Skip to main content
Log in

Comparison of Deterministic and Stochastic Regime in a Model for Cdc42 Oscillations in Fission Yeast

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

A Commentary to this article was published on 07 February 2019

Abstract

Oscillations occur in a wide variety of essential cellular processes, such as cell cycle progression, circadian clocks and calcium signaling in response to stimuli. It remains unclear how intrinsic stochasticity can influence these oscillatory systems. Here, we focus on oscillations of Cdc42 GTPase in fission yeast. We extend our previous deterministic model by Xu and Jilkine to construct a stochastic model, focusing on the fast diffusion case. We use SSA (Gillespie’s algorithm) to numerically explore the low copy number regime in this model, and use analytical techniques to study the long-time behavior of the stochastic model and compare it to the equilibria of its deterministic counterpart. Numerical solutions suggest noisy limit cycles exist in the parameter regime in which the deterministic system converges to a stable limit cycle, and quasi-cycles exist in the parameter regime where the deterministic model has a damped oscillation. Near an infinite period bifurcation point, the deterministic model has a sustained oscillation, while stochastic trajectories start with an oscillatory mode and tend to approach deterministic steady states. In the low copy number regime, metastable transitions from oscillatory to steady behavior occur in the stochastic model. Our work contributes to the understanding of how stochastic chemical kinetics can affect a finite-dimensional dynamical system, and destabilize a deterministic steady state leading to oscillations.

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.

Institutional subscriptions

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

Similar content being viewed by others

References

  • Altschuler SJ, Angenent SB, Wang Y, Wu LF (2008) On the spontaneous emergence of cell polarity. Nature 454(7206):886–889

    Google Scholar 

  • Amiranashvili A, Schnellbächer ND, Schwarz US (2016) Stochastic switching between multistable oscillation patterns of the Min-system. New J Phys 18(9):093049

    Google Scholar 

  • Anderson DF, Enciso GA, Johnston MD (2014) Stochastic analysis of biochemical reaction networks with absolute concentration robustness. J R Soc Interface 11(93):20130943

    Google Scholar 

  • Anderson DF, Cappelletti D, Kurtz TG (2017) Finite time distributions of stochastically modeled chemical systems with absolute concentration robustness. SIAM J Appl Dyn Syst 16(3):1309–1339

    MathSciNet  MATH  Google Scholar 

  • Ashkenazi M, Othmer HG (1978) Spatial patterns in coupled biochemical oscillators. J Math Biol 5(4):305–350

    MathSciNet  MATH  Google Scholar 

  • Barik D, Paul MR, Baumann WT, Cao Y, Tyson JJ (2008) Stochastic simulation of enzyme-catalyzed reactions with disparate timescales. Biophys J 95(8):3563–3574

    Google Scholar 

  • Barik D, Ball DA, Peccoud J, Tyson JJ (2016) A stochastic model of the yeast cell cycle reveals roles for feedback regulation in limiting cellular variability. PLoS Comput Biol 12(12):e1005230

    Google Scholar 

  • Bendezú FO, Vincenzetti V, Vavylonis D, Wyss R, Vogel H, Martin SG (2015) Spontaneous Cdc42 polarization independent of GDI-mediated extraction and actin-based trafficking. PLoS Biol 13(4):e1002097

    Google Scholar 

  • Benzi R, Sutera A, Vulpiani A (1981) The mechanism of stochastic resonance. J Phys A Math Gen 14(11):L453

    MathSciNet  Google Scholar 

  • Bonazzi D, Haupt A, Tanimoto H, Delacour D, Salort D, Minc N (2015) Actin-based transport adapts polarity domain size to local cellular curvature. Curr Biol 25(20):2677–2683

    Google Scholar 

  • Bressloff PC (2010) Metastable states and quasicycles in a stochastic Wilson-Cowan model of neuronal population dynamics. Phys Rev E 82(5):051903

    MathSciNet  Google Scholar 

  • Chang F, Martin SG (2009) Shaping fission yeast with microtubules. Cold Spring Harbor Perspect Biol 1(1):a001347

    Google Scholar 

  • Chiang H-D, Thorp JS (1989) Stability regions of nonlinear dynamical systems: a constructive methodology. IEEE Trans Autom Control 34(12):1229–1241

    MathSciNet  MATH  Google Scholar 

  • Das M, Drake T, Wiley DJ, Buchwald P, Vavylonis D, Verde F (2012) Oscillatory dynamics of Cdc42 GTPase in the control of polarized growth. Science 337(6091):239–243

    Google Scholar 

  • Dauxois T, Di Patti F, Fanelli D, McKane AJ (2009) Enhanced stochastic oscillations in autocatalytic reactions. Phys Rev E 79(3):036112

    Google Scholar 

  • Enciso GA (2016) Transient absolute robustness in stochastic biochemical networks. J R Soc Interface 13(121):20160475

    Google Scholar 

  • Endo M, Shirouzu M, Yokoyama S (2003) The Cdc42 binding and scaffolding activities of the fission yeast adaptor protein Scd2. J Biol Chem 278(2):843–852

    Google Scholar 

  • Erban R, Chapman SJ, Kevrekidis IG, Vejchodskỳ T (2009) Analysis of a stochastic chemical system close to a SNIPER bifurcation of its mean-field model. SIAM J Appl Math 70(3):984–1016

    MathSciNet  MATH  Google Scholar 

  • Etienne-Manneville S, Hall A (2002) Rho GTPases in cell biology. Nature 420(6916):629

    Google Scholar 

  • Forger DB, Peskin CS (2005) Stochastic simulation of the mammalian circadian clock. Proc Natl Acad Sci 102(2):321–324

