Skip to main content
Log in

Cell cycle dynamics: clustering is universal in negative feedback systems

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

Abstract

We study a model of cell cycle ensemble dynamics with cell–cell feedback in which cells in one fixed phase of the cycle \(S\) (Signaling) produce chemical agents that affect the growth and development rate of cells that are in another phase \(R\) (Responsive). For this type of system there are special periodic solutions that we call \(k\)-cyclic or clustered. Biologically, a \(k\)-cyclic solution represents \(k\) cohorts of synchronized cells spaced nearly evenly around the cell cycle. We show, under very general nonlinear feedback, that for a fixed \(k\) the stability of the \(k\)-cyclic solutions can be characterized completely in parameter space, a 2 dimensional triangle \(T\). We show that \(T\) is naturally partitioned into \(k^2\) sub-triangles on each of which the \(k\)-cyclic solutions all have the same stability type. For negative feedback we observe that while the synchronous solution (\(k=1\)) is unstable, regions of stability of \(k \ge 2\) clustered solutions seem to occupy all of \(T\). We also observe bi-stability or multi-stability for many parameter values in negative feedback systems. Thus in systems with negative feedback we should expect to observe cyclic solutions for some \(k\). This is in contrast to the case of positive feedback, where we observe that the only asymptotically stable periodic orbit is the synchronous solution.

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
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

References

  • Bier M, Bakker B, Westerhoff H (2000) How yeast cells synchronize their glycolytic oscillations: a perturbation analytic treatment. Biophys J 78:1087–1093

    Article  Google Scholar 

  • Boczko E, Cooper T, Gedeon T, Mischaikow K (2005) Structure theorems and the dynamics of nitrogen catabolite repression in yeast. PNAS 102:5647–5652

    Article  Google Scholar 

  • Boczko E, Stowers C, Gedeon T, Young T (2010) Ode, rde and sde models of cell cycle dynamics and clustering in yeast. J Biol Dyn 4:328–345

    Article  MathSciNet  Google Scholar 

  • Chen Z, Odstrcil E, Tu B, McKnight S (2007) Restriction of DNA replication to the reductive phase of the metabolic cycle protects genome integrity. Science 316:1916–1919

    Article  Google Scholar 

  • Diekmann O, Gyllenberg M, Verduyn Lunel SM (1993) A cell-cycle model revisited. CWI. Department of Analysis, Algebra and Geometry [AM] Report No. 9305, pp 1–18. http://www.helsinki.fi/~mgyllenb/publications.html

  • Diekmann O, Heijmans H, Thieme H (1984) On the stability of the cell size distribution. J Math Biol 19:227–248

    Article  MATH  MathSciNet  Google Scholar 

  • Finn R, Wilson R (1954) Population dynamic behavior of the chemostat system. Agric Food Chem 2:66–69

    Article  Google Scholar 

  • Hannsgen K, Tyson J (1985) Stability of the steady-state size distribution in a model of cell growth and division. J Math Biol 22:293–301

    Article  MATH  MathSciNet  Google Scholar 

  • Henson M (2005) Cell ensemble modeling of metabolic oscillations in continuous yeast cultures. Comput Chem Eng 29:645–661

    Article  MathSciNet  Google Scholar 

  • Keulers M, Satroutdinov A, Sazuki T, Kuriyama H (1996) Synchronization affector of autonomous short period sustained oscillation of Saccharomyces cerevisiae. Yeast 12:673–682

    Article  Google Scholar 

  • Kilpatrick Z, Ermentrout B (2011) Sparse gamma rhythms arising through clustering in adapting neuronal networks. PLoS Comput Biol 7(11):e1002281.

  • Kjeldsen T, Ludvigsen S, Diers I, Balshmidt P, Sorensen A, Kaarshold N (2002) Engineering-enhanced protein secretory expression in yeast with applications to insulin. J Biol Chem 277:18245–18248

    Article  Google Scholar 

  • Klevecz R (1976) Quantized generation time in mammalian cells as an expression of the cellular clock. PNAS 73:4012–4016

    Article  Google Scholar 

  • Klevecz R, Murray D (2001) Genome wide oscillations in expression. Mol Biol Rep 28:73–82

    Article  Google Scholar 

  • Kopmann A, Diekmann H, Thoma M (1998) Oxygen, ph value, and carbon source induced changes of the mode of oscillation in synchronous continuous culture of Saccharomyces cervisiae. Biotechnol Bioeng 63:410–417

