and relates the compatibility requirement in Assumption 1 to the reduced reactions.

    Example 3.6. We continue with Example 3.4. Consider the walk Γ0 of GU,F consisting of the two edges associated to r1,r2. Then

    r(Γ0)p(Γ0)=E+AE+A,

    a trivial reaction. Likewise, the walk ΓA1 consisting of the edges associated to r1,r3,r4,r2, also results in a trivial reaction,

    r(ΓA1)p(ΓA1)=E+A+BE+A+B.

    In contrast, the walk ΓB1 consisting of the three edges associated to r1,r3 and r5 yields

    r(ΓB1)p(ΓB1)=E+A+BE+P+Q.

    Any other walk of WU,F consists either of a sequence of edges with labels r1,r3,r4,, r3,r4,r3,r5 or a sequence of edges with labels r1,r3,r4,,r3,r4,r2. As the net gain of core species in the contraction of the reactions r3,r4 is zero, the corresponding reduced reactions are the same as that of ΓA1, respectively, ΓB1. Hence, there is only one non-trivial reduced reaction. For future reference, denote the corresponding walks with n instances of r3 by ΓAn, respectively, ΓBn, for n1.

    Lemma 3.7. Suppose Assumption 1 holds. LetΓ=(γ1,γ1,,γl)WU,F. Furthermore, let zZp0 and x=(z,0)Zn0. Then

    λL(γj)(x+j1i=1yL(γi)yL(γi))>0,j=1,,l1,

    if and only if zr(Γ).

    The lemma implies that the reactions corresponding to a walk in WU,F can fire in succession of each other (without other non-fast reactions firing in between), if and only if the present molecule counts of core species is larger or equal to r(Γ). A similar statement appears in [29,Corollary 3.2].

    Lemma 3.7 allows us to define transition intensities of the reduced reactions in a natural way. For Γ=(γ1,γ2,,γl)WU,F, define the function ΛΓ:Zp0R0,

    ΛΓ(z):=λL(γ1)(z,0)lj=2λL(γj)((z,0)+j1i=1ξi)˜γout(s(γj))λL(˜γ)((z,0)+j1i=1ξi), (3.1)

    where ξi:=yL(γi)yL(γi), and out(s(γj)) denotes the set of outgoing edges of s(γj) in GU,F. Each term in the product is the probability that the desired reaction is chosen out of all possible fast reactions with the same non-interacting species in the reactant. The convention 0/0=0 is used. Even if the RN is infinite, then (3.1) is well-defined due to (2.1).

    Definition 3.8. Let N=(C,R) be an RN on S, let US be a non-interacting set of species, and FRU a set of proper fast reactions. The reduced RN NU,F=(CU,F,RU,F) on O=SU, obtained by elimination of (U,F) from N, is the possibly infinite RN defined by

    RU,F=(R(RURU)){r(Γ)p(Γ)ΓWU,F},

    and CU,F being the set of vertices of RU,F.

    For rRU,F, define

    WU,F(r):={ΓWU,Fr(Γ)p(Γ)=r}.

    Definition 3.9. Let N=(C,R) be an SRN on S with transition intensities λr(),rR, satisfying Assumption 1. Let US be a set of non-interacting species, and FRU a set of proper fast reactions. Then, the possibly infinite SRN NU,F=(CU,F,RU,F) on O=SU with transition intensities τr(),rRU,F, given by

    τr(z):={ΓWU,F(r)ΛΓ(z)ifrR(RURU),λr(z,0)+ΓWU,F(r)ΛΓ(z)ifrR(RURU),λr(z,0)ifrR(RURU)andr{r(Γ)p(Γ)ΓWU,F},

    is the reduced SRN, obtained by elimination of (U,F) from N.

    As a reduced SRN has species set O=SU, the associated CTMC lives in Zp0. Even if the reduced SRN has infinitely many reactions, the associated CTMC is always well-defined, that is, it has no instantaneous states [28]. If U consists of intermediate species, then the reduced reaction network and the corresponding transition intensities are easily found [20,Theorem 3.1]. For intermediate species, the number of molecules of core species does not change along a walk, in contrast to for instance Example 3.6. In particular, if the original SRN has mass-action kinetics, then the reduced SRN has mass-action kinetics as well.

    As remarked earlier, one might discard any trivial reaction. However, the transition intensities of the trivial reactions play a crucial role in Assumption 2 in Section 4.1.

    Lemma 3.10. Suppose Assumption 1 holds. Then, the reduced SRN has Q-matrix with all off-diagonal row sums finite.

    Proof. By definition, we need to show that

    rRU,Fτr(z)<.

    For this, we note that

    rRU,Fτr(z)=rR(RURU)λr(z,0)+ΓWU,FΛΓ(z).

    Hence, it is enough to show that the second summand in the second term is finite as the first is finite by assumption. By construction of the reduced transition intensities,

    ΓWU,FΛΓ(z)rRURUλr(z,0)<.

    The left hand side is the sum of all transition intensities of walks (reduced reactions) with initial reaction in RURU. This sum is at most equal to the right hand side: Let Γ=(γ1,,γl) be a walk and yL(γ1) be the reactant of the first reaction in the walk. If yL(γ1)zr(Γ) and r(Γ)z, then the first reaction can fire, but not all reactions in the walk can fire in succession, hence the corresponding ΛΓ(z) is zero. For a concrete illustration, see Example 5.2.

    As a consequence of Lemma 3.7, Assumption 1 holds for a reduced SRN, provided it holds for the original SRN.

    Corollary 3.11. Suppose Assumption 1 holds. Then, for any reaction r:yyRU,F of the reduced SRN, it holds that τr(z)>0 if and only if zy.

    Proof. It is enough to prove it for ΛΓ(z), where ΓWU,F and zZp0, that is,

    ΛΓ(z)>0zr(Γ).

    Only one direction is not obvious. Assume zr(Γ). Using the definition of ΛΓ(), it is sufficient to show that the numerators in the fraction are all non-zero. This holds by Lemma 3.7, hence ΛΓ(z)>0.

    Example 3.12. We continue with Examples 3.4, 3.6, assuming stochastic mass-action kinetics. There are three reduced reactions, the first reaction being non-trivial and the last two trivial,

    s1:E+A+BE+P+Q,s2:E+AE+A,s3:E+A+BE+A+B,

    with infinitely many walks underlying the first and third reduced reaction. Here, si is used to denote the reactions of the reduced SRN, rather than ri, to distinguish the reactions of the reduced RN from those of the original RN. We find

    ΛΓ0(z)=κ1κ2zEzAk3zB+κ2,
    ΛΓAn(z)=κ1κ2zEzAk3zB+κ2(κ3κ4zB(κ4+κ5)(k3zB+κ2))n,n1,
    ΛΓBn(z)=κ1κ3κ5zEzAzB(κ3zB+κ2)(κ4+κ5)(κ3κ4zB(κ4+κ5)(k3zB+κ2))n1,n1,

    such that

    τ1(z)=n=1ΛΓBn(z)=κ1κ3κ5zEzAzB(κ4+κ5)(κ3zB+κ2)κ3κ4zB.

    Analogously, we compute the sums over walks for the trivial reactions. We have τ2(z)=ΛΓ0(z) and

    τ3(z)=κ1κ3κ5zEzAzB(k3zB+κ2)((κ4+κ5)(κ3zB+κ2)κ3κ4zB).

    Furthermore, a simple calculation gives the following identity for zZp0,

    ΓWU,FΛΓ(z)=τ1(z)+τ2(z)+τ3(z)=λ1(z,0).

    An interesting observation is that the kinetics is mass-action-like in the sense that the numerator alone is of mass-action form, while the denominator is positive (for any state). This is, however, not true in general, see Example 5.2.

