Skip to main content
Log in

Extension and Justification of Quasi-Steady-State Approximation for Reversible Bimolecular Binding

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

Abstract

The quasi-steady-state approximation (QSSA) is commonly applied in chemical kinetics without rigorous justification. We provide details of such a justification in the ubiquitous case of reversible two-step bimolecular binding in which molecules as an intermediate step of the reaction form a transient complex. First, we justify QSSA in the regime that agrees with the results in the literature and is characterized by \(\max \{R_0, L_0\} \ll K_m\). Here, \(R_0\) and \(L_0\) are the initial concentrations of reacting receptor and ligand, respectively, and \(K_m\) is the Michaelis constant. We also validate QSSA under an alternative condition that can be viewed as partially irreversible binding, and it does not require a tight bound on \(R_0\) and \(L_0\) but rather requires \(k_2 + k_{-2} \ll k_{-1}\). Here, \(k_{-1}\) is the rate constant of decomposition of the transient complex to the ligand and the receptor, and \(k_2\) and \(k_{-2}\) are the forward and the reverse rate constants of transformation of the complex to the product, respectively. Furthermore, we provide arguments that QSSA can also be accurate in a regime when \(\max \{R_0, L_0\} \approx K_m\) and \(k_2 + k_{-2} \approx k_{-1}\) if \(|R_0 - L_0| \ll K_m\). The derived conditions may be of practical use as they provide weaker requirements for the validity of QSSA compared to the existing results.

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

Similar content being viewed by others

Notes

  1. To complete the proof of the forward invariance of \(\varOmega _0\), it is also necessary to check that the phase flow points inwards at the corners of \(\varOmega _0\). That follows directly from the fact that the phase flow \((\ell ,v)\) is continuous, and it points inwards on the rest of the boundary. The same argument can be used in the full proof of Theorem 1 in Sect. 2.7.

  2. Note that the a priori bounds for \(V(\ell , v)\) are not sharp as the highest value of individual terms is often reached for different values of \(\ell \). The estimates can be further improved by using \(\hbox {d}\ell / \hbox {d}\tau \le 0\) proved in Sect. 2.1. It allows to restrict the range of \(\ell \) in the bounding box \(\varOmega \) to \(\ell ^{*} \le \ell \le 1\).

References

  • Borghans JAM, Boer RJD, Segel LA (1996) Extending the quasi-steady state approximation by changing variables. Bull Math Biol 58:43–63

    Article  MATH  Google Scholar 

  • Briggs GE, Haldane JBS (1925) A note on the kinetics of enzyme action. Biochem J 19:338–339

    Article  Google Scholar 

  • Calder MS, Siegel D (2008) Properties of the Michaelis-Menten mechanism in phase space. J Math Anal Appl 339:1044–1064

    Article  MATH  MathSciNet  Google Scholar 

  • Coddington E, Levison N (1955) Theory of ordinary differential equations. McGraw Hill, New York

    MATH  Google Scholar 

  • Crooke PS, Tanner RD, Aris R (1979) The role of dimensionless parameters in the Briggs-Haldane and Michaelis-Menten approximations. Chem Eng Sci 34:1354–1357

    Article  Google Scholar 

  • Darvey IG, Matlak RF (1967) An investigation of a basic assumption in enzyme kinetics using results of the geometric theory of differential equations. Bull Math Biophys 29:335–341

    Article  MATH  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

  • François P, Voisinne G, Siggia ED, Altan-Bonnet G, Vargassola M (2013) Phenotypic model for early T-cell activation displaying sensitivity, specificity, and antagonism. Proc Natl Acad Sci USA 110:E888–E897

    Article  Google Scholar 

  • Fraser SJ (1988) The steady state and equilibrium approximations: a geometrical picture. J Chem Phys 88:4732–4738

    Article  Google Scholar 

  • Fraser SJ, Roussel MR (1994) Phase-plane geometries in enzyme kinetics. Can J Chem 72:800–812

    Article  Google Scholar 

  • Goeke A, Walcher S (2013) Quasi-steady state: searching for and utilizing small parameters. In: Recent trends in dynamical systems. Springer proceedings in mathematics and statistics, vol. 35, pp 153–178

  • Gorban AN, Shahzad M (2011) The Michaelis-Menten-Stueckelberg theorem. Entropy 13:966–1019

    Article  MATH  MathSciNet  Google Scholar 

  • Hanson SM, Schnell S (2008) Reactant stationary approximation in the enzyme kinetics. J Phys Chem A 112:8654–8658

