Skip to main content
Log in

A multi-time-scale analysis of chemical reaction networks: II. Stochastic systems

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

We consider stochastic descriptions of chemical reaction networks in which there are both fast and slow reactions, and for which the time scales are widely separated. We develop a computational algorithm that produces the generator of the full chemical master equation for arbitrary systems, and show how to obtain a reduced equation that governs the evolution on the slow time scale. This is done by applying a state space decomposition to the full equation that leads to the reduced dynamics in terms of certain projections and the invariant distributions of the fast system. The rates or propensities of the reduced system are shown to be the rates of the slow reactions conditioned on the expectations of fast steps. We also show that the generator of the reduced system is a Markov generator, and we present an efficient stochastic simulation algorithm for the slow time scale dynamics. We illustrate the numerical accuracy of the approximation by simulating several examples. Graph-theoretic techniques are used throughout to describe the structure of the reaction network and the state-space transitions accessible under the dynamics.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. The notation used is defined later.

  2. This form also includes non-ideal mass action rate laws, but the concentrations in (2.7) are then replaced by the activities of the species in the reactant complex, and as a result the flow on a edge may depend on all species in the system.

  3. This formulation applies only to ideal solutions—in nonideal solutions the number of molecules must be replaced by an appropriate measure of its activity in the solution. In particular, this involves a suitable description of diffusion when the solution is not ideal (Othmer 1976; Schnell and Turner 2004).

  4. The case in which there are infinitely many states will be discussed in Sect. 3.2.

  5. A vector space whose elements are infinite sequences of real numbers.

  6. This is the direct sum, but is generally not the orthogonal direct sum.

  7. The reader can show that the presence of inputs or outputs of the form given in Table 1 does not alter the deficiency.

References

  • Anderson DF, Craciun G, Kurtz TG (2010) Product-form stationary distributions for deficiency zero chemical reaction networks. Bull Math Biol 72:1947–1970

    Article  MathSciNet  MATH  Google Scholar 

  • Anderson DF, Kurtz TG (2011) Continuous time markov chain models for chemical reaction networks. In: Koeppl H, Densmore D, Setti G, di Bernardo M (eds) Design and analysis of biomolecular circuits: engineering approaches to systems and synthetic biology. Springer, New York, pp 3–42

  • Aris R (1965) Prolegomena to the rational analysis of chemical reactions. Arch Ration Mech Anal 19(2):81–99

    Article  MathSciNet  Google Scholar 

  • Boucherie RJ, Dijk NMV (1991) Product forms for queueing networks with state-dependent multiple job transitions. Adv Appl Probab 23(1):152–187

  • Bundschuh R, Hayot F, Jayaprakash C (2003) Fluctuations and slow variables in genetic networks. Biophy J 84:1606

    Article  Google Scholar 

  • Campbell SL, Meyer CP (1991) Generalized inverses of linear transformations. Dover Publications, New York

  • Cao Y, Gillespie DT, Petzold LR (2005) The slow-scale stochastic simulation algorithm. J Chem Phys 122:014116

    Article  Google Scholar 

  • Chen WK (1971) Applied graph theory. North-Holland, Amsterdam

    MATH  Google Scholar 

  • Chevalier MW, EI-Samad H (2009) A rigorous framework for multiscale simulation of stochastic cellular networks. J Chem Phys 131(5):054102

    Article  Google Scholar 

  • Cotter S (2015) Constrained approximation of effective generators for multiscale stochastic reaction networks and application to conditioned path sampling. arXiv:1506.02446 (arXiv preprint)

  • Deuflhard P, Huisinga W, Jahnke T, Wulkow M (2008) Adaptive discrete Galerkin methods applied to the chemical master equation. SIAM J Sci Comput 30(6):2990–3011

    Article  MathSciNet  MATH  Google Scholar 

  • E W, Liu D, Vanden-Eijnden E (2005) Nested stochastic simulation algorithm for chemical kinetic systems with disparate rates. J Chem Phys 123:194107