    Example 3.13. Consider the RN,

    r1:S1U1,r2:U1S3,r3:S2+U1U1,

    with U={U1} and F=RU proper. Taken with mass-action kinetics with corresponding rate constants, κ1,κ2,κ3, the reduced SRN has infinitely many reactions, sn:S1+nS2S3, n0, with transition intensities,

    τn(x)=κ1κ2xS1κ2+κ3(xS2n)n1i=0κ3(xS2i)κ2+κ3(xS2i),n0,

    where by convention 1i=0=1. On any particular state, at most finitely many reactions can be active, that is, have non-zero transition intensity.

    For examples from the biochemical literature, we refer to § 5.1.

    We study transient approximability of the CTMC via the reduced SRNs under appropriate assumptions. We cover two settings, the standard time-homogeneous CTMC setting for SRNs, and afterwards the setting where the transition intensities of the SRN are allowed to be time-dependent.

    Let N=(C,R) be an SRN on a species set S (of size n) with non-interacting species US, fast proper reactions FRU, and transition intensities λr,rR. From this SRN, we construct a family of SNRs, indexed by a parameter ϵ>0, with corresponding Q-matrix Qϵ. In particular, the transition intensities of the reactions in F consuming non-interacting species are scaled in ϵ as follows:

    λϵr(x):=1ϵλr(x), rF, that is, for small ϵ the reactions are fast compared to the remaining reactions,

    λϵr(x):=λr(x), rRF, that is, the transition intensities are independent of ϵ, and the reactions are slow.

    Thus, we might write the corresponding Q-matrix as as sum of two terms, a fast part ˜Q, scaled by 1ϵ, and a slow part ˆQ,

    Qϵ=1ϵ˜Q+ˆQ.

    It follows from Definition 3.9 that the reduced SRN, obtained from the SRN with transition intensities λϵr, rR, has transition intensities τr, rRU,F, independent of ϵ.

    Let (Xϵ(t))t0 be Markov chains on Zn0 with generators Qϵ, ϵ>0, respectively. We note that all Qϵ, ϵ>0, are dynamically equivalent in the sense that the connectivity of the state space and the decomposition of the state space Zn0 into communicating classes are independent of ϵ. Furthermore, let (X0(t))t0 be a Markov chain on Zp0 (excluding the m=np coordinates for non-interacting species) with generator Q0, obtained from the transition intensities τr, rRU,F. In the following, we provide conditions that guarantee that the dynamics of (Xϵ(t))t0 is similar (in a sense to be made precise) to the dynamics of (X0(t))t0, whenever ϵ is small. Technically, we will consider the limit as ϵ0.

    We restrict attention to a particular closed set E of (Xϵ(t))t0, such that E(Zp0×{0}). That is, there are states in E with no molecules of non-interaction species being present. If the reduced SRN starts in X0(0)E0:=ρO(E(Zp0×{0})), then by construction of the reduced reactions, X0(t)E0 for all t>0. Furthermore, E0 is a closed set of Q0.

    We require the following.

    Assumption 2. The closed set E is finite, and all zE0 satisfy the following:

     ΓWU,FΛΓ(z)=rRURUλr(z,0) (4.1)

    The sum on the left hand side also includes walks giving rise to trivial reactions, as in Example 3.12. The condition might fail in two ways. Either a walk is blocked because of lack of molecules of core species (for example r1:0U, r2:S+U0; if there is no molecules of S, then the fast chain r1,r2 is blocked and a molecule of U cannot be degraded), or one might be trapped in infinite walks (for example, r1:0U, r2:2S+U3S+U, r3:U0; there is positive probability of an infinite walk r1,r2,r2, without degradation of the U molecule). The latter cannot occur if E, and hence E0, are finite as this implies that at most finitely many (reduced) reactions can be active on any state of E (E0).

    For a sub-conservative RN, any closed set of states E (not necessarily a communicating class) in a directed stoichiometric compatibility class is finite. If Assumption 1 is satisfied, then the conditions in Assumption 2 do not depend on the transition intensities λr, rR, but only on the structure of the underlying RN (together with U and F). Furthermore, as (2.1) holds by assumption, then Assumption 2 is meaningful even for infinite RNs. In particular, Theorem 4.1 also holds for infinite RNs as E is finite.

    For a subset BE(Zp0×{0}), we define B0=ρO(B).

    Theorem 4.1. Let N=(C,R) be an SRN on S with transition intensities λr, rR, US a set of non-interacting species, and FRU a set of fast proper reactions. Suppose Assumptions 1 and 2 hold. Let π be a probability distribution on E(Zp0×{0}), and π0 the induced probability distribution on E0 by omitting the last np coordinates of the states of E. Then, the following holds for any 0<T:

    supt[0,T]|Pπ(Xϵ(t)B)Pπ0(X0(t)B0)|=O(ϵ)forϵ0.

    In particular, for any BE(Zp0×{0}) and any 0<T:

    limϵ0supt[0,T]|Pπ(Xϵ(t)B)Pπ0(X0(t)B0)|=0.

    The proof is in Appendix § B.3. As a consequence of the theorem we have:

    Corollary 4.2. Let N=(C,R) be an SRN on S with transition intensities λr, rR, US a set of non-interacting species, and FRU a set of fast proper reactions. Suppose Assumptions 1 and 2 hold. Let x=(z,0)E(Zp0×{0}) and BE(Zp0×{0}). Then, for all t0 it holds that:

    limϵ0Px(Xϵ(t)B)=Pz(X0(t)B0).

    If the initial state Xϵ(0) has more than one molecule of the non-interacting species, then these will in general be depleted quickly for small ϵ. This is however not always the case, see Example 5.2.

    In the previous section, we considered the scaling limit of time-homogeneous SRNs. While time-constant transition intensities are reasonable in many situations, time-dependence becomes relevant, for example, in connection with variation in experimental setups, cycle-dependent mechanisms or temperature changes [30].

    The setup is essentially the same as in Section 4.1, except we now allow time-dependent transition intensities, λr(t,x), rR, for t in a compact time interval t[0,T]. We make the following assumption.

    Assumption 3. The time-dependent transition intensities, λr(t,x), rR, are such that λr(t,) satisfies Assumption 1 for all t[0,T] and such that for all xZn0, λr(,x), rR, are C1-functions from [0,T]R0 with Lipschitz derivatives.

    The transition intensities are scaled as in the homogeneous case. Under Assumption 3, the reachability properties of the CTMC are invariant over time and match the reachability properties of the homogeneous SRNs under Assumption 1. Assumption 2 holds for all t[0,T], provided it holds for one t, as the assumption is structural.

    With these changes and amendments, then Theorem 4.1 and Corollary 4.2 hold as well for the time-dependent case. A proof is sketched in Appendix C.

    We next illustrate the reduction procedure with various examples. Furthermore, we derive sufficient conditions for Assumption 2 to hold, and compare the long-term behavior of the CTMC of the reduced SRN with the one of the original SRN.

    In this section, we give realistic examples to elaborate on the reduction. All examples are taken with stochastic mass-action kinetics, hence they satisfy Assumption 1. To indicate mass-action kinetics, we put the reaction rate constants κi as labels of the reactions, and omit the reaction names ri. A key source book for realistic examples is [26], that contains reaction networks used in systems biology. Other useful references are [31], which focuses on enzyme kinetics, and [1], which has examples from a range of areas in biology.

    Example 5.1. We first consider the two-substrate Michaelis-Menten mechanism [26,Section 3.1.2]; also discussed in Examples 2.2, 3.4, 3.6, and 3.12, and repeated here for convenience with reaction names replaced by their reaction constants:

    E+Aκ1κ2EA,EA+Bκ3EAB
    EABκ4EA+B,EABκ5E+P+Q.