    Google Scholar 

  • Freisinger T, Klünder B, Johnson J, Müller N, Pichler G, Beck G, Costanzo M, Boone C, Cerione RA, Frey E et al (2013) Establishment of a robust single axis of cell polarity by coupling multiple positive feedback loops. Nat Commun 4:1807

    Google Scholar 

  • Gammaitoni L, Hänggi P, Jung P, Marchesoni F (1998) Stochastic resonance. Rev Mod Phys 70(1):223

    Google Scholar 

  • Gang H, Ditzinger T, Ning CZ, Haken H (1993) Stochastic resonance without external periodic force. Phys Rev Lett 71(6):807

    Google Scholar 

  • Gardiner C (2009) Stochastic methods, vol 4. Springer, Berlin

    MATH  Google Scholar 

  • Geffert PM (2015) Stochastic non-excitable systems with time delay: modulation of noise effects by time-delayed feedback. Springer, Berlin

    MATH  Google Scholar 

  • Geva-Zatorsky N, Rosenfeld N, Itzkovitz S, Milo R, Sigal A, Dekel E, Yarnitzky T, Liron Y, Polak P, Lahav G et al (2006) Oscillations and variability in the p53 system. Mol Syst Biol 2:2006.0033

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

    Google Scholar 

  • Gonze D, Halloy J, Goldbeter A (2002a) Deterministic versus stochastic models for circadian rhythms. J Biol Phys 28(4):637–653

    Google Scholar 

  • Gonze D, Halloy J, Goldbeter A (2002b) Robustness of circadian rhythms with respect to molecular noise. Proc Natl Acad Sci 99(2):673–678

    Google Scholar 

  • Goryachev AB, Leda M (2017) Many roads to symmetry breaking: molecular mechanisms and theoretical models of yeast cell polarity. Mol Biol Cell 28(3):370–380

    Google Scholar 

  • Hegemann B, Unger M, Lee SS, Stoffel-Studer I, van den Heuvel J, Pelet S, Koeppl H, Peter M (2015) A cellular system for spatial signal decoding in chemical gradients. Dev Cell 35(4):458–470

    Google Scholar 

  • Howard M, Rutenberg AD (2003) Pattern formation inside bacteria: fluctuations due to the low copy number of proteins. Phys Rev Lett 90(12):128102

    Google Scholar 

  • Hu J, Kang H-W, Othmer HG (2014) Stochastic analysis of reaction–diffusion processes. Bull Math Biol 76(4):854–894

    MathSciNet  MATH  Google Scholar 

  • Jilkine A, Angenent SB, Wu LF, Altschuler SJ (2011) A density-dependent switch drives stochastic clustering and polarization of signaling molecules. PLoS Comput Biol 7(11):e1002271

    Google Scholar 

  • Johnson JM, Jin M, Lew DJ (2011) Symmetry breaking and the establishment of cell polarity in budding yeast. Curr Opin Genet Dev 21(6):740–746

    Google Scholar 

  • Johnston MD, Anderson DF, Craciun G, Brijder R (2018) Conditions for extinction events in chemical reaction networks with discrete state spaces. J Math Biol 76(6):1535–1558

    MathSciNet  MATH  Google Scholar 

  • Kang H-W, Kurtz TG, Popovic L (2014) Central limit theorems and diffusion approximations for multiscale markov chain models. Ann Appl Probab 24(2):721–759

    MathSciNet  MATH  Google Scholar 

  • Kar S, Baumann WT, Paul MR, Tyson JJ (2009) Exploring the roles of noise in the eukaryotic cell cycle. Proc Natl Acad Sci 106(16):6471–6476

    Google Scholar 

  • Keizer J (1987) Statistical thermodynamics of nonequilibrium processes. Springer, Berlin

    Google Scholar 

  • Kerr RA, Levine H, Sejnowski TJ, Rappel W-J (2006) Division accuracy in a stochastic model of Min oscillations in Escherichia coli. Proc Natl Acad Sci USA 103(2):347–352

    Google Scholar 

  • Kim JK, Josić K, Bennett MR (2014) The validity of quasi-steady-state approximations in discrete stochastic simulations. Biophys J 107(3):783–793

    Google Scholar 

  • Klünder B, Freisinger T, Wedlich-Söldner R, Frey E (2013) GDI-mediated cell polarization in yeast provides precise spatial and temporal control of Cdc42 signaling. PLoS Comput Biol 9(12):e1003396

    Google Scholar 

  • Kuo C-C, Savage NS, Chen H, Wu C-F, Zyla TR, Lew DJ (2014) Inhibitory GEF phosphorylation provides negative feedback in the yeast polarity circuit. Curr Biol 24(7):753–759

    Google Scholar 

  • Kurtz TG (1971) Limit theorems for sequences of jump markov processes approximating ordinary differential processes. J Appl Probab 8(2):344–356

    MathSciNet  MATH  Google Scholar 

  • Kurtz TG (1972) The relationship between stochastic and deterministic models for chemical reactions. J Chem Phys 57(7):2976–2978

    Google Scholar 

  • Kuske R, Gordillo LF, Greenwood P (2007) Sustained oscillations via coherence resonance in SIR. J Theor Biol 245(3):459–469

    MathSciNet  Google Scholar 

  • Lawson MJ, Drawert B, Khammash M, Petzold L, Yi T-M (2013) Spatial stochastic dynamics enable robust cell polarization. PLoS Comput Biol 9(7):e1003139

    Google Scholar 

  • Lipan O, Ferwerda C (2018) Hill functions for stochastic gene regulatory networks from master equations with split nodes and time-scale separation. Phys Rev E 97(2):022413