    Google Scholar 

  • Kuenzi M, Fiechter A (1969) Changes in carbohydrate composition and trehalose activity during the budding cycle of Saccharomyces cerevisiae. Arch Microbiol 64:396–407

    Google Scholar 

  • Mauroy A, Sepulchre R (2008) Clustering behaviors in networks of integrate-and-fire oscillators. Chaos 18(3):037122

  • Meyenburg HV (1969) Energetics of the budding cycle of Saccharomyces cerevisiae during glucose limited aerobic growth. Arch Microbiol 66:289–303

    Google Scholar 

  • Monte SD, d’Ovidio F, Danø S, Sørensen P (2007) Dynamical quorum sensing: population density encoded in cellular dynamics. PNAS 104:18377–18381

  • Pr PP (2003) Oscillatory metabolism of Saccharomyces cerevisiae: an overview of mechanisms and models. Biotechnol Adv 21:183–192

    Article  Google Scholar 

  • Robertson J, Stowers C, Boczko E, Johnson C (2008) Real-time luminescence monitoring of cell-cycle and respiratory oscillations in yeast. PNAS 105:17988–93

    Article  Google Scholar 

  • Rotenberg M (1977) Selective synchrony of cells of differing cycle times. J Theor Biol 66:389–398

    Article  MathSciNet  Google Scholar 

  • Slavov N, Botstein D (2011) Coupling among growth rate response, metabolic cycle, and cell division cycle in yeast. Mol Biol Cell 22:1999–2009

    Article  Google Scholar 

  • Stowers C, Young T, Boczko E (2011) The structure of populations of budding yeast in response to feedback. Hypotheses Life Sci 1:71–84

    Google Scholar 

  • Tu B, Kudlicki A, Rowicka M, McKnight S (2005) Logic of the yeast metabolic cycle: temporal compartmentation of cellular processes. Science 310:1152–1158

    Article  Google Scholar 

  • Uchiyama K, Morimoto M, Yokoyama Y, Shioya S (1996) Cell cycle dependency of rice \(\alpha \)-amylase production in a recombinant yeast. Biotechnol Bioeng 54:262–271

    Article  Google Scholar 

  • Von Meyenburg K (1973) Stable synchronous oscillations in continuous cultures of S. cerevisiae under glucose limitation. In: Biological and Biochemical Oscillators. Academic Press, NY

  • Young T, Fernandez B, Buckalew R, Moses G, Boczko E (2012) Clustering in cell cycle dynamics with general responsive/signaling feedback. J Theor Biol 292:103–115

    Article  MathSciNet  Google Scholar 

  • Zhu G, Zamamiri A, Henson M, Hjortsø M (2000) Model predictive control of continuous yeast bioreactors using cell population balance models. Chem Eng Sci 55:6155–6167

Download references

Acknowledgments

Richard Buckalew (Ohio U.) provided essential help with some of the images. We thank Kara Finley (Ohio U.) for helpful comments. E.B., T.Y. and this work were partially supported by the NIH-NIGMS grant R01GM090207.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nathan Breitsch.

Appendix A: verification for \(k=3\)

Appendix A: verification for \(k=3\)

Referring to the labeling in Fig. 6, we present calculations verifying rigorously the conclusions of the numerically generated plot in Fig. 9 for the case \(k=3\).

Regions 2, 6, 7:

In Fig. 9 for the case \(k=3\) we see that regions 2 and 7 have neutral stability of the 3 cohort cyclic solution and in region 6 it is stable. These three regions are covered by Theorem 2.4.

Region 1:

We wish to confirm that for all parameter values in this region, the \(3\) cohort cyclic solution is neutrally stable.

For parameter values in region 1 both \(x_1\) and \(x_2\) are initially in \(S\). The order of events is that first \(x_2\) reaches \(s\), then it reaches \(r\) and finally it reaches \(1\). The speed of \(x_2\) will be \(1\) until it reaches \(r\). While in \(R\) the cohort \(x_2\) will experience feedback due to \(x_1, x_1 \in S\), and so it’s speed will be \(1+\beta _2\). The time required for \(x_2\) to reach \(1\) is thus:

$$\begin{aligned} t^* = r-x_2 + \frac{1}{1+\beta _2} (1-r). \end{aligned}$$

The positions of \(x_1\) and \(x_1\) at this time will be:

$$\begin{aligned} x_1(t^*) = t^* \quad x_1(t^*) = x_1 + t^*. \end{aligned}$$

We observe that this map is affine in the variables \(x_1\), \(x_2\). Thus the derivative of the map \(F\) in a neighborhood of the initial condition of the cyclic solution is:

$$\begin{aligned} DF = \left( \begin{array}{c@{\quad }c} 0 &{} -1 \\ 1 &{} -1 \end{array} \right) . \end{aligned}$$
(7.1)