    The mechanism is commonly studied as a deterministic mass-action system, using Tikhonov-Fenichel singular perturbation theory or the QSSA, with short-lived intermediate complexes (non-interacting species) U={EA,EAB} and fast reactions F=RU={r2,r3,r4,r5}. Assuming the same in the stochastic setting gives the reduced RN in Example 3.6{or Example 3.12 with} one non-trivial reaction,

    s1:E+A+BE+P+Q,

    with transition intensity

    τ1(z)=κ1κ3κ5zEzAzB(κ4+κ5)(κ3zB+κ2)κ3κ4zB.

    We omit the trivial reactions; further details are given in example 3.12. Assumption 2 is fulfilled for all zZ50 (p=5, see Example 3.12), in particular the original RN is conservative. Then, Theorem 4.1 applies to any closed set E in a directed stoichiometric compatibility class that intersects Z50×{0} non-trivially.

    Example 5.2. Consider the two-substrate Michaelis-Menten mechanism with U={EA,EAB} as above, but with set of fast reactions F={r3,r4,r5}RU. We remove r2 from the fast reactions in the previous example. Then there is only one reduced reaction in contrast to Example 5.1 that additionally had two trivial reactions, given by

    s1:E+A+BE+P+Q,

    with transition intensity

    τ1(z)=κ1zEzA1{zB:zB1}.

    Assumption 2 holds for all z=(zA,zB,zE,zP,zQ)Z50 with zB>0:

    ΓWU,FΛΓ(z)=τ1(z)=κ1zEzA1{zB:zB1}=κ1zEzA=λ1(z,0).

    Also note that for zA1,zE1 and zB=0, we have 0=τ1(z)<κ1zEzA=λ1(z,0)0. This is an example of the inequality in the proof of Lemma 3.10.

    This example does not have mass-action-like kinetics, as τ1(z) cannot be expressed as a fraction with positive denominator such that the numerator is of mass-action form. Alternatively, we might extend by multiplication and division by zB.

    Example 5.3. Consider a non-competitive inhibition network [26,Section 3.2.2]:

    S+Eκ1κ2ESκ3E+P,I+Eκ4κ5EI,
    I+ESκ6κ7ESI,S+EIκ8κ9ESI,

    where E is an enzyme that catalyze the conversion of a substrate S into a substrate P. The conversion is delayed by an inhibitor I that binds to the enzyme and to the substrate through the intermediate complex EI. It is non-competitive in the sense that I does not compete with E for substrate binding, but rather acts on E directly; a phenomenon known as allosteric regulation.

    This example is typically analyzed by means of the QSSA with short-lived species; here U={ES,EI,ESI}. In our setting, taking U to be non-interacting species with F=RU, then the reduced SRN becomes

    s1:S+2IS+I+E,s2:S+IE+P,s3:S+2IE+I+Ps4:I+E+SI+E+P,s5:I+E+S2I+S,

    with transition intensities

    τ1(z)=κ1κ5κ6κ9zSzI(zI1)(κ2+κ3)(κ7(κ5+κ8zS)+κ5κ9)+κ6κ5κ9(zI1)τ2(z)=κ1κ3zSzIκ2+κ3+κ6(zI1)τ3(z)=κ1κ3zSzI(κ7(κ5+κ8zS)+κ5κ9(κ2+κ3)(κ7(κ5+κ8zS)+κ5κ9)+κ6κ5κ9(zI1)1κ2+κ3+κ6(zI1))τ4(z)=κ3κ4κ7κ8(κ7+κ9)zEzIzS[κ5(κ7+κ9)+κ8κ7zS]((κ2+κ3)(κ7+κ9)+κ6κ9zI))κ8κ7κ6κ9zIzSτ5(z)=κ2κ4κ7κ8(κ7+κ9)zEzIzS[κ5(κ7+κ9)+κ8κ7zS]((κ2+κ3)(κ7+κ9)+κ6κ9zI))κ8κ7κ6κ9zIzS.

    \medskip For any closed set E in a directed stoichiometric compatibility class that intersects Z40×{0} non-trivially, Assumption 2 is satisfied.

    Example 5.4. Consider an example of allosteric activation [26,Problem 3.7.8] that models the regulation of an enzyme by binding of an allosteric activator before the enzyme can bind a substrate:

    R+Eκ1κ2ER,ER+Sκ3κ4ERSκ5P+ER.

    Here, R is the allosteric activator, S the substrate, E the enzyme, and P the product. While the RN is reminiscent of the the two-substrate example 5.1, the enzyme-activator complex ER stays intact after the product P dissociates in reaction r5.

    To analyze the behavior of the system, often the QSSA is applied to a mass-action ODE system with short-lived intermediate complexes (non-interacting species) U={ER,ERS}. Choosing analogously F=RU in the stochastic case gives a proper set of fast reactions. Then, the reduced RN has infinitely many reactions, given by:

    s2k1:E+R+kSE+R+kP,s2k:E+R+(k+1)SE+R+kP+S,

    for k1, with transition intensities

    τ2k1(z)=κ1κ2zRzEκ2+κ3(zSk2)k1i=0κ3κ5(zSi)κ2κ4+κ2κ5+κ3κ5(zSi),
    τ2k(z)=τ2k1(z)κ3κ4(zSk)κ2κ4+κ2κ5+κ3κ5(zSk).

    All zZ40 (p=4) satisfy (4.1). As N is sub-conservative, any closed set E of a directed stoichiometric compatibility class is finite. Hence, for any closed set E in a directed stoichiometric compatibility class that intersects Z40×{0} non-trivially, Assumption 2 is satisfied.

    Example 5.5. As a final example, consider a mechanism-based inhibitor system [1,Section 6.4],

    S+Eκ1κ2Xκ3Yκ5E+P,Yκ4Ei,

    where S,P are substrate and product, respectively, X,Y intermediate complexes, E an enzyme in active form (implying it might bind to the substrate) and Ei, the enzyme in its inactivated form. A main interest is to know the final ratio of the product to the inactivated enzyme [1,Section 6.4]. For this, often the QSSA is applied to the short-lived species U={X,Y}. Analogously, we consider the stochastic system with F=RU being fast and proper, and study the corresponding stochastic reduction. The reduced SRN has reactions

    s1:S+EE+P,s2:S+EEi

    with transition intensities:

    τ1(z)=κ1κ3κ5zSzE(κ2+κ3)(κ4+κ5),τ2(z)=κ1κ3κ4zSzE(κ2+κ3)(κ4+κ5).

    As U={X,Y} are intermediate species, all zZ40 (p=4) satisfy (4.1). As N is sub-conservative, for any closed sets E of a directed stoichiometric compatibility class Assumption 2 is satisfied.

    While Assumption 1 and F being proper are both easy to check, Assumption 2 is non-trivial in general. Assumption 2 requires a finite closed set of the CTMC and that molecules of non-interacting species can be degraded through chains of fast reactions. The following gives sufficient conditions for this to hold (see Appendix § A for proof).

    Proposition 5.6. Assume N is a sub-conservative SRN on a species set S, US is a set of non-interacting species, and FRU is a proper set of reactions. Suppose Assumption 1 holds. Furthermore, assume that one of the following holds:

    (a) F(RURU) is weakly reversible.

    (b) FRURU is weakly reversible, and for any reaction r:yy in RURU, the reverse reaction r:yy is in F.

    (c) N is weakly reversible and F=RU.

    Then, for an arbitrary closed set in a (directed) stoichiometric compatibility class, Assumption 2 is satisfied.

    None of the conditions in the theorem are necessary for Assumption 2 to hold. For intermediate species a stronger result holds.