    Google Scholar 

  • Manninen T, Linne M-L, Ruohonen K (2006) Developing Itô stochastic differential equation models for neuronal signal transduction pathways. Comput Biol Chem 30(4):280–291

    MATH  Google Scholar 

  • McKane AJ, Newman TJ (2005) Predator–prey cycles from resonant amplification of demographic stochasticity. Phys Rev Lett 94(21):218102

    Google Scholar 

  • McKane AJ, Nagy JD, Newman TJ, Stefanini MO (2007) Amplified biochemical oscillations in cellular systems. J Stat Phys 128(1–2):165–191

    MathSciNet  MATH  Google Scholar 

  • McKane AJ, Biancalani T, Rogers T (2014) Stochastic pattern formation and spontaneous polarisation: the linear noise approximation and beyond. Bull Math Biol 76(4):895–921

    MathSciNet  MATH  Google Scholar 

  • Othmer HG, Aldridge JA (1978) The effects of cell density and metabolite flux on cellular dynamics. J Math Biol 5(2):169–200

    MathSciNet  MATH  Google Scholar 

  • Pablo M, Ramirez SA, Elston TC (2018) Particle-based simulations of polarity establishment reveal stochastic promotion of Turing pattern formation. PLoS Comput Biol 14(3):e1006016

    Google Scholar 

  • Pavin N, Paljetak HČ, Krstić V (2006) Min-protein oscillations in Escherichia coli with spontaneous formation of two-stranded filaments in a three-dimensional stochastic reaction-diffusion model. Phys Rev E 73(2):021904

    Google Scholar 

  • Pikovsky AS, Kurths J (1997) Coherence resonance in a noise-driven excitable system. Phys Rev Lett 78(5):775

    MathSciNet  MATH  Google Scholar 

  • Press WH, Teukolsky SA, Vetterling WT, Flannery BP (1996) Numerical recipes in C, vol 2. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Reichenbach T, Mobilia M, Frey E (2006) Coexistence versus extinction in the stochastic cyclic Lotka–Volterra model. Phys Rev E 74(5):051907

    MathSciNet  Google Scholar 

  • Samoilov M, Plyasunov S, Arkin AP (2005) Stochastic amplification and signaling in enzymatic futile cycles through noise-induced bistability with oscillations. Proc Natl Acad Sci USA 102(7):2310–2315

    Google Scholar 

  • Schnoerr D, Sanguinetti G, Grima R (2017) Approximation and inference methods for stochastic biochemical kinetics—a tutorial review. J Phys A Math Theor 50(9):093001

    MathSciNet  MATH  Google Scholar 

  • Slaughter BD, Smith SE, Li R (2009) Symmetry breaking in the life cycle of the budding yeast. Cold Spring Harb Perspect Biol 1(3):a003384

    Google Scholar 

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

    Google Scholar 

  • Thomas P, Straube AV, Timmer J, Fleck C, Grima R (2013) Signatures of nonlinearity in single cell noise-induced oscillations. J Theor Biol 335:222–234

    MathSciNet  MATH  Google Scholar 

  • Toner DLK, Grima R (2013) Molecular noise induces concentration oscillations in chemical systems with stable node steady states. J Chem Phys 138(5):02B602

    Google Scholar 

  • Tostevin F, Howard M (2005) A stochastic model of Min oscillations in Escherichia coli and Min protein segregation during cell division. Phys Biol 3(1):1

    Google Scholar 

  • Ushakov OV, Wünsche H-J, Henneberger F, Khovanov IA, Schimansky-Geier L, Zaks MA (2005) Coherence resonance near a Hopf bifurcation. Phys Rev Lett 95(12):123903

    Google Scholar 

  • Van Kampen NG (1992) Stochastic processes in physics and chemistry, vol 1. Elsevier, Amsterdam

    MATH  Google Scholar 

  • Vellela M, Qian H (2007) A quasistationary analysis of a stochastic chemical reaction: Keizer’s paradox. Bull Math Biol 69(5):1727–1746

    MathSciNet  MATH  Google Scholar 

  • Wheatley E, Rittinger K (2005) Interactions between Cdc42 and the scaffold protein Scd2: requirement of SH3 domains for GTPase binding. Biochem J 388(1):177–184

    Google Scholar 

  • Wilkie J, Wong YM (2008) Positivity preserving chemical langevin equations. Chem Phys 353(1–3):132–138

    Google Scholar 

  • Wu C-F, Lew DJ (2013) Beyond symmetry-breaking: competition and negative feedback in GTPase regulation. Trends Cell Biol 23(10):476–483

    Google Scholar 

  • Xu B, Bressloff PC (2016) A PDE–DDE model for cell polarization in fission yeast. SIAM J Appl Math 76(5):1844–1870

    MathSciNet  MATH  Google Scholar 

  • Xu B, Jilkine A (2018) Modeling Cdc42 oscillation in fission yeast. Biophys J 114(3):711–722

    Google Scholar 

  • Zakharova A, Vadivasova T, Anishchenko V, Koseska A, Kurths J (2010) Stochastic bifurcations and coherencelike resonance in a self-sustained bistable noisy oscillator. Phys Rev E 81(1):011106

    Google Scholar 

  • Zakharova A, Feoktistov A, Vadivasova T, Schöll E (2013) Coherence resonance and stochastic synchronization in a nonlinear circuit near a subcritical Hopf bifurcation. Eur Phys J Spec Top 222(10):2481–2495

    Google Scholar 

Download references

Acknowledgements

BX is supported by the Robert and Sara Lumpkins Endowment for Postdoctoral Fellows in Applied and Computational Math and Statistics at the University of Notre Dame. HWK is supported by NSF Grant DMS-1620403. AJ is supported by NSF Grant DMS-1615800. AJ and BX acknowledge the assistance of the Notre Dame Center for Research Computing (CRC).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Bin Xu or Alexandra Jilkine.