The eigenvalues of this matrix are \(-\frac{1}{2} \pm i \frac{\sqrt{3}}{2}\) which implies that the solution is linearly neutrally stable. Since the map is affine in a neighborhood of the initial condition, the cyclic solution is neutral.

Region 3:

In this region we wish to confirm that the 3 cyclic solution is stable.

For these parameter values first \(x_2\) reaches \(r\), then \(x_1\) reaches \(s\) and finally \(x_2\) reaches \(1\). Thus \(x_2\) will experience feedback of \(\beta _2\) for the times \(r-x_2 < t < s - x_1\). For \(t > s-x_1\) it will be subject to feedback \(\beta _1\). The trajectory of \(x_2\) is thus:

$$\begin{aligned} x_2(t) = {\left\{ \begin{array}{ll} x_2 + t &{}\quad \text {for }\quad 0 \le t \le r - x_2, \\ r + (1+\beta _2)(t - r+x_2) &{}\quad \text {for }\quad r-x_2 \le t \le s - x_1, \\ r + (1+\beta _2)(s - x_1 - r+x_2) &{}\quad \text {for }\quad s-x_1 \le t \le t^* .\\ \quad + (1 + \beta _1)(t - s + x_1) \end{array}\right. } \end{aligned}$$

It follows that \(t^*\) satisfies:

$$\begin{aligned} r + (1+\beta _2)(s - x_1 - r+x_2) + (1 + \beta _1)(t ^*- s + x_1) = 1 \end{aligned}$$

and so

$$\begin{aligned} t^* = \frac{\beta _2 - \beta _1}{1+ \beta _1} x_1 - \frac{1 + \beta _2}{1 + \beta _1} x_2. \end{aligned}$$

The local affine map is given by:

$$\begin{aligned} x_1(t^*) = t^*, \qquad x_1(t^*) = x_1 + t^*, \end{aligned}$$

which gives us:

$$\begin{aligned} DF = \left( \begin{array}{c@{\quad }c} \frac{\beta _2 - \beta _1}{1+ \beta _1} &{} - \frac{1 + \beta _2}{1 + \beta _1} \\ 1+ \frac{\beta _2 - \beta _1}{1+ \beta _1} &{} - \frac{1 + \beta _2}{1 + \beta _1} \end{array} \right) = \left( \begin{array}{c@{\quad }c} \frac{\beta _2 - \beta _1}{1+ \beta _1} &{} - \frac{1 + \beta _2}{1 + \beta _1} \\ \frac{1+ \beta _2}{1+ \beta _1} &{} - \frac{1 + \beta _2}{1 + \beta _1} \end{array} \right) . \end{aligned}$$
(7.2)

The characteristic polynomial for this matrix simplifies to:

$$\begin{aligned} \lambda ^2 + \lambda + \frac{1 + \beta _2 }{1+\beta _1} = 0. \end{aligned}$$

One can check that the roots of this polynomial are less than 1 in modulus if and only if \(\beta _2 < \beta _1\). This is the case for monotone negative feedback and so the cyclic solution is stable.

Region 4:

We confirm that the 3-cyclic solution is neutrally stable in this region. We first compute the time it takes \(x_2\) to reach 1; \(x_2\) travels with speed \((1+\beta _1)\) until time \(t=s-x_1\), then travels at speed \((1+\beta _1)\), allowing us to set up the equation:

$$\begin{aligned} 1 = x_2 + (1+\beta _2)(s-x_1) + (1+\beta _2)(t^* - s + x_1). \end{aligned}$$

Solving, we find \(t^* = \displaystyle \frac{1+(\beta _1 - \beta _2)s}{1+\beta _1} - \displaystyle \frac{1}{1+\beta _1}x_2 + \displaystyle \frac{\beta _2 - \beta _1}{1+\beta _1}x_1\).

cohort \(x_1\) travels at speed 1; \(x_1\) travels at speed 1 until \(t=r-x_1\), then at speed \((1+\beta _1)\):

$$\begin{aligned} x_1(t^*) = t^* \quad x_1(t^*) = r+(1+\beta _1)(t^* - r + x_1) \end{aligned}$$

Gathering for convenience the constant terms as a single \(C\), \(x_1(t^*) = (1+\beta _1)x_1 + (\beta _2 - \beta _1)x_1 - x_2 + C= (1+\beta _2)x_1 - x_2 + C\); thus

$$\begin{aligned} DF = \left( \begin{array}{c@{\quad }c} \frac{\beta _2 - \beta _1}{1+\beta _1} &{} -\frac{1}{1+\beta _1} \\ 1+\beta _2 &{} -1 \end{array} \right) . \end{aligned}$$

The characteristic polynomial of DF is given by \(\lambda ^2 + \lambda (\frac{\beta _1 - \beta _2}{1+\beta _1} + 1) + 1\). To simplify the following expressions, we denote \(w = \frac{\beta _2 - \beta _1}{1+\beta _1}\). The roots of the polynomial are then \(\lambda _{\pm }=\frac{-w-1 \pm \sqrt{(w+1)^2 - 4}}{2}\). Observing that \(w<1\) iff \(-1<\beta _2\), which is true, we see that the roots are complex. We compute the norm of \(\lambda _{+}\) as \(\sqrt{\frac{1}{4}w^2 + \frac{1}{4}(4-w^2)} = 1\)

Region 5:

We wish to confirm that the 3 cyclic solution is neutrally stable in this region.

For this region \(x_1\) reaches \(s\) before \(x_1\) or \(x_2\) reach \(r\). This implies that no cell actually experiences any feedback. The map \(F\) in a neighborhood of the cyclic solution is thus given by the trivial affine expression:

$$\begin{aligned} x_1(t^*) = t^* = 1-x_2, \quad x_1(t^*) = x_1 +1 - x_2. \end{aligned}$$

The derivative \(DF\) is then the same as (6.1) and the solution is thus neutrally stable.

Region 8:

For parameter values in this region \(x_2\) travels with speed \(1+\beta _1\) until time \(t=s\), then at speed 1. The time it takes to reach 1 can thus be calculated:

$$\begin{aligned} 1 = x_2 + (1+\beta _1)s+t^* - s \rightarrow 1 - x_2 - \beta _1 s = t^*. \end{aligned}$$

Clearly \(x_1(t^*) = t^*\); bearing in mind that \(x_1\) enters \(R\) before \(x_1\) leaves \(S\), we calculate:

$$\begin{aligned} x_1(t^*) = r+ (1+\beta _1)(s-r+x_1)+t^* - s. \end{aligned}$$

This gives rise to the matrix

$$\begin{aligned} DF = \left( \begin{array}{c@{\quad }c} 0 &{} -1\\ 1+\beta _1 &{} -1 \end{array} \right) , \end{aligned}$$

yielding a characteristic polynomial of \(\lambda ^2 + \lambda + (1+\beta _ 1)\) and roots of \(\lambda _{\pm } = \frac{-1 \pm \sqrt{-3 - 4\beta _1}}{2}\). If the roots are real, the maximum root (in absolute value), \(\lambda _{-}\), is clearly contained in the interval \((-1,-\frac{1}{2})\); if the roots are complex, their norm can be explicitly calculated as \(\sqrt{1+\beta _1}\).

Region 9:

cohort \(x_2\) travels at rate \((1+\beta _1)\) until time \(t=s\), whereafter it travels with speed 1. We can then calculate the time it takes \(x_2\) to reach 1:

$$\begin{aligned} 1 = x_2 + (1+\beta _1)s+t^* -s, \end{aligned}$$

yielding \(t^* = 1 - x_2 - \beta _1 s\).

cohort \(x_1\) experiences feedback exactly when \(x_2\) does, and thus travels the same distance. The local affine map is given by:

$$\begin{aligned} x_1(t^*) = t^* \quad x_1(t^*) = x_1+1 -x_2. \end{aligned}$$

The resultant map DF is exactly that in (A.1), so the solution is neutrally stable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Breitsch, N., Moses, G., Boczko, E. et al. Cell cycle dynamics: clustering is universal in negative feedback systems. J. Math. Biol. 70, 1151–1175 (2015). https://doi.org/10.1007/s00285-014-0786-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-014-0786-7

Mathematics Subject Classification

Navigation