    Lemma 5.7. Assume N is an SRN on a species set S, and that US is a set of intermediate species. Suppose Assumption 1 holds. Then, FRU is proper if and only if all zZp0 satisfy (4.1) of Assumption 2.

    We summarize the properties of the examples from § 5.1. In examples 5.1, 5.3 and 5.4 we can directly conclude by Proposition 5.6 (b) that Assumption 2 holds.

    A natural question following the transient approximability of SRNs with non-interacting species is whether the reduced SRN approximates the stationary behavior in the limit as ϵ0. First one might ask whether states x=(z,0)Zp0×{0} and z are of the same type (positive recurrent/null recurrent/transient) for N and NU,F, respectively. Furthermore, the stationary distribution (if it exists) for the scaled SRN N as ϵ0 might match the stationary distribution of NU,F. The following examples show that such correspondences do not hold in general, even for reduction by intermediates.

    Example 5.8. Consider the following SRN N with mass-action kinetics,

    S1κ1κ2U1κ3S2κ4S1,U1κ5S3.

    Choosing U={U1}, F={r3}, then the reduced SRN NU,F is S1S2 with transition intensities satisfying Assumption 1 by Corollary 3.11. Any (a,b,c,0) with a+b1 is transient for N, but recurrent for NU,F. Reaction r5 in N causes absorption into (0,0,a+b+c+d,0). As r5 is not present in the reduced SRN, this cannot happen in NU,F.

    Example 5.9. Consider the SRN N with mass-action kinetics,

    S1κ1κ2U1κ3κ4S2

    Choosing U={U1}, F={r3}, then the reduced SRN NU,F becomes S1S2. The state (a,b,0) with a1 is recurrent for N, but transient for NU,F.

    The next result follows from [29,Theorem 5.6]. We recall that irreducible components of sub-conservative RNs are finite [27]. For an irreducible component EZn0, define E0=ρO(E(Zp0×{0})).

    Theorem 5.10. Let N=(C,R) be a sub-conservative SRN on S, US a set of intermediate species, and F=RU a set of proper reactions. Suppose Assumption 1 holds. Then x=(z,0)Zp0×{0} is transient for N if and only if z is transient for NU,F, and the same holds for positive recurrence. In particular, if a subset E is an irreducible component for N then E0 is an irreducible component for NU,F.

    From Theorem 5.10, Lemma 5.7 and [32,Theorem 3], we get convergence of the stationary distribution.

    Corollary 5.11. Let N=(C,R) be a sub-conservative SRN on S, US a set of intermediate species, and F=RU a set of fast proper reactions. Suppose Assumption 1 holds. Let E be an irreducible component, and let E0 be the corresponding projected set. Furthermore, denote by πϵ the unique stationary distribution of Qϵ on E, and let π0 be the unique stationary distribution of the reduced SRN on E0. Then, πϵ((z,0))π0(z), for zE0, and ϵ0.

    We compare our setting for reduction of SRNs with non-interacting species to other approaches of model simplification. Then we discuss assumptions, extensions and examples that are not covered by our approach.

    In our treatment, we consider SRNs with fast reactions consuming non-interacting species, generalizing earlier settings in three ways: 1) extending from intermediate species to non-interacting species, 2) allowing arbitrary transition intensities satisfying a compatibility condition, and 3) allowing non-homogeneous Markov dynamics, that is, time-dependent transition intensities. As the set-up is based mostly on singular perturbation of Markov chains with transient states [33], which is also available in the book [22], we loose the possibility to study multiple time scales, but are restricted to two scales.

    Reductions of SRNs with non-interacting species resembles reduction by intermediate species as intermediate species form a special class of non-interacting species [20]. In contrast to our setting, [20] allows a multi-scale setting. Other multi-scale reductions require the parameters to fulfill a balancing equation between scales of species concentrations and scales of transition intensities [13,16,17,18,36], where [36] treats the classical scaling for convergence to the ODE. This is not required in our setting (nor in [20] for the reactions involving intermediate species), in fact such equation will not hold. Balancing is violated because fast reactions do not have 'limits' themselves as ϵ0, but only jointly through contraction of sequences of reactions. Furthermore, at the process-level, we do not have convergence in the Skorohod topology, see [20,Example 5.3] or [18,§ 6.5], because even for small ϵ, molecules of non-interacting species are created and exist for small amounts of time, while this is not possible in the reduced SRN. Convergence in Skorohod topology typically require that species abundance is measured in concentrations, rather than in molecule numbers. In contrast, we might have convergence in the Jakubowski topology from [34], also see [18,Remark A.14] for more in the context of reaction networks. Whenever, a molecule of a non-interacting species is produced, the Markov chain (Xϵ(t))t0 will jump a finite number of times (stochastically bounded in ϵ) in a small time interval before the molecule is degraded again. This implies the (so-called) up- and downcrossing numbers of [34] is bounded, a prerequisite for convergence.

    The results on the transient dynamics in our two-scale setting could potentially be extended by allowing general infinite state space (that is, weaken Assumption 2). In so, the transition rates of the original CTMC become unbounded for mass-action kinetics, and the theory on singular perturbations for CTMCs are not applicable [22]. Furthermore, the original process might have sample paths that diverge to infinity in a finite amount of time. Hence, theory applicable to unbounded cases as well as conditions ensuring non-explositivity would be of interest to develop. We note that even 1-dimensional CTMCs arising from RNs can have both positive recurrent and explosive irreducible components [35], implying that different parts of the state space might have to be analyzed separately.

    Another generalization would be to allow many scales rather than two scales. This could be done using [32] in the finite state space case. However, it becomes difficult to find the reduced reactions, as these depend on the concrete form of the transition intensities and their scalings. Also, a reduction might not be interpreted as SRNs fulfilling Assumption 1. As an example, consider

    S1κ1U11ϵκ2S2,U1+S31ϵ2κ3S4

    with stochastic mass-action kinetics. The reduction for small ϵ can be written as two reactions

    S1S2,S1+S3S4

    with transition intensities κ1zS11{0}(zS3) for the first reaction and κ1zS11N{0}(zS3) for the second. In particular, the first transition intensity invalidates Assumption 1.

    In our approach, it is important that non-interacting species are produced and degraded. Consider the SRN with reactions

    U1κ1κ2U2,S1+U1κ3κ4S2+U2.

    Neither U1 nor U2 are produced nor degraded, hence there does not exist a proper set of reactions F and the example falls outside our setting. As a matter of fact, the scaling parameter ϵ would apply uniformly to all reactions as they all transform one non-interacting species into another. Hence, the distribution of the CTMC approaches the stationary distribution within a short time span (for small ϵ). This distribution is not concentrated on the part of the state space without molecules of non-interacting species. Rescaling time by ϵ would retrieve the original chain.

    The following lemma proves Proposition 5.6 on sufficient conditions for Assumption 2.

    Lemma A.1. Assume N=(C,R) is a sub-conservative SRN on a species set S with transition intensities λr, rR, that US is a set of non-interacting species, and FRU is a proper set of fast reactions. Suppose Assumption 1 holds.Then either of the following are sufficient for Assumption 2 to be satisfied for an arbitrary closed set in a (directed) stoichiometric compatibility class:

    (a) F(RURU) is weakly reversible.

    (b) FRURU is weakly reversible, and for any reaction rRURU, the inverse reaction is in F.

    (c) N is weakly reversible and F=RU.

    Proof. As N is sub-conservative, any closed set of a directed stoichiometric compatibility class is finite. Hence, it is enough to show that (4.1) holds for arbitrary x=(z,0)Zp0×{0}. As (2.1) holds by assumption, it is enough to show that for arbitrary rRURU with λr(z,0)0 the following holds:

    ΓWU,FΓstartswithrΛΓ(z)=λr(z,0).

