Skip to main content
Log in

On the Validity of the Stochastic Quasi-Steady-State Approximation in Open Enzyme Catalyzed Reactions: Timescale Separation or Singular Perturbation?

  • Perspectives
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

The quasi-steady-state approximation is widely used to develop simplified deterministic or stochastic models of enzyme catalyzed reactions. In deterministic models, the quasi-steady-state approximation can be mathematically justified from singular perturbation theory. For several closed enzymatic reactions, the homologous extension of the quasi-steady-state approximation to the stochastic regime, known as the stochastic quasi-steady-state approximation, has been shown to be accurate under the analogous conditions that permit the quasi-steady-state reduction in the deterministic counterpart. However, it was recently demonstrated that the extension of the stochastic quasi-steady-state approximation to an open Michaelis–Menten reaction mechanism is only valid under a condition that is far more restrictive than the qualifier that ensures the validity of its corresponding deterministic quasi-steady-state approximation. In this paper, we suggest a possible explanation for this discrepancy from the lens of geometric singular perturbation theory. In so doing, we illustrate a misconception in the application of the quasi-steady-state approximation: timescale separation does not imply singular perturbation.

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

Similar content being viewed by others

Notes

  1. The use of “same” here is slightly abusive; deterministic and stochastic rate constants differ in terms of their units, and the state space is discrete–as opposed to continuous–in the realm of the CME.

  2. From this point onward, we often will refer to the curve defined by (6) as the QSS manifold.

  3. Locally, near a specific value of s, the metric \(e_T/(K_M+s)\ll 1\) is often sufficient to validate the sQSSA. For more details, please consult (Segel 1988; Segel and Slemrod 1989; Eilertsen and Schnell 2020).

  4. \(N_z\) is the range, \({\mathcal {R}}\), of the Jacobian, Dw(z), for \(z\in M\).

  5. For a review describing the different forms of the QSSA, we invite the reader to consult (Schnell and Maini 2003; Eilertsen and Schnell 2020)

  6. In fact, one can rigorously make this claim by analyzing the Lie derivative and demanding that the vector field be nearly tangent to the QSS manifold; see (Schauer and Heinrich 1979; Noethen and Walcher 2011; Goeke et al. 2017; Eilertsen et al. 2021).

  7. We say “almost” because a solution would require the QSS manifold to be perfectly invariant.

  8. The “\(\delta \)” is included in the definition on \({\mathcal {M}}\) so that the compactness requirement of Fenichel theory is satisfied.

  9. Fenichel theory (Fenichel 1979) ensures the eventual validity of (7) as \(e_T\) goes to zero with \(k_0\) constrained to lie along the ray \(k_0=\alpha e_Tk_2\), provided initial conditions are sufficiently close to the critical manifold and all other parameters are bounded away from zero. However, neither TFPV theory nor Fenichel theory indicate how small \(e_T\) must be to obtain an accurate QSS reduction. Moreover, Fenichel applies only to compact, normally hyperbolic critical manifolds.

  10. The projection operator \(\Pi ^M\) is oblique since the Dw(z)-invariant subspaces, \(N_z=\;\text {range}\; Dw(z), \;\;z\in M, \;\) and \(\ker Dw(z)=T_zM, \;\;z\in M\), are not necessarily orthogonal to one another.

References

  • Agarwal A, Adams R, Castellani GC, Shouval HZ (2012) On the precision of quasi steady state assumptions in stochastic dynamics. J Chem Phys 137(4):044105

    Article  Google Scholar 

  • Ball K, Kurtz TG, Popovic L, Rempala G (2006) Asymptotic analysis of multiscale approximations to reaction networks. Ann Appl Probab 16(4):1925–1961

    Article  MathSciNet  Google Scholar 

  • Bobadilla AVP, Bartmanski BJ, Grima R, Othmer HG (2020) The Status of the QSSA Approximation in Stochastic Simulations of Reaction Networks, in (2018) MATRIX Annals. Springer International Publishing 137–147

  • Choi B, Rempala GA, Kim JK (2017) Beyond the Michaelis-Menten equation: accurate and efficient estimation of enzyme kinetic parameters. Sci Rep 7:1708

    Article  Google Scholar 

  • Eilertsen J, Schnell S (2020) The quasi-steady-state approximations revisited: timescales, small parameters, singularities, and normal forms in enzyme kinetics. Math Biosci 325:108339

    Article  MathSciNet  Google Scholar 

  • Eilertsen J, Roussel M, Schnell S, Walcher S (2021) On the quasi-steady-state approximation in an open Michaelis-Menten reaction mechanism. AIMS Math 6:6781–6814

    Article  MathSciNet  Google Scholar 

  • Eilertsen J, Schnell S (2021) Stochastic enzyme kinetics and the quasi-steady-state reductions: Application of the slow scale linear noise approximation à la Fenichel, Under ReviewarXiv:2101.04814

  • Fenichel N (1979) Geometric singular perturbation theory for ordinary differential equations. J Differ Equ. 31:53–98

    Article  MathSciNet  Google Scholar 

  • Goeke A, Walcher S, Zerz E (2015) Determining “small parameters’’ for quasi-steady state. J Differ Equ 259:1149–1180

    Article  MathSciNet  Google Scholar 

  • Goeke A, Walcher S, Zerz E (2017) Classical quasi-steady state reduction—a mathematical characterization. Physica D 345:11–26

    Article  MathSciNet  Google Scholar 

  • Heineken FG, Tsuchiya HM, Aris R (1967) On the mathematical status of the pseudo-steady hypothesis of biochemical kinetics. Math Biosci 1:95–113

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Kampen NV (2007) Chapter X. The expansion of the master equation, in: Stochastic Processes in Physics and Chemistry (\(3^{\text{rd}}\) Edition), North-Holland Personal Library, Elsevier, Amsterdam, 244–272

  • Kang H-W, Kurtz TG (2013) Separation of time-scales and model reduction for stochastic reaction networks. Ann Appl Probab 23(2):529–583

    Article  MathSciNet  Google Scholar 

  • Kang H-W, KhudaBukhsh WR, Koeppl H, Rempala GA (2019) Quasi-steady-state approximations derived from the stochastic model of enzyme kinetics. Bull Math Biol 81(5):1303–1336

    Article  MathSciNet  Google Scholar 

  • Kim J, Josić K, Bennett M (2014) The Validity of Quasi-Steady-State approximations in discrete stochastic simulations. Biophys J 107(3):783–793

    Article  Google Scholar 

  • Kim JK, Josić K, Bennett MR (2015) The relationship between stochastic and deterministic quasi-steady state approximations. BMC Syst Biol 9:87

    Article  Google Scholar 

  • Kim JK, Rempala GA, Kang H-W (2017) Reduction for stochastic biochemical reaction networks with multiscale conservations. Multiscale Model Simul 15(4):1376–1403

    Article  MathSciNet  Google Scholar 

  • Kuehn C (2015) Multiple Time Scale Dynamics. Springer, Berlin

    Book  Google Scholar 

  • MacNamara S, Bersani AM, Burrage K, Sidje RB (2008) Stochastic chemical kinetics and the total quasi-steady-state assumption: Application to the stochastic simulation algorithm and chemical master equation. J Chem Phys 129(9):095105

    Article  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 

  • Noethen L, Walcher S (2009) Quasi-steady state and nearly invariant sets. SIAM J Appl Math 70:1341–1363

    Article  MathSciNet  Google Scholar 

  • Noethen L, Walcher S (2011) Tikhonov‘s theorem and quasi-steady state. Discrete Contin Dyn Syst Ser B 16:945–961

    MathSciNet  MATH  Google Scholar 

  • Palsson BO (1987) On the dynamics of the irreversible Michaelis-Menten reaction mechanism. Chem Eng Sci 42:447–458

    Article  Google Scholar 

  • Palsson BO, Lightfoot EN (1984) Mathematical modelling of dynamics and control in metabolic networks. I. On Michaelis-Menten kinetics. J Theor Biol 111:273–302

    Article  MathSciNet  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 

  • Sanft K, Gillespie DT, Petzold LR (2011) The legitimacy of the stochastic Michaelis-Menten approximation. IET Syst Biol 5:58–69

    Article  Google Scholar 

  • Schauer M, Heinrich R (1979) Analysis of the quasi-steady-state approximation for an enzymatic one-substrate reaction. J Theor Biol 79:425–442

    Article  Google Scholar 

  • Schnell S, Maini P (2003) A century of enzyme kinetics: reliability of the \({K}_{M}\) and \(v_{{\rm max}}\) estimates. Comm Theor Biol 8(2):169–187

    Google Scholar 

  • Segel LA (1988) On the validity of the steady state assumption of enzyme kinetics. Bull Math Biol 50:579–593

    Article  MathSciNet  Google Scholar 

  • Segel LA, Slemrod M (1989) The Quasi-Steady-State assumption: a case study in perturbation. SIAM Rev 31:446–477

    Article  MathSciNet  Google Scholar 

  • Stoleriu I, Davidson FA, Liu JL (2004) Quasi-steady state assumptions for non-isolated enzyme-catalysed reactions. J Math Biol 48(1):82–104

    Article  MathSciNet  Google Scholar 

  • Stroberg W, Schnell S (2016) On the estimation errors of \({K}_{M}\) and \(v\) from time-course experiments using the Michaelis-Menten equation. Biophys Chem 219:17–27

    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 

  • Thomas P, Straube AV, Grima R (2012a) The slow-scale linear noise approximation: an accurate, reduced stochastic description of biochemical networks under timescale separation conditions. BMC Sys Biol 6(1):39

    Article  Google Scholar 

  • Thomas P, Grima R, Straube AV (2012b) Rigorous elimination of fast stochastic variables from the linear noise approximation using projection operators. Phys Rev E 86:041110

    Article  Google Scholar 

  • Tikhonov A (1952) Systems of differential equations containing small parameters in their derivatives. Mat Sb (N.S.) 31:575–586

    Google Scholar 

  • Wechselberger M (2020) Geometric Singular Perturbation Theory Beyond the Standard Forms. Springer, Berlin

    Book  Google Scholar 

Download references

Acknowledgements

Justin Eilertsen was partially supported by the University of Michigan Postdoctoral Pediatric Endocrinology and Diabetes Training Program “Developmental Origins of Metabolic Disorder” (NIH/NIDDK Grant: T32 DK071212).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Santiago Schnell.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A. Fenichel Theory: Projecting Onto a Slow Manifold

Appendix A. Fenichel Theory: Projecting Onto a Slow Manifold

In this brief appendix, we briefly describe how to construct \(\Pi ^M\), the projection operator. For more technical details, we invite the reader to consult (Wechselberger 2020; Goeke et al. 2015, 2017). Following (Wechselberger 2020), we begin with the perturbation form

$$\begin{aligned} {\dot{z}}= w(z) + \varepsilon G(z,\varepsilon ),\quad z\in {\mathbb {R}}^n. \end{aligned}$$

and let M be a compact subset of \(w(z)=0\) that forms a k-dimensional manifold (\(k<n\)) comprised of non-isolated stationary points such that:

  • For all \(z\in M\), the algebraic and geometric multiplicities of the zero eigenvalues of Dw(z) are equal with \(T_zM = \ker Dw(z)\) and \(\dim \ker Dw(z)= k\),

  • If \(\lambda _j\) is a nonzero eigenvalue of Dw(z), then \(\mathfrak {Re}(\lambda _j) < 0\quad \forall z\in M\),

then, M is locally attracting and there exists a splitting

$$\begin{aligned} {\mathbb {R}}^n:=T_zM \oplus N_z \quad \forall z \in M, \end{aligned}$$
(A.1)

where \(T_zM\) is the tangent space of M at z given by \(\{ x \in {\mathbb {R}}^n: x \in \ker Dw(z)\}\), and \(N_z\) is the complement to \(T_zM\) and coincides with range of the Jacobian, \({\mathcal {R}}(Dw(z))\). The objective from the point forward will be to exploit the splitting (A.1) and construct an obliqueFootnote 10 projection operator:

$$\begin{aligned} \Pi ^M: {\mathbb {R}}^n \mapsto T_zM, \qquad I-\Pi ^M: {\mathbb {R}}^n \mapsto N_z. \end{aligned}$$

To construct \(\Pi ^M\), we invoke the factorization

$$\begin{aligned} w(z) = P(z)f(z), \end{aligned}$$
(A.2)

where P(z) is a rectangular matrix function with full rank at every point in M, and the zero level set of f(z) coincides with the critical manifold, M. The derivative of f(z), Df(z), has full rank \(\forall z \in M\) and:

  • The columns of P(z) form a basis for the range of the Jacobian \({\mathcal {R}}(Dw(z))=N_z\), for all \(z\in M\).

  • The rows of Df(z) form a basis for the orthogonal complement of \(\ker Dw(z)=(T_zM)^{\perp }\) for all \(z \in M\).

It follows that

since \(T_zM\) and \(N_z\) are complementary subspaces, the matrix

$$\begin{aligned} \Pi ^M := I- P(Df P)^{-1}Df \end{aligned}$$
(A.3)

defines the oblique projection onto \(T_zM\). To leading order in \(\varepsilon \), the reduced flow on the slow manifold is

$$\begin{aligned} {\dot{z}} = \Pi ^M G(z,0), \end{aligned}$$
(A.4)

which is the QSS reduction obtained from Fenichel theory (Fenichel 1979).

Remark 4

The critical manifold is said to be normally hyperbolic if the matrix DfP (the eigenvalues of which are the non-trivial eigenvalues of Dw(z)) is hyperbolic \(\forall z\in M\), meaning that \(\mathfrak {Re}(\lambda _i) \ne 0.\) In most applications, we are interested in attracting critical manifolds, and thus, we often require \(\mathfrak {Re}(\lambda _i) < 0.\) However, normally hyperbolic critical manifolds are repelling if \(\mathfrak {Re}(\lambda _i) > 0\), or of saddle type if the real parts of the eigenvalues of DfP are both positive and negative.

As an example, we will explicitly compute the sQSSA for the open MM reaction mechanism. In standard form, we have

$$\begin{aligned} \begin{pmatrix} {\dot{s}} \\ {\dot{c}}\end{pmatrix} = \begin{pmatrix}k_{-1} +k_1s \\ -k_{-1}-k_2-k_1s \end{pmatrix}c + \varepsilon \begin{pmatrix} k_0^*-k_1e_T^*s\\ k_1e_T^*s \end{pmatrix} \end{aligned}$$
(A.5)

so that \(f(s,c)=c\) and

$$\begin{aligned} P \equiv \begin{pmatrix}k_{-1} + k_1s \\ -k_{-1}-k_2-k_1s\end{pmatrix}. \end{aligned}$$
(A.6)

The derivative of f(sc) is \([0 \;\;1]\), and thus \((DfP)^{-1}\) is

$$\begin{aligned} -\frac{1}{k_1(K_M+s)}, \end{aligned}$$
(A.7)

which is of course a scalar. The product, PDf, is given by

$$\begin{aligned} \begin{pmatrix} 0 &{} k_1(K_S+s)\\ 0 &{} -k_1(K_M+s)\end{pmatrix}, \quad K_S:=k_{-1}/k_1, \quad K_M:=(k_{-1}+k_2)/k_1. \end{aligned}$$
(A.8)

Computing \(\Pi ^M\) from (A.3) yields

$$\begin{aligned} \Pi ^M:=\begin{pmatrix} 1 &{} \frac{(K_S+s)}{(K_M+s)}\\ 0 &{} 0\end{pmatrix}. \end{aligned}$$
(A.9)

To recover the sQSSA, we simply project the perturbation onto the tangent space of the critical manifold:

$$\begin{aligned} \begin{pmatrix} {\dot{s}} \\ {\dot{c}}\end{pmatrix} = \Bigg [\begin{pmatrix} 1 &{} \frac{(K_S+s)}{(K_M+s)}\\ 0 &{} 0\end{pmatrix}\begin{pmatrix} k_0^*-k_1e_T^*s\\ k_1e_T^*s \end{pmatrix}\Bigg ]\Bigg |_{c=0} =\begin{pmatrix} k_0^*-\frac{k_1k_2 e_T^*s}{k_{-1}+k_2+k_1s}\\ 0\end{pmatrix}.\nonumber \\ \end{aligned}$$
(A.10)

The QEA (26) that corresponds to small \(k_0\) and small \(k_2\) is computed in a similar manner. For more details see (Eilertsen et al. 2021).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Eilertsen, J., Schnell, S. On the Validity of the Stochastic Quasi-Steady-State Approximation in Open Enzyme Catalyzed Reactions: Timescale Separation or Singular Perturbation?. Bull Math Biol 84, 7 (2022). https://doi.org/10.1007/s11538-021-00966-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-021-00966-5

Keywords

Navigation