    Article  Google Scholar 

  • Henri V (1902) Théorie générale de laction de quelques diastases. C R Acad Sci Paris 135:916–919

    Google Scholar 

  • Henri V (1903) Lois Générales de l’Action des Diastases. Librairie Scientifique A, Hermann, Paris

    Google Scholar 

  • Jones CKRT (1994) Geometric singular perturbation theory. In: Arnold L (ed) Dynamical systems, Montecatini Terme. Springer, Berlin. Volume 1609 of Lecture Notes in Mathematics, pp 44–118

  • Kaper HG, Kaper TJ (2002) Asymptotic analysis of two reduction methods for systems of chemical reactions. Phys D 165:66–93

    Article  MATH  MathSciNet  Google Scholar 

  • Kaper TJ (1999) An introduction to geometric methods and dynamical systems theory for singular perturbation problems. In: Cronin J, R E O’Mailey J (eds) Analyzing multiscale phenomena using singular perturbation methods. American Mathematical Society, Providence, pp 85–131

    Chapter  Google Scholar 

  • Kollár R, Bodová K, Tomáška L, Nosek J (2014) Mathematical model of alternative mechanism of telomere length maintenance. Phys Rev E 89:032701

    Article  Google Scholar 

  • McKeithan TW (1995) Kinetic proofreading in T-cell receptor signal transduction. Proc Natl Acad Sci USA 92:5042–5046

    Article  Google Scholar 

  • Michaelis L, Menten ML (1913) Die kinetik der invertinwirkung. Biochem Z 49:333–369

    Google Scholar 

  • Miller WG, Alberty RA (1958) Kinetics of the reversible Michaelis-Menten mechanism and the applicability of the steady-state approximation. J Am Chem Soc 80:5146–5151

    Article  Google Scholar 

  • Mischenko EF, Kolesov YS, Kolesov AY, Rozov NK (1994) Asymptotic methods in singularly perturbed systems. Consultants Bureau, New York

    Book  Google Scholar 

  • Murray JD (2002) Mathematical biology. I: an introduction. Springer, NewYork

    Google Scholar 

  • Ninio J (1975) Kinetic amplification of enzyme discrimination. Biochemie 57:587–595

    Article  Google Scholar 

  • Nipp K (1991) Numerical integration of stiff odes of singular perturbation type. Zeitsch Angew Math Phys 42:53–79

    Article  MATH  MathSciNet  Google Scholar 

  • Noethen L, Walcher S (2007) Quasi-steady state in the Michaelis-Menten system. Nonlinear Anal RealWorld Appl 8:1512–1535

    Article  MATH  MathSciNet  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

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

    Article  MATH  MathSciNet  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 

  • Perelson A, Rocklin S, Goldstein B (1980) Optimal strategies in immunology III. the IGM-IGG switch. J Math Biol 10:209–256

    Article  MATH  Google Scholar 

  • Roussel MR, Fraser SJ (1990) Geometry of the steady-state approximation: perturbation and accelerated convergence methods. J Chem Phys 93:1072–1081

    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 (2014) Validity of the Michaelis-Menten equation—steady-state or reactant stationary assumption: that is the question. FEBS J 281:464–472

    Article  Google Scholar 

  • Schnell S, Maini PK (2000) Enzyme kinetics at high enzyme concentrations. Bull Math Biol 62:483–499

    Article  Google Scholar 

  • Schnell S, Maini PK (2002) Enzyme kinetics far from the standard quasi-steady-state and equilibrium approximations. Math Comput Model 35:137–144

    Article  MATH  MathSciNet  Google Scholar 

  • Schnell S, Maini PK (2003) A century of enzyme kinetics: reliability of the \(k_m\) and \(v_{\max }\) estimates. Comm Theor Biol 8:169–197

    Article  Google Scholar 

  • Schnell S, Mendoza C (1997) Closed form solution for time-dependent enzyme kinetics. J Theor Biol 187:207–212

    Article  Google Scholar 

  • Segel IH (1975) Enzyme kinetics: behavior and analysis of rapid equilibrium and steady-state enzyme systems. Wiley, New York

    Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

  • Segel LA, Edelstein-Keshet L (2013) A primer on mathematical models in biology. SIAM, Philadelphia

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

    Article  MATH  MathSciNet  Google Scholar 

  • Šišková K (2014a) Rigorous analysis of quasi-steady state for chemical kinetics. Master’s thesis. Comenius University. Bratislava