    Consider the SRN with species set S and reactions F, denoted N|F, that we take with the same transition intensities as N. Consider the associated discrete-time Markov chain (DTMC) from the jump chain of N|F (cf. [28,p. 82]), which is the sequence of states taken by the CTMC. This will be denoted by (Yn)nN. By definition of ΛΓ(), ΛΓ(z)λr(z,0) corresponds to the probability that the transitions of the jump chain (Yn)nN when starting from (z,0)+yy comes from the sequence of reactions in Γ.

    Denote the set of reachable states from (z,0)+yy via N|F by N|F((z,0)+yy). As N is sub-conservative, N|F is sub-conservative and N|F((z,0)+yy) is finite. In the following, let eiZn0 denote the vector with Ui-coordinate entry equal to one and all other entries equal to zero.

    (a) Assume F(RURU) is weakly reversible. Then, if (z,0)+eiN|F((z,0)+yy), then also (z,0)N|F((z,0)+ei). Hence, any such (z,0)+ei is transient in the jump chain (Yn)nN of N|F. On the other hand by definition, any state of the form (z,0) is absorbing for the jump chain of N|F. As the set of reachable states is finite, the time until absorption is a.s. finite. So overall the jump chain ends in one of the absorbing states after a finite time, and

    ΓWU,FΓstartswithrΛΓ(z)λr(z,0)=1.

    For (b) and (c), the proofs are similar to (a).

    The proof of Theorem 4.1 is based on singularly perturbed and watched Markov chains. We therefore first introduce these concepts for the convenience of the reader.

    Singular perturbation theory for Markov chains is a well-developed topic in probability theory. For the convenience of the reader we restate and summarize results of [22,33] for homogeneous CTMCs on a finite state space, using their notation. For more on the two-scale setting for CTMCs, we refer to [22,37], where [37] is a book more generally treating multiscale methods in mathematics.

    The setting we introduce has a Markov chain with fast and slow components subject to so-called weak and strong interactions, where the fast components consist of absorbing, weakly irreducible or transient states of the fast dynamics. These cases are treated in detail in [22,§ 4.3,§ 4.4,§ 4.5].

    Let a sequence of homogeneous generators of a CTMC be given by

    Qϵ=1ϵ˜Q+ˆQ,ϵ>0,

    on a finite state space I. By rearrangement, the states can be divided into blocks of states IA,IW,IT according to their role in the Markov chain associated to ˜Q, which are absorbing, irreducible, and transient, respectively. The matrices ˜Q,ˆQ are partitioned accordingly into block-matrices and take the form

    ˜Q=(˜QAA˜QAW˜QAT˜QWA˜QWW˜QWT˜QTA˜QTW˜QTT)=(0000˜QWW0˜QTA˜QTW˜QTT), (B.1)
    ˆQ=(ˆQAAˆQAWˆQATˆQWAˆQWWˆQWTˆQTAˆQTWˆQTT).

    The states in IA with ˜QAA=0,˜QAW=0,˜QAT=0 correspond to absorbing states of ˜Q. The states in IT correspond to transient states of ˜Q (hence ˜QTT is a Hurwitz matrix, furthermore stable and non-singular see Lemma B.5). Finally, IW consists of the states that are part of (non-trivial) closed irreducible components, so ˜QWW in (B.1) can further be decomposed as

    ˜QWW=diag(˜Q1,˜Q2,,˜Ql)=(˜Q1˜Q2˜Ql), (B.2)

    where ˜Qi is an mi×mi matrix. We denote by νi the stationary distribution of ˜Qi for i=1,,l (which exist by assumptions on the state space).

    Remark B.1. Our presentation differs slightly from [22]. In that work, so-called weakly irreducible classes is used [22,Definition 2.7a]; classes containing exactly one closed irreducible component and possibly other states that lead to this component (which are then transient) in the case of time-homogeneous CTMCs. As we include transient states of ˜Q via IT, this is not a restriction. We note that our setting is included in [22,§ 4.5].

    For ϵ>0, the forward equation of the CTMC gives an ODE

    pϵ(t)dt=p(t)Qϵ,p(0)=π0,

    where π0 is the initial probability distribution.

    In [22], they construct sequences of functions that approximate pϵ(t) in the limit for ϵ0 uniformly on [0,T] for different powers of ϵ. One contribution comes from the so-called outer expansion, that approximates pϵ(t) for t>0. Another part comes from the so-called initial-layer correction, that approximates pϵ(t) in a neighborhood of t=0. This might be ignored as it is zero in our case [22]. Hence, we will focus on the zeroth order outer expansion as this is enough to obtain an approximation to order O(ϵ) (see Proposition B.3).

    Following [22,Theorem 4.45], the distribution of the CTMC converges to the zeroth order outer expansion φ0(t) for t>0. In the case we consider, it is given as

    φ0(t)=(φA(t),φW(t),φT(t))=(φA(t),φW(t),0T),=(φA(t),ϑ1(t)ν1,,ϑl(t)νl,0T),

    where ϑi(t) are scalar functions and νi are the stationary distributions from matrix (B.2), and 0T is the vector of zero-entries. Then, let (with m0=0)

    ˜1=diag(˜1m1,,˜1ml)

    where

    ˜1mi=(0,,0m1++mi11,,1mi,0,,0mi+1++ml),i=1,,l,

    and define the matrix

    ¯Q=diag(1,,1|IA|,ν1,,νl)((ˆQAAˆQAWˆQWAˆQWW)+(ˆQATˆQWT)˜Q1TT(˜QTA˜QTW))diag(1,,1|IA|,˜1). (B.3)

    Note that ¯Q is not the same as Q0 (which was previously defined for SRNs).

    Remark B.2. The matrices in the product above have different dimensions, i.e. diag(1,,1,ν1,,νl) is a (|IA|+l)×(|IA|+|IW|) matrix, the middle matrix is a (|IA|+|IW|)×(|IA|+|IW|) matrix whereas diag(1,,1,˜1) is a (|IA|+|IW|)×(|IA|+l) matrix.

    Then the zeroth order outer expansion φ0(t) (corresponding to [22,(4.86)]) with initial distribution π0=(pA,pW,pT) is determined by the following ODE,

    {ddt(φA(t),ϑ1(t),,ϑl(t))=(φA(t),ϑ1(t),,ϑl(t))¯Q,φA(0)=pA+pT(˜QTT)1˜QTA,(ϑ1(0),,ϑl(0))=(pW+pT(˜QTT)1˜QTW)˜1, (B.4)

    which defines the generator of a CTMC that is given by ¯Q. This CTMC has reduced state space IA{s1,,sl}, where each state si corresponds to the assembled irreducible component, which has stationary distribution νi from ˜QWW, see (B.2). Then, denoting the projection onto the reduced state space of the initial value of (B.4) by Pr(π0):=(φA(0),ϑ1(0),,ϑl(0)), we get a solution

    (φA(t),ϑ1(t),,ϑl(t))=Pr(π0)e¯Qt.

    Denote the sequence of CTMCs associated to Qϵ by Xϵ(t) and the CTMC from ¯Q by X0(t). Under these conditions the following holds [22,Theorem 4.45].

    Proposition B.3. Let π0 be a probability distribution on IAIWIT with support on IA. Then, for all T>0 and all BIA:

    supt[0,T]|Pπ0(Xϵ(t)B)PPr(π0)(X0(t)B)|=O(ϵ),forϵ0.

    The contribution of the zeroth order outer expansion is sufficient to approximate the above events to order O(ϵ) as the initial-layer corrections are zero [22]. Hence, the CTMC generated by ¯Q can be seen as the limit CTMC (in the sense of Proposition B.3). The above will be the main reference of this section for the proof of Theorem 4.1.