  • Gadgil C, Lee CH, Othmer HG (2005) A stochastic analysis of first-order reaction networks. Bull Math Biol 67:901–946

    Article  MathSciNet  MATH  Google Scholar 

  • Gillespie DT (2007) Stochastic simulation of chemical kinetics. Annu Rev Phys Chem 58:35–55

    Article  Google Scholar 

  • Goutsias J (2005) Quasiequilibrium approximation of fast reaction kinetics in stochastic biochemical systems. J Chem Phys 122:184102

    Article  Google Scholar 

  • Goutsias J, Jenkinson G (2013) Markovian dynamics on complex reaction networks. Phys Rep 529:199–264

    Article  MathSciNet  Google Scholar 

  • Haseltine EL, Rawlings JB (2002) Approximate simulation of coupled fast and slow reactions for stochastic chemical kinetics. J Chem Phys 117(15):6959–6969

    Article  Google Scholar 

  • Hellander A, Lötstedt P (2007) Hybrid method for the chemical master equation. J Comput Phys 227:100–122

    Article  MATH  Google Scholar 

  • Horn F, Jackson R (1972) General mass action kinetics. Arch Ration Mech Anal 48:81

    MathSciNet  Google Scholar 

  • Hu J, Kang H-W, Othmer HG (2013) Stochastic analysis of reaction–diffusion processes. Bull Math Biol 76:854–894

  • Huang C, Liu D (2014) Strong convergence and speed up of nested stochastic simulation algorithm. Commun Comput Phys 15(4):1207–1236

    Article  MathSciNet  Google Scholar 

  • Jahnke T, Huisinga W (2007) Solving the chemical master equation for monomolecular reaction systems analytically. J Math Biol 54:1–26

    Article  MathSciNet  MATH  Google Scholar 

  • Janssen JAM (1989a) The elimination of fast variables in complex chemical reactions II. J Stat Phys 57:171–186

    Article  Google Scholar 

  • Janssen JAM (1989b) The elimination of fast variables in complex chemical reactions. III. Mesoscopic level (irreducible case). J Stat Phys 57:187–198

    Article  MathSciNet  Google Scholar 

  • Kato T (1966) Perturbation theory for linear operators. Springer-Verlag, Berlin

    Book  MATH  Google Scholar 

  • Kazeev V, Khammash M, Nip M, Schwab C (2014) Direct solution of the chemical master equation using quantized tensor trains. PLoS Comput Biol 10(3):e1003359

  • Kim JK, Josic K, Bennett MR (2014) The validity of quasi-steady-state approximations in discrete stochastic simulations. Biophys J 107:783–793

    Article  Google Scholar 

  • Lee CH, Lui R (2009) A reduction method for multiple time scale stochastic reaction networks. J Math Chem 46:1292–1321

    Article  MathSciNet  MATH  Google Scholar 

  • Lee CH, Othmer HG (2009) A multi-time-scale analysis of chemical reaction networks: I. Deterministic systems. J Math Biol 60(3):387–450

    Article  MathSciNet  MATH  Google Scholar 

  • Mastny EA, Haseltine EL, Rawlings JB (2007) Two classes of quasi-steady-state model reductions for stochastic kinetics. J Chem Phys 127(9):094106

    Article  Google Scholar 

  • McQuarrie DA (1967) Stochastic approach to chemical kinetics. J Appl Probab 4(3):413–478

    Article  MathSciNet  MATH  Google Scholar 

  • Mélykúti B, Hespanha JP, Khammash M (2014) Equilibrium distributions of simple biochemical reaction systems for time-scale separation in stochastic reaction networks. J R Soc Interface 11(97):20140054

    Article  Google Scholar 

  • Menz S, Latorre JC, Schutte C, Huisinga W (2012) Hybrid stochastic-deterministic solution of the chemical master equation. Multiscale Model Simul 10(4):1232–1262

    Article  MathSciNet  MATH  Google Scholar 

  • Norris JR (1998) Markov chains. Cambridge University Press, Cambridge

  • Othmer HG (1976) Nonuniqueness of equilibria in closed reacting systems. Chemical Eng Sci 31:993–1003