Additional information

Publisher's Note

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

Appendices

Appendices

Scaling of the Deterministic and Stochastic Models

1.1 Dimensionalization of Deterministic Model

Note that our deterministic model (Xu and Jilkine 2018) used dimensionless units including all the concentrations and the cell size parameter L. Hence, for the stochastic simulation we need to reconstruct the dimensional model with parameters that have appropriate units. Recall that there are three types of variables with different units: length, time, and concentration. For concentrations, we have both membrane and cytosolic concentrations. The choices for these two concentrations are proteins per unit area (\(\upmu \hbox {m}^{-2}\)) and protein copy numbers per unit volume (\(\upmu \hbox {m}^{-3}\)).

Denote \(\tilde{x}\) and \(\tilde{t}\) by the length and time variables in dimensions. Denote \(\tilde{C}(\tilde{G})\) and \(\tilde{c}(\tilde{g})\) by the cytosolic and membrane concentrations of Cdc42(GEF) in dimensions. Consider the change of variables

$$\begin{aligned} \tilde{x}=\alpha x, \quad \tilde{t}=\beta t,\quad \tilde{C}(\tilde{x},\tilde{t})=\gamma C(x,t), \quad \tilde{c}_{1,2}(\tilde{t})=\gamma _m c_{1,2}(t). \end{aligned}$$
(34)

Here, \(\alpha \) represents the units of length. Since \(L=1\) corresponds to \(9\,\upmu \hbox {m}\), we set \(\alpha =9\,\upmu \hbox {m}\). Also, the Cdc42 dissociation rate \(k^-=1\) corresponds to \(4\min ^{-1}\) (Das et al. 2012); hence, we take \(\beta =1/4\min =15\,s\). We will discuss the units of concentrations (i.e., values of \(\gamma , \gamma _m\)) at a later time.

Using Eq. (34), we have

$$\begin{aligned} \frac{\partial C}{\partial t}=\frac{\beta }{\gamma }\frac{\partial \tilde{C}}{\partial \tilde{t}},\quad \frac{\partial C}{\partial x}=\frac{\alpha }{\gamma }\frac{\partial \tilde{C}}{\partial \tilde{x}},\quad \frac{\partial ^2C}{\partial x^2}=\frac{\alpha ^2}{\gamma }\frac{\partial ^2\tilde{C}}{\partial \tilde{x}^2}, \quad \frac{\mathrm{d}c_i}{\mathrm{d}t}=\frac{\beta }{\gamma _m}\frac{\mathrm{d}\tilde{c_i}}{\mathrm{d}\tilde{t}}. \end{aligned}$$
(35)

Hence, from Eq. (1), the diffusion equation and ODEs with dimensions are given by

$$\begin{aligned} \frac{\partial \tilde{C}}{\partial \tilde{t}}(\tilde{x},\tilde{t})&=\frac{D_C\alpha ^2}{\beta }\frac{\partial ^2\tilde{C}}{\partial \tilde{x}^2}, \end{aligned}$$
(36a)
$$\begin{aligned} \frac{\mathrm{d}\tilde{c}_1}{\mathrm{d}\tilde{t}}(\tilde{t})&=\left( \frac{k_0}{\gamma \beta }+\frac{k_{\mathrm{cat}}}{\gamma \gamma _m^2\beta }\tilde{c}_1^2\right) \tilde{g}_1\tilde{C}(0,\tilde{t})-\frac{k^-}{\beta }\tilde{c}_1(\tilde{t}),\end{aligned}$$
(36b)
$$\begin{aligned} \frac{\mathrm{d}\tilde{g}_1}{\mathrm{d}\tilde{t}}(\tilde{t})&=\frac{k^\mathrm{on_\mathrm{max}}/\beta }{1+\kappa \tilde{c}_1^2/\gamma _m^2}\frac{\gamma _m}{\gamma }\tilde{G}(0,\tilde{t})-\frac{k^\mathrm{off}}{\beta }\tilde{g}_1(\tilde{t}). \end{aligned}$$
(36c)

We list the dimensions of rates below.

$$\begin{aligned}&\tilde{k}_0=\frac{k_0}{\gamma \beta } \,\frac{1}{{\mathrm{[conc.]}}\times \sec }, \quad \tilde{k}_{\mathrm{cat}}=\frac{k_{\mathrm{cat}}}{\gamma \gamma _m^2 \beta } \frac{1}{{\mathrm{[conc.]}}_{\mathrm{m}}^2\times {\mathrm{[conc.]}}\times \sec }, \end{aligned}$$
(37a)
$$\begin{aligned}&\tilde{k}^\mathrm{on}_\mathrm{max}=\frac{k^{\mathrm{on}}_{\mathrm{max}}}{\beta }\frac{\gamma _m}{\gamma }\,\frac{{\mathrm{[conc.]}}_{{\mathrm{m}}}}{{\mathrm{[conc.]}}\times \sec }, \quad \tilde{\kappa }= \frac{\kappa }{\gamma _m^2}\,\frac{1}{{\mathrm{[ conc.]}}_{\mathrm{m}}^2}, \quad \tilde{k}^{{\mathrm{off}}}=\frac{k^{{\mathrm{off}}}}{\beta } \frac{1}{\sec }. \end{aligned}$$
(37b)

Here, [conc.] represents cytosolic concentration and \([\mathrm{conc.]}_m\) represents membrane concentration. It remains to determine the units of concentrations \(\gamma \) and \(\gamma _m\) and the corresponding order of magnitude.

1.2 Units of Concentration and Model Geometry

For the stochastic model geometry, we model the rod-shaped cell as a cylinder. In this case, each cell tip is a small disk with a radius r and a small thickness. This is the model geometry that is consistent with our one-dimensional deterministic model. For molecules attached to the cell membrane, we measure the concentration using molecular copy number per cross-sectional area since the thickness is fixed. For cytosolic concentration, we use the units of molecular copy number per volume (\(\upmu \hbox {m}^3\)).

Let \(\gamma _m=\frac{\#}{\mathrm{area}}=n_1 \upmu \hbox {m}^{-2}\) and \(\gamma =\frac{\#}{\mathrm{volume}}=n_c \upmu \hbox {m}^{-3}\). Here, \(n_1\) and \(n_c\) are numbers that represent the order of magnitude. The number of molecules in each compartment is given by

$$\begin{aligned} n_1= \tilde{c}_1 A,\quad m_1=\tilde{g}_1A,\quad n_c(\tilde{t})=\tilde{C}(\tilde{t})V,\quad m_c(\tilde{t})=\tilde{G}(\tilde{t})V. \end{aligned}$$
(38)

Note that in the last two equations, we drop the spatial dependence. This holds when diffusion in the cytosol is fast. The conservation equation is then given by

$$\begin{aligned} N_\mathrm{cdc}&=n_1+n_2+n_c \Rightarrow \tilde{c}_1 A+ \tilde{c}_2 A+ \tilde{C}(\tilde{t})V=\tilde{C}_\mathrm{tot}V,\nonumber \\&\Rightarrow \tilde{C}(\tilde{t})=\frac{\tilde{C}_\mathrm{tot}V/A-\tilde{c}_1 - \tilde{c}_2}{V/A}. \end{aligned}$$
(39)

Since we model the cell as a cylinder, the ratio V / A is given by

$$\begin{aligned} V\approx A\times (\alpha L), \end{aligned}$$
(40)

where \( \alpha =9\,\upmu \hbox {m}\). Equation (39) becomes

$$\begin{aligned} \tilde{C}(\tilde{t})=\frac{\tilde{C}_\mathrm{tot}\alpha L-\tilde{c}_1 - \tilde{c}_2}{\alpha L}. \end{aligned}$$
(41)

We now rewrite the above equation in terms of the dimensionless variables. Using the change of variables (34) to remove the tilde notations gives

$$\begin{aligned} C(t)\gamma&=\frac{C_{\mathrm{tot}}\gamma \alpha {L}-c_1\gamma _m - c_2\gamma _m}{\alpha L},\nonumber \\ \Rightarrow C(t)\frac{\alpha \gamma }{\gamma _m}&=\frac{C_{\mathrm{tot}} L(\alpha \gamma /\gamma _m)-c_1 - c_2}{L} . \end{aligned}$$
(42)

Suppose we choose \(\gamma \) and \(\gamma _m\) such that

$$\begin{aligned} \frac{\alpha \gamma }{\gamma _m}=1. \end{aligned}$$
(43)

Then, we can recover the conservation equation (6).

1.3 Derivation of Stochastic Rates

We now write down the equation for the number of molecules in each compartment using Eq. (38) and the differential equations (36). The ODEs for the number of Cdc42 at tip1 (\(n_1\)) and the number of GEF at tip1 (\(m_1\)) are given by

$$\begin{aligned} \frac{\mathrm{d}n_1}{\mathrm{d}\tilde{t}}(\tilde{t})&=\left( \frac{k_0}{\gamma \beta }+\frac{k_{\mathrm{cat}}}{\gamma \gamma _m^2\beta }\frac{n_1^2}{A^2}\right) m_1\frac{n_c(\tilde{t})}{V}-\frac{k^-}{\beta }n_1(\tilde{t}), \end{aligned}$$
(44a)
$$\begin{aligned} \frac{\mathrm{d}m_1}{\mathrm{d}\tilde{t}}(\tilde{t})&=\frac{k^\mathrm{on}_\mathrm{max}/\beta }{1+\kappa n_1^2/(\gamma _m^2A^2)}\frac{\gamma _m}{\gamma }\frac{A}{V}m_c(\tilde{t})-\frac{k^\mathrm{off}}{\beta }m_1(\tilde{t}). \end{aligned}$$
(44b)

Our goal is to derive the stochastic rates. The above ODEs imply the rescaling relations

$$\begin{aligned}&k'_0=\frac{\tilde{k}_0}{V}=\frac{k_0}{\gamma \beta V} , \end{aligned}$$
(45a)
$$\begin{aligned}&k'_\mathrm{cat}=\frac{\tilde{k}_{\mathrm{cat}}}{A^2V}=\frac{k_{\mathrm{cat}}}{\gamma \gamma _m^2 \beta }\frac{1}{A^2V} , \end{aligned}$$
(45b)
$$\begin{aligned}&k'_\mathrm{on}=\frac{\tilde{k}^\mathrm{on}A}{V}=\frac{k^\mathrm{on}_\mathrm{max}\gamma _m}{\gamma \beta }\frac{A}{ V},\end{aligned}$$
(45c)
$$\begin{aligned}&\kappa '=\frac{\tilde{\kappa }}{A^2}=\frac{\kappa }{\gamma _m^2A^2}, \end{aligned}$$
(45d)
$$\begin{aligned}&k'_{-}=\tilde{k}_{-}=\frac{k_{-}}{\beta } ,\quad k'_\mathrm{off}=\tilde{k}_\mathrm{off}=\frac{k^\mathrm{off}}{\beta }. \end{aligned}$$
(45e)

Here, the rates with the \('\) notation represent the rates we use in stochastic simulation, the rates with the tilde notation represent the rates in physical units, and the rates without any notation are dimensionless.

Using the relations \(V=A\alpha L\) and \(\alpha \gamma =\gamma _m\), we simplify the rescaling relation as

$$\begin{aligned}&k'_0=\frac{k_0}{\gamma \beta \alpha LA} =\frac{k_0}{\gamma _m\beta LA},\end{aligned}$$
(46a)
$$\begin{aligned}&k'_\mathrm{cat}=\frac{k_{\mathrm{cat}}}{\gamma _m^3A^3L\beta } , \end{aligned}$$
(46b)
$$\begin{aligned}&k'_\mathrm{on}=\frac{k^\mathrm{on}_\mathrm{max}\gamma _m}{\gamma \beta }\frac{A}{V}=\frac{k^\mathrm{on}_\mathrm{max}}{\beta L},\end{aligned}$$
(46c)
$$\begin{aligned}&\kappa '=\frac{\kappa }{\gamma _m^2A^2}, \end{aligned}$$
(46d)
$$\begin{aligned}&k'_{-}=\frac{k_{-}}{\beta } ,\quad k'_\mathrm{off}=\frac{k^\mathrm{off}}{\beta }. \end{aligned}$$
(46e)

Setting \(\gamma _m=602\), we recover Eq. (10). Note that the unit of all stochastic rate constants except for \(\kappa '\) is per \(1/4\min \) (\(15\,\)s) due to the factor \(1/\beta \) in (46).

1.4 Derivation of the Reduced ODE Model From the Stochastic Model

Remind that the numbers of molecules of Cdc42 at tip1, at tip2, and in the cytosol at time \(\tilde{t}\) are \(n_1(\tilde{t})\), \(n_2(\tilde{t})\), and \(n_c(\tilde{t})\). Similarly, the numbers of molecules of GEF at tip1, tip2, and in the cytosol at time \(\tilde{t}\) are \(m_1(\tilde{t})\), \(m_2(\tilde{t})\), and \(m_c(\tilde{t})\). Then, the species copy numbers are governed by the following stochastic equations

$$\begin{aligned} n_i(\tilde{t})= & {} n_i(0) + Y_{i,1}\left( \int _0^{\tilde{t}} \left( k_0'+k_\mathrm{cat}'n_i(\tilde{s})^2\right) m_i(\tilde{s})n_c(\tilde{s}) \,\hbox {d}\tilde{s}\right) - Y_{i,2}\left( \int _0^{\tilde{t}} k_{-}'n_i(\tilde{s})\,\hbox {d}\tilde{s}\right) ,\quad i=1,2, \nonumber \\ n_c(\tilde{t})= & {} n_c(0) +\sum _{i=1}^{2}\left[ - Y_{i,1}\left( \int _0^{\tilde{t}} \left( k_0'+k_\mathrm{cat}'n_i(\tilde{s})^2\right) m_i(\tilde{s})n_c(\tilde{s}) \,\hbox {d}\tilde{s}\right) + Y_{i,2}\left( \int _0^{\tilde{t}} k_{-}'n_i(\tilde{s})\,\hbox {d}\tilde{s}\right) \right] ,\nonumber \\ m_i(\tilde{t})= & {} m_i(0) + Y_{i,3}\left( \int _0^{\tilde{t}} \frac{k_\mathrm{on}'}{1+\kappa ' n_i(\tilde{s})^2} m_c(\tilde{s}) \,\hbox {d}\tilde{s}\right) - Y_{i,4}\left( \int _0^{\tilde{t}} k_\mathrm{off}' m_i(\tilde{s})\,\hbox {d}\tilde{s}\right) ,\quad i=1,2, \nonumber \\ m_c(\tilde{t})= & {} m_c(0) + \sum _{i=1}^{2}\left[ - Y_{i,3}\left( \int _0^{\tilde{t}} \frac{k_\mathrm{on}'}{1+\kappa ' n_i(\tilde{s})^2} m_c(\tilde{s}) \,\hbox {d}\tilde{s}\right) + Y_{i,4}\left( \int _0^{\tilde{t}} k_\mathrm{off}' m_i(\tilde{s})\,\hbox {d}\tilde{s}\right) \right] ,\nonumber \\ \end{aligned}$$
(47)

where \(Y_k\)’s are independent Poisson processes. Adding the equations for CDC42 (or GEF) in (47), the total number of molecules of Cdc42 (or GEF) is conserved as

$$\begin{aligned} \begin{aligned}&n_1(\tilde{t})+n_2(\tilde{t})+n_c(\tilde{t})=n_1(0)+n_2(0)+n_c(0) \equiv N_{\mathrm{cdc}}, \\&m_1(\tilde{t})+m_2(\tilde{t})+m_c(\tilde{t})=m_1(0)+m_2(0)+m_c(0) \equiv N_{gef}. \\ \end{aligned} \end{aligned}$$
(48)

Plugging in \(\tilde{t}=\beta t\) and using the change of variables (\(\tilde{s}=\beta s\)), (47) become

$$\begin{aligned} n_i(\beta t)= & {} n_i(0) + Y_{i,1}\left( \int _0^{t} \left( k_0'+k_\mathrm{cat}'n_i(\beta s)^2\right) m_i(\beta s)n_c(\beta s) \beta \,\hbox {d}s\right) \nonumber \\&- Y_{i,2}\left( \int _0^{t} k_{-}'n_i(\beta s)\beta \,\hbox {d}s\right) , \nonumber \\ n_c(\beta t)= & {} n_c(0) +\sum _{i=1}^{2}\left[ - Y_{i,1}\left( \int _0^{t} \left( k_0'+k_\mathrm{cat}'n_i(\beta s)^2\right) m_i(\beta s)n_c(\beta s) \beta \,\hbox {d}s\right) \right. \nonumber \\&\left. + Y_{i,2}\left( \int _0^{t} k_{-}'n_i(\beta s)\beta \,\hbox {d}s\right) \right] ,\nonumber \\ m_i(\beta t)= & {} m_i(0) + Y_{i,3}\left( \int _0^{t} \frac{k_\mathrm{on}'}{1+\kappa ' n_i(\beta s)^2} m_c(\beta s) \beta \,\hbox {d}s\right) - Y_{i,4}\left( \int _0^{t} k_\mathrm{off}' m_i(\beta s) \beta \,\hbox {d}s\right) ,\nonumber \\ m_c(\beta t)= & {} m_c(0) + \sum _{i=1}^{2}\left[ - Y_{i,3}\left( \int _0^{t} \frac{k_\mathrm{on}'}{1+\kappa ' n_i(\beta s)^2} m_c(\beta s) \beta \,\hbox {d}s\right) + Y_{i,4}\left( \int _0^{t} k_\mathrm{off}' m_i(\beta s) \beta \,\hbox {d}s\right) \right] ,\nonumber \\ \end{aligned}$$
(49)

for \(i=1,2\). Assuming all species copy numbers are of the same order, we normalize the species copy numbers by a scaling parameter \(N=\gamma _m A=602A\). Define normalized variables as

$$\begin{aligned} c_i^N(t)=\frac{n_i(\beta t)}{N},&\quad&C^N(t)=\frac{n_c(\beta t)}{NL},\qquad g_i^N(t)=\frac{m_i(\beta t)}{N},\qquad G^N(t)=\frac{m_c(\beta t)}{NL}, \nonumber \\ \end{aligned}$$
(50)

for \(i=1,2\).

We express the stochastic rate constants in (46) using N as the following:

$$\begin{aligned} \begin{aligned} k_0'&= \frac{k_0}{N \beta L},\quad k_\mathrm{cat}' = \frac{k_{\mathrm{cat}}}{N^3 \beta L},\quad k_\mathrm{on}' = \frac{k^\mathrm{on}_\mathrm{max}}{\beta L}, \quad \kappa ' = \frac{\kappa }{N^2},\quad k_{-}'=\frac{k_{-}}{\beta },\quad k_\mathrm{off}'=\frac{k^\mathrm{off}}{\beta }. \end{aligned} \end{aligned}$$
(51)

Then, plugging in the normalized variables and scaled rate constants in (50) and (51), (49) becomes

$$\begin{aligned} c_i^N(t)= & {} c_i^N(0) + \frac{1}{N} Y_{i,1}\left( \int _0^t N\left( k_0+k_{\mathrm{cat}}c_i^N(s)^2\right) g_i^N(s)C^N(s) \,\hbox {d}s\right) \nonumber \\&- \frac{1}{N} Y_{i,2}\left( \int _0^t N k_{-}c_i^N(s)\,\hbox {d}s\right) ,\nonumber \\ C^N(t)= & {} C^N(0) +\sum _{i=1}^2 \frac{1}{NL} \left[ -Y_{i,1}\left( \int _0^t N\left( k_0+k_{\mathrm{cat}}c_i^N(s)^2\right) g_i^N(s)C^N(s) \,\hbox {d}s\right) \right. \nonumber \\&\left. + Y_{i,2}\left( \int _0^t N k_{-}c_i^N(s)\,\hbox {d}s\right) \right] ,\nonumber \\ g_i^N(t)= & {} g_i^N(0) + \frac{1}{N} Y_{i,3}\left( \int _0^t \frac{Nk^\mathrm{on}_\mathrm{max}}{1+\kappa c_i^N(s)^2} G^N(s) \,\hbox {d}s\right) - \frac{1}{N} Y_{i,4}\left( \int _0^t N k^\mathrm{off} g_i^N(s)\,\hbox {d}s\right) ,\nonumber \\ G^N(t)= & {} G^N(0) + \sum _{i=1}^2 \frac{1}{NL} \left[ -Y_{i,3}\left( \int _0^t \frac{Nk^\mathrm{on}_\mathrm{max}}{1+\kappa c_i^N(s)^2} G^N(s) \,\hbox {d}s\right) + Y_{i,4}\left( \int _0^t N k^\mathrm{off} g_i^N(s)\,\hbox {d}s\right) \right] ,\nonumber \\ \end{aligned}$$
(52)

for \(i=1,2\). The strong law of large numbers states that, for a unit Poisson process Y, \(\frac{1}{N}Y(Nu)\rightarrow u\) almost surely as \(N\rightarrow \infty \). Therefore, assuming that \(c_i^N(0)\rightarrow c_i(0)\), \(C^N(0)\rightarrow C(0)\), \(g_i^N(0)\rightarrow g_i(0)\), and \(G^N(0)\rightarrow G(0)\) as \(N\rightarrow \infty \), \(X^N\equiv \left( c_1^N,c_2^N,C^N, g_1^N,g_2^N,G^N\right) \) converges to the limit which is a solution of

$$\begin{aligned} \begin{aligned} c_i(t)&= c_i(0) + \int _0^t \left[ \left( k_0+k_{\mathrm{cat}}c_i(s)^2\right) g_i(s) C(s) - k_{-}c_i(s) \right] \,\hbox {d}s, \\ C(t)&= C(0) + \frac{1}{L} \int _0^t \sum _{i=1}^{2} \left[ -\left( k_0+k_{\mathrm{cat}}c_i(s)^2\right) g_i(s) C(s) + k_{-}c_i(s) \right] \,\hbox {d}s, \\ g_i(t)&= g_i(0) + \int _0^t \left[ \frac{k^\mathrm{on}_\mathrm{max}}{1+\kappa c_i(s)^2} G(s) - k^\mathrm{off} g_i(s) \right] \,\hbox {d}s, \\ G(t)&= G(0) + \frac{1}{L} \int _0^t \sum _{i=1}^{2} \left[ -\frac{k^\mathrm{on}_\mathrm{max}}{1+\kappa c_i(s)^2} G(s) + k^\mathrm{off} g_i(s) \right] \,\hbox {d}s,\\ \end{aligned} \end{aligned}$$
(53)

for \(i=1,2\) as \(N\rightarrow \infty \). Note that \(c_i\) and \(g_i\) are the solution of the reduced ODE model for Cdc42 and GEF given in (7) and the total concentrations of Cdc42 and GEF are conserved as

$$\begin{aligned} \begin{aligned} \frac{c_1(t)+c_2(t)}{L}+C(t) \equiv C_{\mathrm{tot}}, \\ \frac{g_1(t)+g_2(t)}{L}+G(t) \equiv G_{\mathrm{tot}}, \end{aligned} \end{aligned}$$
(54)

which is consistent with (6). Setting \(N=602 A\) and \(t=0\) in (50), we obtain that

$$\begin{aligned} n_i(0)= & {} 602A\, c_i(0),\quad n_c(0) = 602AL\, C(0),\quad m_i(0) = 602A\, g_i(0),\nonumber \\ m_c(0)= & {} 602AL\, G(0), \end{aligned}$$
(55)

for \(i=1,2\). Then, (55), (48), and (54) provide how \(C_{\mathrm{tot}}\) and \(N_{\mathrm{cdc}}\) (\(G_{\mathrm{tot}}\) and \(N_{gef}\)) are related as follows:

$$\begin{aligned} \begin{aligned} N_{\mathrm{cdc}} = 602A(c_1(t)+c_2(t)+LC(t)) = 602AL\, C_{\mathrm{tot}}, \\ N_{gef} = 602A(g_1(t)+g_2(t)+LG(t)) = 602AL\, G_{\mathrm{tot}}. \end{aligned} \end{aligned}$$
(56)

Power Spectrum Analysis

1.1 Discrete Fourier Transform

Here, we use the fast Fourier transform to estimate the temporal power spectrum density of a discrete time series \(\{n_1(t_i)\}_{i=1}^{N_t}\) generated by Gillespie’s algorithm. We subtract the steady state from the time series and replace \(n_1(t_i)\) by \(n_1(t_i)-n_1^*\). The default time interval is given by \(t=0:\Delta t:t_\mathrm{end}\) with \(t_\mathrm{end}=200\) and a uniform time step \(\Delta t=0.001\). For the frequency interval, we set the sample frequency \(f_s=1/\Delta t=1000\) and define the frequency domain by

$$\begin{aligned} ff=[0:N_t-1]*\Delta f,\quad \Delta f=\frac{f_s}{N_t}. \end{aligned}$$
(57)

Here, \(N_t\) is the length of the time series. For each frequency \(f_k\), the discrete Fourier transform of \(n_1(t)\) is given by

$$\begin{aligned} \tilde{n}_1(k):=\sum _{j=0}^{N_t-1}n_1(t_j)\mathrm{e}^{-2\pi ijk/N}. \end{aligned}$$
(58)

Define the one-sided power spectrum P(k) by Press et al. (1996); Toner and Grima (2013)

$$\begin{aligned} P(k) = {\left\{ \begin{array}{ll} \displaystyle \frac{|\tilde{n}_1(k)|^2}{N_t^2}, &{} k=0,\,N_t/2,\\ \displaystyle \frac{|\tilde{n}_1(k)|^2+|\tilde{n}_1({N_t-k})|^2}{N_t^2}, &{} 1\le k\le N_t/2-1. \end{array}\right. } \end{aligned}$$
(59)

The power spectral density (PSD) over the frequency range \((f_k-f_s/(2N_t),f_k+f_s/(2N_t))\) is given by

$$\begin{aligned} S(f_k\pm \frac{1}{2}f_s)=\frac{P(k)}{\Delta f}= \frac{|\tilde{n}_1(k)|^2+|\tilde{n}_1({N_t-k})|^2}{N_tf_s}. \end{aligned}$$
(60)

The corresponding PSD over the angular frequency range \((\omega _k-\Delta \omega /2, \omega _k+\Delta \omega /2)\) is

$$\begin{aligned} S(\omega _k\pm \frac{1}{2}\Delta \omega )=\frac{P(k)}{\Delta \omega }. \end{aligned}$$
(61)

Here, the angular frequency increment is \(\Delta \omega =2\pi f_s/N_t\). Finally, we average the PSD of a single realization r over R realizations to obtain a numerical approximation for the PSD. Denote the PSD of a single realization r by \(S^r(\omega _k)\). The averaged PSD is given by

$$\begin{aligned} S(\omega _k)=\frac{1}{R}\sum _{r=1}^RS^r(\omega _k). \end{aligned}$$
(62)

1.2 MATLAB Code

figure a

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, B., Kang, HW. & Jilkine, A. Comparison of Deterministic and Stochastic Regime in a Model for Cdc42 Oscillations in Fission Yeast. Bull Math Biol 81, 1268–1302 (2019). https://doi.org/10.1007/s11538-019-00573-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-019-00573-5

Keywords

Navigation