    Remark B.4. In the inhomogeneous case, the generator of the scaled CTMC of [22] has the form Qϵt=1ϵ˜Qt+ˆQt, where ˜Qt,ˆQt satisfy some regularity conditions on t[0,T]. The conclusions we have stated are essentially the same in this case. In particular, the error O(ϵ) for the zeroth order outer expansion is as in Proposition B.3.

    We introduce watched CTMCs on a finite state space. Watched Markov chains appear as a restriction of a bigger Markov chain, which we observe only if the chain is in a specific subset of the state space [38,39]. For a textbook treatment see [38], and for a review and connections to linear algebra we refer to [39].

    In the following, we consider a CTMC (X(t))t0 with generator Q on a finite state space S with E1S a subset of the state space. Writing the Q-matrix in block-form according to the sets E1 and E2=SE1, we decompose Q as

    Q=(QE1E1QE1E2QE2E1QE2E2).

    Lemma B.5. If under

    Q:=(QE1E1QE1E200)

    all states in E1 are transient, then QE1E1 is stable and non-singular.

    Proof. Note that no diagonal entries of QE1E1 are zero. Since all states are transient, then all communicating component of the matrix have a "leak" (the row sum of QE1E1for a least one state in a component is non-zero) and the result follows from [40].

    The censored Markov chain on E2 is defined as the process that has sample paths following the process (X(t))t0 as long as it is in E2, and ignoring the parts where X(t)E1. This so-defined process is then again a Markov process, where under the assumption above, the transition rates of the watched Markov chain are given by

    Qwatched:=QE2E2QE2E1Q1E1E1QE1E2.

    Remark B.6. In the above formula, the second summand catches the contribution from when the original CTMC enters E1 until it returns to E2. By definition of the watched CTMC, we have

    (QE2E1Q1E1E1QE1E2)v,w=xE1qv,xPx(Xτ=w),

    where τ=inf{t0|X(t)E2}.

    The proof consists of several parts. We first show without loss of generality, the following might be assumed:

    the closed set E has a particular form,

    the initial distribution has support on a special part of the state space,

    the RN N is finite.

    Then we apply singularly perturbed CTMCs of § B.1 to the SRN. As a third step, we simplify by showing, the dynamics can be restricted to a smaller part of the state space. Finally, we connect the limit CTMC to watched CTMCs of § B.2, which shows how the reduced reactions are derived from the limit CTMC of the scaled SRNs.

    In preparation for the proof, we consider the following. Let N be an SRN with non-interacting species set U, fast proper reactions FRU, and transition intensities λr,rR. For xZn0, denote by N(x) the set of reachable states from x of N, and for zZp0, denote by NU,F(z) the set of reachable states from z of NU,F.

    We introduce a surrogate model M that reflects the behavior of the SRN of N when ϵ is very small. Specifically, this model is a CTMC on Zn0 determined by the Q-matrix

    Q(x,x+ξ):=r:yyR:yy=ξ˜λr(x),

    with transition intensities,

    ˜λr(x):={λr(x)ifrF,1N0(x)λr(x)ifrRF.

    Similarly, for xZn0, we denote by M(x) the set of reachable states from x of M.

    Let Ni:={x=(z,u)Zn0|mj=1uj=i} for i0. Then, we have the following.

    Proposition B.7. Suppose Assumption 1 holds, that F is proper, and let x=(z,0)N0. Then, ρO(M(x)N0))=NU,F(z) and M(x)N(x)

    For the proof of Theorem 4.1, the following Lemma will be useful. The proof follows by contradiction and is omitted.

    Lemma B.8. Suppose Assumption 1 holds and that F is proper. If Assumption 2 holds for some closed set E, then for any xEN0, the states in M(x)N1 are transient for the CTMC of the SRN obtained by considering the RN defined by the reaction set F (and taking the same transition intensities as in N).

    Consider the family of scaled CTMCs as in the setting of § 4 with notation introduced in § B.3.1. Furthermore, consider a finite closed set EZn0. As E is finite, we can assume wlog that N is finite.

    For any probability distribution π on E and any state wE, Pπ(X(t)=w)=xEπxPx(X(t)=w). Therefore it is enough to prove the statement for:

    a closed set of the form ˜E:=N(x) with xN0E. This follows as E=xEN0N(x) by definition.

    an initial distribution with support on M(x)N0 for x˜EN0. This follows as by definition ˜EN0=(x˜EN0M(x))N0, see Proposition B.7.

    Hence, we might assume that N is finite, that E:=N(x) for some xN0 and that π has support on M(x)N0E for xE. We denote M:=M(x).

    With the preparations above, the matrix ˜Q from § B.1 is comprised of the transition intensities from reactions of F with a non-interacting species of U in the reactant and ˆQ is comprised of the other reactions RF, where we divide the states of E as follows:

    IS1=MN0,IS2=(EN0)M,
    IF1=MN1,IF2=E(N0(MN1)),

    where by definition E=IS1IS2IF1IF2. We next illustrate the possible slow transitions (that is, transitions from ˆQ) outgoing from IS1 in blue on the left and all possible fast transitions (that is, transitions from ˜Q) displayed in red on the right by their transition state diagrams.

    Note that there are other possible slow transitions (that is, transitions from ˆQ) outgoing from IS2,IF1,IF2, but the transitions from IS1 are enough for our considerations.

    The transition diagrams above determine the zero/non-zero parts of the corresponding Q-matrices in block-form as follows.

    Lemma B.9. The following holds for the fast and slow parts of the Q-matrices (notation as in § B.1) divided into blocks:

    ˆQ|IS1×S=(ˆQS1S1ˆQS1S2ˆQS1F1ˆQS1F2)=(ˆQS1S10ˆQS1F10), (B.5)
    ˜Q=(˜QS1S1˜QS1S2˜QS1F1˜QS1F2˜QS2S1˜QS2S2˜QS2F1˜QS2F2˜QF1S1˜QF1S2˜QF1F1˜QF1F2˜QF2S1˜QF2S2˜QF2F1˜QF2F2)=(00000000˜QF1S10˜QF1F10˜QF2S1˜QF2S2˜QF2F1˜QF2F2)

    Proof. Recall that the slow transitions come from RF and the fast ones from F by assumption. We go through the slow transitions and then the fast ones.

    By contradiction there are no slow transitions from IS1 to IS2 or IF2.

    By Assumption 1, ˜Q has zero entries on E0=IS1IS2 (the coordinates of non-interacting species are zero). By the definition of the sets IF1,IS2 and IF2 there are no fast transitions from IF1 to IS2 or IF2.

    Following § B.1 (see also [22]), the distribution of the CTMC converges to the zeroth order outer expansion φ0(t) for t0, which we consider as

    φ0(t):=(φS1(t),φS2(t),φF1(t),φF2(t)).

    By assumption, the initial distribution π0 has support contained in IS1. We next go through the roles of the states in IS1,IS2,IF1,IF2.

    States in IF1 are transient in ˜Q by Assumption 2 and Lemma B.8, and ˜Q|IF1×IF1 is non-singular (see Lemma B.5). States in IS1,IS2 are absorbing in ˜Q. States in IF2 can be transient, absorbing or part of a closed communicating class in ˜Q. Correspondingly, φ0(t) is zero on IF1 and on the transient part of IF2.

    Next, we make the following general observation.

    Remark B.10. Let (X(t))t0 be a CTMC with generator Q on a state space S. Let π0 be a probability distribution with supp(π)ZS, where |Z|<. Assume that there are no xZ,ySZ such that xy. Then, Q|Z×Z is a Q-matrix, and denoting by Xrestr the corresponding CTMC, we have

    Pπ0(X(t)B)=Pπ0(Xrestr(t)B),forBS,t0.

    Then, it is enough to consider the restricted Q-matrix Q|A×A and Xrestr for computations of such probabilities.