    Article  Google Scholar 

  • Othmer HG (1979) A graph-theoretic analysis of chemical reaction networks Lecture Notes, Rutgers University. http://math.umn.edu/~othmer/graphrt.pdf

  • Othmer HG (2005) Analysis of complex reaction networks Lecture Notes, University of Minnesota

  • Othmer HG, Aldridge JA (1978) The effects of cell density and metabolite flux on cellular dynamics. J Math Biol 5:169–200

    Article  MathSciNet  MATH  Google Scholar 

  • Peles S, Munsky B, Khammash M (2006) Reduction and solution of the chemical master equation using time scale separation and finite state projection. J Chem Phys 125:204104

    Article  Google Scholar 

  • Rao CV, Arkin AP (2003) Stochastic chemical kinetics and the quasi-steady state assumption: application to the Gillespie algorithm. J Chem Phys 118(11):4999–5010

    Article  Google Scholar 

  • Salis H, Kaznessis Y (2005) Accurate hybrid stochastic simulation of a system of coupled chemical or biochemical reactions. J Chem Phys 122:054103

    Article  Google Scholar 

  • Schnell S, Turner T (2004) Reaction kinetics in intracellular environments with macromolecular crowding: simulations and rate laws. Prog Biophys Mol Biol 85(2):235–260

    Article  Google Scholar 

  • Smith S, Cianci C, Grima R (2015) Model reduction for stochastic chemical systems with abundant species. arXiv:1510.03172 (arXiv preprint)

  • Srivastava R, Haseltine EL, Mastny E, Rawlings JB (2011) The stochastic quasi-steady-state assumption: reducing the model but not the noise. J Chem Phys 134(15):154109

    Article  Google Scholar 

  • Thomas P, Straube AV, Grima R (2011) Communication: limitations of the stochastic quasi-steady-state approximation in open biochemical reaction networks. J Chem Phys 135(18):181103

    Article  Google Scholar 

  • Wilhelm T (2009) The smallest chemical reaction system with bistability. BMC Syst Biol 3(1):90

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xingye Kan.

Additional information

All authors contributed equally to this work.

Supported in part by NSF Grants DMS # 9517884 and 131974 and NIH Grant # GM 29123 to H. G. Othmer and by National Research Foundation of Korea (2014R1A1A2054976) to C. H. Lee.

Appendix

Appendix

1.1 The general case in Example 10

When there are \(N_0\) molecules, there are \(N_0+1\) fast components and the transition matrix K is block tridiagonal, i.e.

$$\begin{aligned} K = \left[ \begin{array}{cccccc} \dfrac{1}{\epsilon }K_1^f+K_1^s &{} K_{1,2}^s &{} &{} &{} &{}\\ K_{2,1}^s &{} \dfrac{1}{\epsilon }K_2^f+K_2^s &{} K_{2,3}^s &{} &{} &{}\\ &{} K_{3,2}^s &{} \ddots &{} \ddots &{} &{} \\ &{} &{} \ddots &{} \ddots &{} \ddots &{} \\ &{} &{} &{} \ddots &{} \ddots &{} K_{N_0,N_0+1}^s\\ &{} &{} &{} &{} K_{N_0+1,N_0}^s &{} \dfrac{1}{\epsilon }K_{N_0+1}^f+K_{N_0+1}^s \end{array} \right] , \end{aligned}$$