  • Šišková K (2014b) Rigorous analysis of quasi-steady state for chemical kinetics. In: Plachetka T, Vinař T (eds) Proceedings of the student science conference. Library and Publishing Centre, FMPI Comenius University, Bratislava

  • Tzafriri AR (2003) Michaelis-Menten kinetics at high enzyme concentrations. Bull Math Biol 65:1111–1129

    Article  Google Scholar 

  • Tzafriri AR, Edelman ER (2004) The total quasi-steady-state approximation is valid for reversible enzyme kinetics. J Theor Biol 226:303–313

    Article  MathSciNet  Google Scholar 

  • Tzafriri AR, Edelman ER (2005) On the validity of the quasi-steady state approximation of bimolecular reactions in solution. J Theor Biol 233:343–350

    Article  Google Scholar 

  • Walter C (1974) The validity of using a quasi-steady-state approximation for the reversible Michaelis-Menten mechanism of enzyme action. J Theor Biol 44:1–5

    Article  Google Scholar 

Download references

Acknowledgments

The authors are grateful to the anonymous referees who significantly helped to improve the manuscript. The authors would also like to thank Pavol Brunovský and Katarína Bod’ová for many helpful suggestions and gratefully acknowledge the financial contribution of European Commission under Marie Curie International Reintegration Grant 239429 and of the Scientific Grant Agency of the Slovak Republic under the grant 1/0459/13.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Richard Kollár.

Appendix

Appendix

1.1 Monotonicity

First we prove that \(\hbox {d}L/\hbox {d}t \le 0\) for all \(t\ge 0\) for the solution of (12)–(13) subject to the initial condition \((L(0), C(0)) = (L_0, 0)\). The same property then holds for the solution of (6)–(9) subject to (10). Let us assume the contrary, i.e., there is a trajectory (L(t), C(t)) starting at \((L_0, 0)\) such that \(dL / \hbox {d}t(t_1) > 0\) for some \(t_1 > 0\). Since the vector field \((\hbox {d}L/\hbox {d}t, \hbox {d}C/\hbox {d}t)\) points downwards on the C–nullcline for \(L > L^{*}\), the trajectory (L(t), C(t)) can only cross the nullcline for \(L \in (0,L^{*})\). Therefore, there exist \(t_2 \in (0, t_1)\) such that the trajectory (L(t), C(t)) crosses the line segment \(L = L^{*}, C \in [0,C^{*})\) in the phase plane for the first time, i.e., \(L(t) > L^{*}\) for all \(t \in [0, t_2)\). On that segment, \(L(t_2) = L^{*}\) and \(C(t_2) < C^{*}\), and thus by (11) also \(P(t_2) > P^{*}\). Since \(P(0) = 0\) by continuity, there must exist \(t_3 \in (0, t_2)\) such that \(P(t_3) = P^{*}\) and \(\hbox {d}P / \hbox {d}t (t_3) \ge 0\). But (9) implies \(0 \le \hbox {d}P / \hbox {d}t (t_3) = k_2 C(t_3) - k_{-2} P^{*}\), i.e., \(C(t_3) \ge k_{-2}/k_2 P^{*} = C^{*}\). Therefore, \(P(t_3) = P^{*}, C(t_3) \ge C^{*}\), and hence \(L(t_3) \le L^{*}\), yielding a contradiction because \(L(t) > L^{*}\) for all \(t \in [0,t_2)\). Therefore, \(dL/dt \le 0\) for all \(t >0\).

Furthermore, we prove \(\hbox {d}P/\hbox {d}t \ge 0\) for all \(t\ge 0\) for the solution satisfying (6)–(9) subject to (10) by using an argument similar to Darvey and Matlak (1967). Using (6)–(9), the initial conditions, and continuity of the solution, we derive

$$\begin{aligned} \frac{\hbox {d}}{\hbox {d}t} (k_2 C - k_{-2} P) = k_2 k_1 L R - k_2 (k_{-1} + k_2 + k_{-2}) C + k_{-2}(k_2 + k_{-2}) P \ge 0\nonumber \\ \end{aligned}$$
(62)

for all small enough positive t. The stoichiometry restrictions (11) allow us to reduce the system (6)–(9) to the two-dimensional dynamical system in variables (LP) of the form

$$\begin{aligned} \frac{\hbox {d}L}{\hbox {d}t}= & {} -k_1 L(L + R_0 - L_0) + k_{-1} (L_0 - L - P), \\ \frac{\hbox {d}P}{\hbox {d}t}= & {} k_2 (L_0 - L) - (k_2 + k_{-2}) P. \end{aligned}$$

The P-nullcline is the line connecting the initial condition \((L_0, 0)\) and the point \((0, k_2L_0/(k_2 + k_{-2}))\) in the (LP) phase plane. Furthermore, the L-nullcline is given by the parabola \(P = L_0 - L - k_1L(L+R_0 - L_0)/k_{-1}\) that contains points \((0, L_0), (L_0,-k_1L_0R_0/k_{-1})\) and that is decreasing for all \(L \ge 0\). Therefore, there is a unique equilibrium of the system in the positive quadrant. The condition (62) and the fact that \(dP /dt = 0\) at \(t=0\) imply that the solution (L(t), P(t)) lies under the P-nullcline for small positive t. Since the solution (L(t), P(t)) cannot cross neither of the nullclines, it remains wedged in between them for all positive t. That immediately implies \(\hbox {d}P / \hbox {d}t \ge 0\) for all \(t \ge 0\).

1.2 Estimates

Estimate of \(t_{L}\). The numerator in the expression (16) for \(t_L\) is equal to the range of values of L given by \(L_\mathrm{max}-L_\mathrm{min}\). It can be estimated by \(L_0\) as \(L + C + P = L_0\), and both P and C are nonnegative. Since \(\hbox {d}L / \hbox {d}t \le 0\) for all t, a sharper estimate \(L_\mathrm{max} - L_\mathrm{min} = L_0 - L^{*}\), where \(L^{*}\) is the equilibrium concentration of L, can be used as well. On the other hand, the denominator in (16) that is equal to the maximal value of \(\hbox {d}L / \hbox {d}t\) can be estimated in the following way. By (12) at the point of its maximum, the function

$$\begin{aligned} U(t) = -k_1 L(t) (L(t) + R_0 - L_0) + k_{-1}C(t) \end{aligned}$$

has its local extreme, i.e., \(\hbox {d}U/ \hbox {d}t = 0\). Hence,

$$\begin{aligned} k_{-1} \frac{\hbox {d}C}{\hbox {d}t} = k_1 \frac{\hbox {d}L}{\hbox {d}t} (2L + R_0 - L_0). \end{aligned}$$

Using (12)–(13), it reduces to an expression for \(C = C(L)\) that is valid at the point of the local maximum:

$$\begin{aligned} C = \frac{\frac{k_{-1}k_{-2}}{k_1} (L_0 - L) + L(L+R_0-L_0)\left[ k_{-1} + k_1(2L + R_0 - L_0)\right] }{k_{-1}(2 L + R_0 - L_0 + K_m)}. \end{aligned}$$

The value of \(\hbox {d}L/\hbox {d}t\) at its eventual point of local maximum is then

$$\begin{aligned} \frac{\hbox {d}L}{\hbox {d}t} \Big |_{\mathrm{max}} = \frac{-(k_2 + k_{-2})L(L+R_0 - L_0) + \frac{k_{-1}k_{-2}}{k_1}(L_0 - L)}{2L + R_0 - L_0 + K_m}\, . \end{aligned}$$
(63)

If we consider the right-hand side of (63) as a function of L, it is easy to see that it is a decreasing function on its natural domain \(L \in [0, L_0]\), and thus, it reaches its maximum absolute value at one of its end points. Its value at \(L = 0\) is positive and at \(L = L_0\) negative. Since we have proved that \(\hbox {d}L / \hbox {d}t \le 0\) for all \(t \ge 0\), it is enough to consider the value at \(L = L_0\) given by

$$\begin{aligned} \Big |\frac{\hbox {d}L}{\hbox {d}t}\Big |_{\mathrm{max}, L = L_0} = (k_2 + k_{-2}) \frac{L_0R_0}{R_0 + L_0 + K_m}. \end{aligned}$$

The maximum possible value of |dL / dt| is then bounded by the maximum of the values of U(t) at \(t = 0 (L = L_0\) and \(C = 0)\), at \(t = \infty \), i.e., at the system equilibrium for which \(U(t) = 0\) (this follows from the analysis performed in Sect. 2.5) and at its eventual local extreme. Since at \(t = 0, |U(0)|= k_1 L_0R_0\) and \(k_1 K_m \ge k_2 + k_{-2}\), it easy to see that

$$\begin{aligned} \Big |\frac{\hbox {d}L}{\hbox {d}t}\Big |_{\mathrm{max}} \le \max \Big \{k_1 L_0R_0, \frac{(k_2 + k_{-2})L_0R_0}{R_0 + L_0 + K_m}\Big \} = k_1 L_0R_0, \end{aligned}$$

that leads to (17).

Estimate of \(\triangle L / L_0\). Next, we estimate the relative depletion of the ligand during the initial short timescale \(t \in [0, t_C]\), where \(t_C\) is given by (15). The formula for \(\triangle L = L_0 - L(t_C)\) follows from (12):

$$\begin{aligned} \triangle L = \left| -\int _0^{t_C} \frac{\hbox {d}L}{\hbox {d}t} \, \hbox {d}t \right| = \left| -k_{-1}\int _0^{t_C} C \, \hbox {d}t + k_1 \int _0^{t_C} L (L + R_0- L_0)\, \hbox {d}t \right| . \end{aligned}$$
(64)

Here, both integrals in (64) are positive. Next, we estimate the first integral in (64). Note that \(\hbox {d} / \hbox {d}L (\hbox {d}C/\hbox {d}t) > 0\); therefore, the production of C is increasing with concentration of L. Since \( L \le L_0\) and at the concentration \(L = L_0\) the evolution of C is governed by (14), it follows that

$$\begin{aligned} C(t) \le \frac{L_0R_0}{K_m} \left( 1 - \hbox {e}^{-(k_{-1}+k_2 + k_{-2})t} \right) . \end{aligned}$$

Therefore,

$$\begin{aligned} k_{-1}\int _0^{t_C} C \, \hbox {d}t \le k_{-1}\frac{L_0R_0}{K_m} \int _0^{t_C} 1 - \hbox {e}^{-(k_{-1}+k_2 + k_{-2})t} \hbox {d}t = \frac{L_0R_0}{K_m}\frac{1-\kappa }{\hbox {e}}. \end{aligned}$$

The second integral in (64) can be estimated by the trapezoid rule using the fact that \(L(t_C) = L_0 - \triangle L\):

$$\begin{aligned} k_1 \int _0^{t_C} L(L + R_0 - L_0) \hbox {d}t\approx & {} k_1 t_C \frac{1}{2} \left( L_0 R_0 + (L_0 - \triangle L)(R_0 - \triangle L) \right) \\\approx & {} \frac{L_0R_0}{K_m} - \frac{\triangle L}{K_m} \frac{L_0 + R_0}{2}, \end{aligned}$$

where we neglected the quadratic terms \(O(\triangle ^2 L)\). The quantity \(\triangle L\) in (64) can be then estimated as

$$\begin{aligned} \triangle L \approx -\frac{L_0 R_0}{K_m} \frac{1-\kappa }{\hbox {e}} + \frac{L_0R_0}{K_m} - \triangle L \frac{L_0 + R_0}{2K_m}, \end{aligned}$$

that can be rewritten as

$$\begin{aligned} \triangle L \left( 1 +\frac{L_0 + R_0}{2K_m}\right) \approx \frac{L_0 R_0}{K_m} \left( 1 - \frac{1-\kappa }{e}\right) . \end{aligned}$$

Note that \(\kappa \in [0,1]\), i.e., \(1 - (1-\kappa ) / e= O(1)\). Hence,

$$\begin{aligned} |\triangle L| \approx O\left( \frac{L_0R_0}{2K_m + L_0+R_0}\right) . \end{aligned}$$
(65)

The reactant stationary assumption \(\triangle L/L_0 \ll 1\) can be then expressed as

$$\begin{aligned} \frac{R_0}{2K_m + L_0+R_0} \ll 1 \end{aligned}$$

that is for \(L_0 \le R_0\) equivalent to \(R_0 / K_m \ll 1\). However, note that for any \(R_0 - L_0 < 2K_m\), one has

$$\begin{aligned} \frac{R_0}{2K_m + L_0+R_0} = \frac{R_0}{2K_m + 2R_0 + (L_0 - R_0)} < \frac{1}{2}. \end{aligned}$$

For values of \(\kappa \) close to \(0^+\), the neglected factor \((1-\kappa ) /\hbox {e}\) in (65) is smaller than 2 / 3, i.e., \(\triangle L / L_0 < 1/3\), regardless of the value of \(R_0/K_m\). For moderate values of \(R_0/K_m \approx \frac{1}{2}\) and \(R_0 = L_0\), the ratio \(\triangle L / L_0\) can even reach the value 0.1.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kollár, R., Šišková, K. Extension and Justification of Quasi-Steady-State Approximation for Reversible Bimolecular Binding. Bull Math Biol 77, 1401–1436 (2015). https://doi.org/10.1007/s11538-015-0090-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-015-0090-8

Keywords

Navigation