    Lemma B.11. The CTMC of the generator ¯Q (see (B.3) of § B.1) in our setting has the property of Remark B.10 with Z:=IS1=MN0.

    Proof. First, we decompose IF2 into three different parts F2,t,F2,w and F2,a according to whether a state is transient, part of a (non-trivial) irreducible component or absorbing in ˜Q. Then, writing

    A=IS1IS2F2,a,T=IF1F2,t,W=F2,w, (B.6)

    corresponds to the situation of § B.1.

    Let S be the reduced state space of ¯Q (cf., § B.1). Clearly AS, as these are the absorbing states in ˜Q, and at least the transient states have been eliminated from S. Hence IS1S. We next focus on ¯Q|IS1×S, which contains the possible outgoing transitions from IS1. It is enough to show that ¯Q|IS1×SIS1 has only zero entries. For this, it suffices to consider the matrices in the middle of (B.3), which we recall here,

    ((ˆQAAˆQAWˆQWAˆQWW)+(ˆQATˆQWT)˜Q1TT(˜QTA˜QTW)). (B.7)

    We next check the right summand of (B.7), which we denote

    R=(ˆQATˆQWT)˜Q1TT(˜QTA˜QTW)=(ˆQAT˜Q1TT˜QTAˆQAT˜Q1TT˜QTWˆQWT˜Q1TT˜QTAˆQWT˜Q1TT˜QTW).

    Then, consider the restriction to R|IS1×S. Keeping the decomposition of A,T,W in (B.6) in mind, the following hold for the involved matrices ˆQAT, ˜Q1TT, ˜QTA, ˜QTW by Lemma B.9:

    ˜Q1TT=(˜QF1F10)1=(˜Q1F1F10),˜QTW=(˜QF1F2,w˜QF2,tF2,w)=(0),
    ˆQAT=(ˆQS1F10),˜QTA=(˜QF1S100),

    and we get that

    R|S1×ST=(RS1S1RS1S2RS1F2,aRS1S2,w)(ˆQS1F1˜Q1F1F1˜QF1S1000) (B.8)

    Hence, the property holds for R, that is, the right hand side of (B.7) has the required property.

    Next we look at the left hand side of (B.7), and again restrict to IS1×ST. By Lemma B.9, the slow part has only non-zero transitions to IS1 or IF1, hence also this matrix has the required property. As this holds for both summands of (B.7), we have shown that the property holds for ¯Q.

    Finally, noting the form of equation (B.5) and equation (B.8) we see that only the slow and fast parts outgoing from IF1 and IS1 contribute to ¯Q|IS1×IS1. Then we can compute ¯Q for the state space IF1IS1 (i.e. with restricted Qϵ,˜Q,ˆQ) or for the state space IS1IS2IF1IF2. However, in both cases the expression we get for ¯Q|IS1×IS1 is the same, hence the following holds.

    Lemma B.12. If the initial distribution has support on IS1, for the computation of φ0(t), it is enough to restrict the state space of Qϵ to IF1IS1.

    We treat the restriction of the state space to IF1IS1 in the following. Restricting to IF1IS1, we have

    ˜Q=(˜QS1S1˜QS1F1˜QF1S1˜QF1F1)=(00˜QF1S1˜QF1F1),ˆQ=(ˆQS1S1ˆQS1F1ˆQF1S1ˆQF1F1).

    where the states IS1 are absorbing in ˜Q, and the states of IF1 are transient states. Note that this setting corresponds to the situation of § B.1 where we removed the irreducible part (that is, IW in the notation of § B.1), IF1 corresponds to T and IS1 corresponds to A of § B.1.

    The scaled CTMC converges to the zeroth order outer expansion φ0(t) for t>0, giving φ0(t):=(φS1(t),φF1(t))=(φS1(t),0F1) which satisfies the ODE from § B.1 with

    {˙φS1(t)=φS1(t)(ˆQS1S1+ˆQS1F1(˜QF1F1)1˜QF1S1),φS1(0)=pS1+pF1(˜QF1F1)1˜QF1S1=Pr(π0), (B.9)

    with initial distribution π0=(pS1,pF1). As the initial distribution has support on IS1, we get the following from Proposition B.3, where again Xϵ(t) is the CTMC associated to Qϵ while X0(t) is the CTMC from C (corresponding to ¯Q from § B.1).

    Lemma B.13. Let π0 be a probability distribution with support on IS1, and BIS1. Then the following holds for all T>0:

    supt[0,T]|Pπ0(Xϵ(t)B)PPr(π0)(X0(t)B)|=O(ϵ)forϵ0.

    Finally, observing the form of the matrix ˆQS1S1+ˆQS1F1(˜QF1F1)1˜QF1S1 in (B.9), we can view it as the watched Markov chain of § B.2, where it is watched when in IS1 with Q-matrix

    Q=(ˆQS1S1ˆQS1F1˜QF1S1˜QF1F1).

    By the interpretation of the censored Markov chain available from Remark B.6, the part ˆQS1F1(˜QF1F1)1˜QF1S1 corresponds to transition intensities given by the rates entering a state in IF1 from IS1 times the exit probabilities to IS1. As ˜QF1S1,˜QF1F1 have the transition intensities from the reactions F and ˆQS1F1 from RF, this corresponds exactly to the transition rates of the defined reduced RN of § 3.2. Hence Theorem 4.1 follows.

    The proof follows the same steps as the proof of Theorem 4.1, hence it is enough to check that Assumptions 2 and 3 together with F being proper are sufficient for the assumptions made in [22] to hold. Recall that since the λyy(t,) satisfy Assumption 1 for each t[0,T], the state space decomposition of the ordinary Markov process for an RN and Qt for t fix agree. Hence, by Assumption 2 and Lemma B.8, any xE with one molecule of a non-interacting species is transient in ˆQt. Furthermore by Assumption 3, ˜Qt,ˆQt are once continuous differentiable with Lipschitz derivative ([22,Assumption A4.4(respectively,A4.5)]). With these observations we can conclude.

    LH acknowledges funding from the Swiss National Science Foundations Early Postdoc.Mobility grant (P2FRP2_188023). The work presented in this article is supported by Novo Nordisk Foundation, grant NNF19OC0058354.

    The authors declare there are no conflicts of interest.