where the fast blocks are given by

$$\begin{aligned} K_1^f= & {} \left[ \begin{array}{cccccc} -N_0k_1 &{} k_2 &{} &{} &{} &{}\\ N_0k_1 &{} -[(N_0-1)k_1+k_2] &{} 2k_2 &{} &{} &{}\\ &{} (N_0-1)k_1 &{} -[(N_0-2)k_1+2k_2] &{} 3k_2 &{} &{} \\ &{} &{} \ddots &{} \ddots &{} \ddots &{} \\ &{} &{} &{} \ddots &{} \ddots &{} N_0k_2\\ &{} &{} &{} &{} k_1 &{} -N_0k_2 \end{array} \right] _ {(N_0+1)\times (N_0+1)}\\ K_2^f= & {} \left[ \begin{array}{cccccc} -(N_0-1)k_1 &{} k_2 &{} &{} &{} &{}\\ (N_0-1)k_1 &{} -[(N_0-2)k_1+k_2] &{} 2k_2 &{} &{} &{}\\ &{} (N_0-2)k_1 &{} -[(N_0-3)k_1+2k_2] &{} 3k_2 &{} &{} \\ &{} &{} \ddots &{} \ddots &{} \ddots &{} \\ &{} &{} &{} \ddots &{} \ddots &{} (N_0-1)k_2\\ &{} &{} &{} &{} k_1 &{} -(N_0-1)k_2 \end{array} \right] _ {N_0\times N_0} \cdots \\ K_{N_0}^f= & {} \left[ \begin{array}{cc} -k_1 &{} k_2\\ k_1 &{} -k_2 \end{array} \right] , K_{N_{0}+1}^f = 0. \end{aligned}$$

The slow off-diagonal blocks are given by

$$\begin{aligned}&K_{1,2}^s=\left[ \begin{array}{cccc} k_5 &{} &{} &{} \\ k_4 &{} k_5 &{} &{} \\ &{} \ddots &{} \ddots &{} \\ &{} &{} \ddots &{} k_5\\ &{} &{} &{} k_4 \end{array}\right] _{(N_0+1)\times N_0}, K_{2,3}^s=\left[ \begin{array}{cccc} 2k_5 &{} &{} &{} \\ 2k_4 &{} 2k_5 &{} &{} \\ &{} \ddots &{} \ddots &{} \\ &{} &{} \ddots &{} 2k_5\\ &{} &{} &{} 2k_4 \end{array}\right] _{N_0\times (N_0-1)}, \cdots ,\\&K_{N_{0},N_{0}+1}^s = \left[ \begin{array}{c}N_0k_5\\ N_0k_4 \end{array}\right] . \end{aligned}$$

For the lower diagonal blocks,

$$\begin{aligned} K_{2,1}^s= & {} \left[ \begin{array}{ccccc} N_0k_6 &{} k_3 &{} &{} &{} \\ &{} (N_0-1)k_6 &{} 2k_3 &{} &{}\\ &{} &{} \ddots &{} \ddots &{}\\ &{} &{} &{} k_6 &{} N_0k_3 \end{array}\right] _{N_0\times (N_0+1)} \\ K_{3,2}^s= & {} \left[ \begin{array}{ccccc} (N_0-1)k_6 &{} k_3 &{} &{} &{} \\ &{} (N_0-2)k_6 &{} 2k_3 &{} &{}\\ &{} &{} \ddots &{} \ddots &{}\\ &{} &{} &{} k_6 &{} (N_0-1)k_3 \end{array}\right] _{(N_0-1)\times N_0}, \end{aligned}$$

\(\cdots , \qquad K_{N_0+1,N_0}^s = \left[ \begin{array}{cc}k_6&k_3 \end{array}\right] \). Finally, the \(K_i^s\) along the diagonal are diagonal matrices of the same dimension as \(K_i^f\), and the (jj)-th entry of \(K_i^s\) is the negative sum of rates leaving j-th node of the i-th fast component.

In this case the transition rate is given by

$$\begin{aligned} \tilde{k}_{i,j}^s=L_iK_{i,j}^s\varPi _j. \end{aligned}$$

1.2 Moment equations of the invariant distributions

The low-order moments of the distributions for the fast systems play a role in the QSS reduction in Sect. 4, and here we consider the low-order moment equations.

Theorem 20

Let r be the total number of the reactions in the system. Then the invariant (steady-state) distribution of P(nt), which we denote P(n), satisfies

$$\begin{aligned} \sum _{\ell =1}^{r} \mathcal{R}_{\ell }(n-\nu \mathcal{E}_{(\ell )} P(n-\nu \mathcal{E}_{(\ell )}) = \sum _{\ell =1}^{r} \mathcal{R}_{\ell }(n) P(n) \end{aligned}$$
(8.1)

and the first two moment equations lead to

$$\begin{aligned}&E[ \nu \mathcal{E}\mathcal{R}(n)]=0, \end{aligned}$$
(8.2)
$$\begin{aligned}&\sum _{i=1}^{r}\big [ \nu \mathcal{E}_{(i)} \otimes E[n \mathcal{R}_i(n)] + E[n \mathcal{R}_i(n)] \otimes \nu \mathcal{E}_{(i)} + \nu \mathcal{E}_{(i)} \otimes \nu \mathcal{E}_{(i)} E[ \mathcal{R}_i(n)] \big ] = 0. \nonumber \\ \end{aligned}$$
(8.3)

Proof

At the steady-state

$$\begin{aligned} \sum _{\ell =1}^{r} \mathcal{R}_{\ell }(n-\nu \mathcal{E}_{(\ell )}) P(n-\nu \mathcal{E}_{(\ell )}) = \sum _{\ell =1}^{r} \mathcal{R}_{\ell }(n) P(n) \end{aligned}$$
(8.4)

By multiplying by n and summing over all the values of \(n\in \mathcal{L}(n_0)\), we obtain

$$\begin{aligned} \sum _{n}\sum _{\ell =1}^{r} n \mathcal{R}_{\ell }(n-\nu \mathcal{E}_{(\ell )}) P(n-\nu \mathcal{E}_{(\ell )}) = \sum _{n}\sum _{\ell =1}^{r} n \mathcal{R}_{\ell }(n) P(n). \end{aligned}$$

Using the transformation \( n-\nu \mathcal{E}_{(\ell )} \rightarrow n\) on the left side, we obtain

$$\begin{aligned} \sum _{n}\sum _{\ell =1}^{r} (n+\nu \mathcal{E}_{(\ell )}) \mathcal{R}_{\ell }(n) P(n) = \sum _{n}\sum _{\ell =1}^{r} n \mathcal{R}_{\ell }(n) P(n). \end{aligned}$$

By subtracting the right side from the left one,

$$\begin{aligned} \sum _{\ell =1}^{r} \nu \mathcal{E}_{(\ell )} \sum _{n} \mathcal{R}_{\ell }(n) P(n) = \sum _{\ell =1}^{r} \nu \mathcal{E}_{(\ell )} E[\mathcal{R}_{\ell }(n)]= 0. \end{aligned}$$

Thus we conclude that

$$\begin{aligned} \nu \mathcal{E}E[ \mathcal{R}(n)]= 0. \end{aligned}$$

If the deficiency \(\delta \equiv \rho (\mathcal{E}) - \rho (\nu \mathcal{E})\) is zero, then \(E[ \mathcal{R}(n)]\) is a cycle in the graph (Othmer 1979).Footnote 7

At the next order we multiply Eq. (8.4) by a tensor product \(n\otimes n\) and sum over n. Then by a similar argument, we obtain

$$\begin{aligned}&\sum _{i=1}^{r}\big [ \nu \mathcal{E}_{(i)} \otimes \sum _{n} n \mathcal{R}_i(n) p(n) + \sum _{n} n \mathcal{R}_i(n) p(n)\otimes \nu \mathcal{E}_{(i)}\\&\qquad +\, \nu \mathcal{E}_{(i)} \otimes \nu \mathcal{E}_{(i)} \sum _{n} \mathcal{R}_i(n) p(n) \big ]\\&\quad = \sum _{i=1}^{r}\big [ \nu \mathcal{E}_{(i)} \otimes E[n \mathcal{R}_i(n)] + E[n \mathcal{R}_i(n)] \otimes \nu \mathcal{E}_{(i)} + \nu \mathcal{E}_{(i)} \otimes \nu \mathcal{E}_{(i)} E[ \mathcal{R}_i(n)] \big ]\\&\quad = 0. \end{aligned}$$

If \(\delta =0\), the two lowest moment equations can be simplified to

$$\begin{aligned}&E[ \mathcal{R}(n)]=0 \end{aligned}$$
(8.5)
$$\begin{aligned}&\quad \sum _{i=1}^{r}\big [ \nu \mathcal{E}_{(i)} \otimes E[n \mathcal{R}_i(n)] + E[n \mathcal{R}_i(n)] \otimes \nu \mathcal{E}_{(i)} \big ] = 0. \end{aligned}$$
(8.6)

When all reactions are linear the problem is much simpler, and the evolution equations for the first and second moments can be written explicitly in terms of those moments (Gadgil et al. 2005).

As a consequence of Theorem 20, similar equations can be obtained for the quasi-steady-state of the probability distribution for the fast subsystem in a two-time scale stochastic network. We first define the expectation of a function f(n) over a discrete reaction simplex \(\mathcal{L}_f\) for the fast subsystem as follows.

$$\begin{aligned} E_{\mathcal{L}_f}[f(n)] \equiv \sum _{n\in \mathcal{L}_f} f(n) p(n). \end{aligned}$$

Corollary 21

Let \(r_f\) be the total number of fast reactions. Then at the steady-state of the fast subsystem, the governing equation is given by

$$\begin{aligned} \sum _{i=1}^{r_f} \mathcal{R}^f_i(n-\nu \mathcal{E}^f_{(i)}) P(n-\nu \mathcal{E}^f_{(i)}) = \sum _{i=1}^{r_f} \mathcal{R}^f_i(n) P(n) \end{aligned}$$
(8.7)

and for each discrete reaction simplex \(\mathcal{L}_f\),

$$\begin{aligned}&E_{\mathcal{L}_f} [ \mathcal{R}^f(n)]=0, \end{aligned}$$
(8.8)
$$\begin{aligned}&\sum _{i=1}^{r_f}\big [ \nu \mathcal{E}^f_{(i)} \otimes E_{\mathcal{L}_f}[n \mathcal{R}^f_{i}(n)] + E_{\mathcal{L}_f}[n \mathcal{R}^f_i(n)] \otimes \nu \mathcal{E}^f_{(i)} \big ] = 0. \end{aligned}$$
(8.9)

Proof

The ‘state-wise’ form of the master Eq. (3.1) can be written

$$\begin{aligned} \dfrac{d}{dt}P(n,t)= & {} \sum _{\ell }\dfrac{1}{{\epsilon }} \big [\mathcal{R}^f_{\ell }(n-\nu \mathcal{E}^f_{(\ell )})\cdot P(n-\nu \mathcal{E}^f_{(\ell )},t)-\mathcal{R}^f_{\ell }(n)\cdot P(n,t)\big ] \nonumber \\ \nonumber \\&+ \sum _{k} \big [\mathcal{R}^s_{k}(n-\nu \mathcal{E}^s_{(k)})\cdot P(n-\nu \mathcal{E}^s_{(k)},t)- \mathcal{R}^s_{k}(n)\cdot P(n,t)\big ], \end{aligned}$$
(8.10)

where \(\mathcal{R}^f\) and \(\mathcal{R}^s\) are the transition rates of fast and slow reactions, respectively and \(\mathcal{E}^f\) and \(\mathcal{E}^s\) are incidence matrices for fast and slow reactions, respectively.

In the previous theorem, substitute \(\mathcal{E}^f, \mathcal{R}^f\) and \(E_{\mathcal{L}_f}[\cdot ]\) into \(\mathcal{E}, \mathcal{R}\) and \(E[\cdot ]\) and use the full rank assumption on \( \nu \mathcal{E}^f\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kan, X., Lee, C.H. & Othmer, H.G. A multi-time-scale analysis of chemical reaction networks: II. Stochastic systems. J. Math. Biol. 73, 1081–1129 (2016). https://doi.org/10.1007/s00285-016-0980-x

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-016-0980-x

Keywords

Mathematics Subject Classification

Navigation