    [1] J. D. Murray, Mathematical Biology I. An Introduction, volume 17 of Interdisciplinary Applied Mathematics, Springer, New York, 3 edition, (2002). doi: 10.1007/b9886
    [2] E. Weinan, Principles of Multiscale Modeling, Cambridge University Press, (2011).
    [3] L. Segal, M. Slemrod, The quasi-steady-state assumption: A case study in perturbation, SIAM Rev., 31 (1989), 446–477. doi: 10.1137/1031091 doi: 10.1137/1031091
    [4] N. Fenichel, Geometric singular perturbation theory for ordinary differential equations, J. Diff. Eqns., 31 (1979), 53–98. doi: 10.1016/0022-0396(79)90152-9 doi: 10.1016/0022-0396(79)90152-9
    [5] A. N. Tikhonov, Systems of differential equations containing a small parameter multiplying the derivative (in Russian), Math. Sb., 31 (1952), 575–586.
    [6] A. Goeke, S. Walcher, E. Zerz, Classical quasi-steady state reduction – a mathematical characterization, Phys. D Nonlinear Phenom., 345 (2017), 11–26. doi: 10.1016/j.physd.2016.12.002 doi: 10.1016/j.physd.2016.12.002
    [7] F. G. Heineken, H. M. Tsuchiya, R. Aris, On the mathematical status of the pseudo-steady state hypothesis of biochemical kinetics, Math. Biosci.., 1 (1967), 95–113.
    [8] H.-W. Kang, W. R. KhudaBukhsh, H. Koeppl, G. A. Rempała, Quasi-steady-state approximations derived from the stochastic model of enzyme kinetics, Bull. Math. Biol., 81 (2019), 1303–1336. doi: 10.1007/s11538-019-00574-4 doi: 10.1007/s11538-019-00574-4
    [9] E. Feliu, C. Wiuf, Variable elimination in chemical reaction networks with mass-action kinetics, SIAM J. Appl. Math., 72 (2012), 959–981. doi: 10.1137/110847305 doi: 10.1137/110847305
    [10] E. Feliu, S. Walcher, C. Wiuf, Quasi-steady state and singular perturbation reduction for reaction networks with non-interacting species, SIAM J. Appl. Dyn. Syst., In press.
    [11] A. Gorban, Model reduction in chemical dynamics: Slow invariant manifolds, singular perturbations, thermodynamic estimates, and analysis of reaction graph, Curr. Opin. Chem. Eng., 21 (2018), 48–59. doi: 10.1016/j.coche.2018.02.009 doi: 10.1016/j.coche.2018.02.009
    [12] X. Kan, Chang Hyeong Lee, H. G. Othmer, A multi-time-scale analysis of chemical reaction networks: Ii. stochastic systems, J. Math. Biol., 73 (2016), 1081–1129. doi: 10.1007/s00285-016-0980-x doi: 10.1007/s00285-016-0980-x
    [13] D. Schnoerr, G. Sanguinetti, R. Grima, Approximation and inference methods for stochastic biochemical kinetics—a tutorial review, J. Phys. A Math., 50 (2017), 093001. doi: 10.1088/1751-8121/aa54d9 doi: 10.1088/1751-8121/aa54d9
    [14] J. A. M. Janssen, The elimination of fast variables in complex chemical reactions. ii. mesoscopic level (reducible case), J. Stat. Phys., 57 (1989), 171–185. doi: 10.1007/BF01023639 doi: 10.1007/BF01023639
    [15] J. A. M. Janssen, The elimination of fast variables in complex chemical reactions. iii. mesoscopic level (irreducible case), J. Stat. Phys., 57 (1989), 187–198. doi: 10.1007/BF01023640 doi: 10.1007/BF01023640
    [16] T. G. Kurtz, Approximation of Population Processes, Society for Industrial and Applied Mathematics, (1981). doi: 10.1137/1.9781611970333
    [17] K. Ball, T. G. Kurtz, L. Popovic, G. Rempala, Asymptotic analysis of multiscale approximations to reaction networks, Ann. Appl. Probab., 16 (2006), 1925–1961. doi: 10.1214/105051606000000420 doi: 10.1214/105051606000000420
    [18] H.-W. Kang, T. G. Kurtz, Separation of time-scales and model reduction for stochastic reaction networks, Ann. Appl. Probab., 23 (2013), 529–583. doi: 10.1214/12-AAP841 doi: 10.1214/12-AAP841
    [19] P. Pfaffelhuber, L. Popovic, Scaling limits of spatial compartment models for chemical reaction networks, Ann. Appl. Probab., 25 (2015), 3162–3208. doi: 10.1214/14-AAP1070 doi: 10.1214/14-AAP1070
    [20] D. Cappelletti, C. Wiuf, Elimination of intermediate species in multiscale stochastic reaction networks, Ann. Appl. Probab., 26 (2016), 2915–2958. doi: 10.1214/15-AAP1166 doi: 10.1214/15-AAP1166
    [21] M. Sáez, C. Wiuf, E. Feliu, Graphical reduction of reaction networks by linear elimination of species, J. Math. Biol., 74 (2017), 195–237. doi: 10.1007/s00285-016-1028-y doi: 10.1007/s00285-016-1028-y
    [22] G.G. Yin, Q. Zhang, Continuous-Time Markov Chains and Applications: A Two-Time-Scale Approach, Stochastic Modelling and Applied Probability, Springer New York, (2012). doi: 10.1007/978-1-4614-4346-9
    [23] D. Freedman, Approximating Countable Markov Chains, Springer New York, (2012).
    [24] X. Chen, C. Jia, Limit theorems for generalized density-dependent Markov chains and bursty stochastic gene regulatory networks, J. Math. Biol., 80 (2020), 959–994. doi: 10.1007/s00285-019-01445-1 doi: 10.1007/s00285-019-01445-1
    [25] S. Be'er, M. Assaf, Rare events in stochastic populations under bursty reproduction, J. Stat. Mech–Theory E., (2016), 113501. doi: 10.1088/1742-5468/2016/11/113501 doi: 10.1088/1742-5468/2016/11/113501
    [26] B. Ingalls, Mathematical Modeling in Systems Biology, Cambridge, Massachusetts: MIT Press, (2013).
    [27] D. F. Anderson, T. G. Kurtz, Stochastic Analysis of Biochemical Systems, Springer Publishing Company, Incorporated, (2015). doi: 10.1007/978-3-319-16895-1
    [28] J. R. Norris, Markov Chains, Cambridge University Press, Cambridge, (1997). doi: 10.1017/CBO9780511810633
    [29] L. Hoessly, C. Wiuf, P. Xia, On the sum of chemical reactions, (2021). arXiv: 2105.04353.
    [30] D. F. Anderson, A modified next reaction method for simulating chemical systems with time dependent propensities and delays, J. Chem. Phys., 127 (2007), 214107. doi: 10.1063/1.2799998 doi: 10.1063/1.2799998
    [31] A. Cornish-Bowden, Fundamentals of Enzyme Kinetics, Wiley, (2013).
    [32] C. Jia, Reduction of markov chains with two-time-scale state transitions, Stochastics, 88 (2016), 73–105. doi: 10.1080/17442508.2015.1036433 doi: 10.1080/17442508.2015.1036433
    [33] G. Yin, Q. Zhang, G. Badowski, Asymptotic properties of a singularly perturbed markov chain with inclusion of transient states, Ann. Appl. Probab., 10 (2000), 549–572.
    [34] A. Jakubowski, A non-Skorohod topology on the Skorohod space, Electron. J. Probab., 2 (1997), 1–21. doi: 10.1214/EJP.v2-18 doi: 10.1214/EJP.v2-18
    [35] C. Xu, M. C. Hansen, C. Wiuf, Dynamics of continuous time markov chains with applications to stochastic reaction networks, (2019). arXiv: 1909.12825.
    [36] T. Kurtz, The relationship between stochastic and deterministic models for chemical reactions, J. Chem. Phys.., 57 (1972), 2976–2978. doi: 10.1063/1.1678692 doi: 10.1063/1.1678692
    [37] G. Pavliotis, A. Stuart, Multiscale Methods: Averaging and Homogenization, volume 53. Springer, 01 (2008). doi: 10.1007/978-0-387-73829-1
    [38] J. G. Kemeny, J. L. Snell, Finite Markov Chains: With a new appendix "Generalization of a Fundamental Matrix". Undergraduate Texts in Mathematics, Springer New York, (1983).
    [39] C. Meyer, Stochastic complementation, uncoupling markov chains, and the theory of nearly reducible systems, SIAM Rev.., 31 (1995), 09. doi: 10.1137/1031050 doi: 10.1137/1031050
    [40] R. J. Plemmons, M-matrix characterizations.i—nonsingular m-matrices, Linear Algebra Its Appl., 18 (1977), 175–188. doi: 10.1016/0024-3795(77)90073-8 doi: 10.1016/0024-3795(77)90073-8
  • This article has been cited by:

    1. LINARD HOESSLY, CARSTEN WIUF, PANQIU XIA, On the sum of chemical reactions, 2023, 34, 0956-7925, 303, 10.1017/S0956792522000146
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2383) PDF downloads(77) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog