Abstract
The cell cycle duration is a variable cellular phenotype that underlies long-term population growth and age structures. By analyzing the stationary solutions of a branching process with heritable cell division times, we demonstrate existence of a phase transition, which can be continuous or first-order, by which a non-zero fraction of the population becomes localized at a minimal division time. Just below the transition, we demonstrate coexistence of localized and delocalized age-structure phases, and power law decay of correlation functions. Above it, we observe self-synchronization of cell cycles, collective divisions, and slow ‘aging’ of population growth rates.
The duration of a cell-cycle, or inter-division time (IDT), is a fluctuating quantity in cellular populations, and its statistical properties are thought to result from biological mechanisms that regulate cell growth and division [1–5]. In most observations on mammalian cells, and in a subset of bacterial experiments, positive mother-daughter correlations of IDTs have been measured (see Table S1 in [6]). In general, heritability of a trait (i.e. positive parent-offspring correlation) enables selection to act on the trait distribution in a population to increase the long-term population growth rate. In this Letter, we take up the question of how selection and cell cycle heritability interact to determine long-term population dynamics, a fundamental step toward understanding how evolution has shaped cell cycle control mechanisms. Our work makes connections between dynamics of age-structured populations [7] and error-threshold phenomena of evolutionary theory [8, 9], and is applicable in experimental analyses of cellular population dynamics.
Model and stationary solutions.
We model a proliferating population by a branching process in which K (τ, τ′) is the transition probability density from τ′ to τ, where τ′ and τ are the parent and offspring IDTs, respectively. We focus on the analytically tractable Lebowitz-Rubinow model [10], which uses the transition kernel
| (1) |
where δ (·) is the Dirac delta function showing accurate inheritance of the parent’s IDT, h represents the heritability of IDTs in the model (0 ≤ h < 1), and k (τ) is a probability density function on the interval [τ0, ∞), with 0 < τ0 < ∞. A cell produces offsprings at every division, where is independently drawn from a fixed probability distribution and its average is denoted by z > 0. The dynamics of cell divisions in the population are governed by
| (2) |
where ndiv (τ; t) dτdt is the expected number of dividing cells (cells at the termination of cell-cycle) between times t and t + dt with IDT between τ and τ + dτ [11]. The expected number of divisions occurring between t and t + dt is given by , and the time-dependent population growth rate is defined by , where N (t) is the expected population size at time t (see [6] for detailed derivations).
The steady-state, exponentially growing solution of (2) is characterized by the time-independent growth rate Λ and the probability density pdiv (τ) of IDTs of dividing cells, which can be found by substituting the form ndiv (τ; t) = Ndiv (t) pdiv (τ), where Ndiv (t) is proportional to eΛt, yielding
| (3) |
Using (1) and solving the above equation one obtains
| (4) |
where
| (5) |
which is a valid solution if there exists Λ such that qh,Λ(τ) ≥ 0 for all τ ∈ [τ0, ∞), and which normalizes pdiv(τ), i.e. Λ is the unique real solution of Q(h, Λ) = 1, where
| (6) |
If z = 1, for example, then Λ = 0 and pdiv (τ) = k (τ) for any 0 ≤ h < 1; this can be realized as an isolated single cell, where offspring at each division event are removed.
In the absence of IDT heritability (h = 0), one recovers the well-known result pdiv (τ) = ze−Λτ k (τ) where Λ is the unique real root of the integral equation . For h > 0, one additionally must have 1 − hze−Λτ > 0 for all τ in the support of k(τ) to ensure qh,Λ (τ) ≥ 0. If k(τ) > 0 for all τ > τ0, we find
| (7) |
(see [6] for k(τ) with bounded support). Since Q(h, Λ) is a monotonically decreasing function of Λ tending to zero as Λ → ∞, Q(h, Λ) = 1 has a unique root Λ = Λ(h) provided that Q(h, ω0 (h)) ≥ 1. This condition holds for h < hc, where hc is a heritability threshold defined by
| (8) |
such that for h > hc, Q(h, ω0 (h)) < 1 and Q(h, Λ) = 1 does not admit a real root Λ. For h > hc, the solution pdiv (τ) given in (4) is incomplete, as there is missing probability 1 − Q. For z > 1, the full solution is
| (9) |
and substitution into (3) yields for the steady-state growth rate when h > hc [6].
Further analysis of (8) shows that a heritability threshold hc < 1 exists if and only if converges [6]; e.g. if k(τ) ~ (τ − τ0)γ near τ0 for some γ > 0. The steady state population growth rate Λ(h) qualitatively changes as h crosses the threshold: it depends on k (τ) for h < hc, and becomes independent of it for h > hc. The expected number of offspring having the same IDT as their parent is hz, and only parents with τ′ = τ0 can generate offspring with τ = τ0. Thus, the fraction of the population with IDT τ0 grows with rate . If we add a small fraction of τ0 cells to a population, they go extinct if hz < 1, while if hz > 1, they can constitute a giant cluster in the population’s genealogy. The subpopulation localized at τ = τ0 will be outcompeted by the rest of the population for z−1 < h < hc, with Λ(h) determined by (6); or it will dominate the population for h > hc, and thus dictate its growth rate to be . In Fig. 1A, we show a range of examples of k(τ) which admit a threshold hc. Increasing h from 0 to 1, Λ(h) increases monotonically with a shallow slope, while past hc, the slope of the growth rate changes markedly.
FIG. 1.

Stationary solutions and their dependence on the heritability parameter h. Solid curves show results using z = 2 for (A) population growth rate, Λ(h); (B) reciprocal mean block size, ; and (C) reciprocal ancestral mean IDT, . Dotted line in (B,C) indicates the result for isolated single cells (i.e. z = 1). The IDT distribution k (τ) (shown at inset of A) is chosen to be a gamma distribution, shifted by τ0 = 0.2, with shape parameters α = 1.5 (blue), 4 (orange), and 20 (green); for τ ≥ τ0. For the mean IDT to be 1, the scale parameter θ is chosen as τ0 +αθ = 1. hc = 0.68, 0.60 and 0.58 respectively for α = 1.5, 4, and 20.
Localization phase transition of population age structure.
We now show that the threshold behavior identified above constitutes a phase transition in the strict sense. We map the population to a statistical mechanical ensemble, as follows. From the viewpoint of single cell lineages – i.e. the history of an individual and all of its ancestors – an age-structured population is an ensemble of trajectories: the sequence of IDTs along a lineage (…, τi−1, τi, τi+1, …) is analogous to a microscopic state of a large system (e.g. configuration of spins on a lattice, conformation of a polymer in space, etc.); while a single ancestral cell division τi specifies the state of a single component (e.g a spin, or a monomer) [6, 13].
To analyze the structure of lineages observed above and below hc, we consider the number of generations with which the same IDT is consecutively inherited, which we denote by m and call the “block size”. The probability distribution of m over lineages, plin(m), is analogous to a correlation function, and its mean measures the typical correlation length. From (3) one can infer that the joint probability distribution of block size m and IDT τ is
| (10) |
For h < hc, the mean block size is finite, while for h > hc, diverges. The probability distribution of IDTs on lineages [6] is given by
| (11) |
The expression indicates that the IDT distribution on lineages in the population is localized entirely at τ = τ0 above hc, despite the fact that the IDT distribution of isolated lineages remains k (τ).
The mean block size serves as an order parameter that diverges for h > hc, and the continuity of the phase transition can be characterized by its behavior near hc. If , the transition is a continuous phase transition, while if the transition is referred to as discontinuous, or ‘first-order’. Analogously, the lineage mean IDT, , which is given by ∂Λ/∂ ln z, exhibits the same type of phase transition. In terms of thermodynamics, the length of lineages t defines the system size and ln z can be seen as a generalized force (e.g. chemical potential). Then tΛ is the thermodynamic potential (e.g. grand potential) whose first-order derivative by ln z coincides with the expected number of divisions on lineages averaged across population, which is the extensive variable associated with ln z [6]. Examples of the h dependence of and are shown in Fig. 1B and C. Assuming the law of large numbers, and hold where D and S denote respectively the number of divisions and the number of switches to different values of τ on lineage [6].
Using the form of k (τ) given in Fig. 1, where k(τ) ~ (τ − τ0)γ for τ near τ0, and computing m(h) we find that the phase transition is continuous for 0 < γ ≤ 1, and becomes discontinuous for γ > 1 (see Fig. 1B). For a discontinuous transition, one expects to observe coexistence of two phases at the transition point, which is seen in numerical simulations of finite populations shown below (Fig. 2A). We can also compute the probability distribution of block sizes, which decays exponentially below the transition, and follows power law statistics, plin (m) ~ m−γ−1 for large m, in the limit as expected for correlation functions in the vicinity of a phase transition [6]. Such an appearance of long memory of IDT inheritance at the transition point is likewise observed in simulations (Fig. 2A).
FIG. 2.

Dynamics and aging of growth rates in a finite population. (A,B) Time courses of reciprocal empirical minimal IDT, (top panels, orange, solid), compared with the minimal IDT, (gray dashed line); and growth rates (bottom panels, blue line) compared with predicted stationary value (gray dashed line). (A) Coexistence of delocalized and localized states near the transition (h = 0.7); (B) aging dynamics above the transition (h = 0.8). Parameters are as in Fig. 1 using α = 4, which exhibits a first-order localization transition. Population size in simulations is maintained at N = 100, and the effective heritability threshold is hc,N = 0.73; see [6] for simulation details.
Dynamics and aging of population growth rate.
In finite sized populations, the stationary distribution (9) for h > hc is not achievable because any parent cell with IDT τ′ > τ0 has probability zero of generating offspring with IDT τ0. Additionally, the distribution (9) cannot be maintained as a steady-state because cells with IDT τ0 can be lost from the population within finite time due to coalescence. To observe dynamics in finite populations, we conducted exact stochastic simulations in which cells are randomly removed to maintain a fixed population size [6]. We sampled the initial population with size N = 100 independently from the stationary probability distribution with h = 0, which we refer to as the delocalized state, and simulated the population forward in time for a given value of h > 0.
In simulations with h < hc, growth rates Λt fluctuate around the expected steady-state growth rate (see [6]). Near the heritability threshold, however, Λt exhibits sudden transitions between two distinct, long-lived states (Fig. 2A), which is expected for systems near a first-order phase transition. One of these states represents localization at the empirical minimum IDT, denoted by , which is the minimum IDT among all the dividing cells within each time bin [6]. In this state, the growth rate fluctuates around over a sufficient period during which is constant. The other phase represents delocalization, where exhibits large fluctuations. For higher values of h, the observed growth rate increases stepwise and fluctuates around (Fig. 2B). In this case, a fraction of the population localized at is maintained over a significant time interval until a new value of replaces the current empirical minimum IDT. The time intervals between these replacement events become increasingly long as approaches τ0, because IDTs that are shorter than the current minimum become increasingly rare. Plotting values of Λt at different simulation times as function of h, the curve possesses an inflection point hc,N slightly greater than hc, which we refer to as the effective transition point at fixed population size N (see [6] for further details).
Collective divisions, cell cycle synchronization, and noisy inheritance.
In addition to aging dynamics, above the transition divisions occur collectively and periodically in finite populations. The number of divisions that occur, binned in short intervals, is plotted over time (Fig. 3, left panels), along with its auto-correlation function (Fig. 3, right panels). The auto-correlation function does not oscillate below the transition, but exhibits decaying oscillations near the transition point and sustained oscillations above the transition. The fact that the period of the autocorrelation function is approximately τ0 reflects localization at close to τ0. We note that division rate oscillations have also been predicted to arise in cell size control models, due to negative IDT correlations [14].
FIG. 3.

Cell cycle synchronization in finite populations. Parameters are as in Fig. 2. Number of divisions B(t, δt) binned over δt = 0.1 doubling time (A, C, E) and their auto-correlations (B, D, F) are shown. Simulations were run over tmax = 104 time units and the last 10 time units of the series are shown for A, C and E, where time point 0 indicates the simulation end. Auto-correlation function ACF(Δt) was computed over entire time series. The correlation coefficient was normalized to equal 1 at Δt = 0. Dotted line indicates zero correlation. (A,B) Below the transition (h = 0.6) timing of divisions is not synchronized. (C, D) Near the transition point (h = 0.7), timing of divisions is partially synchronized, observed as decaying oscillation of the auto-correlation function. (E, F) Above the transition (h > hc,N), collective divisions are observed indicating self-synchronization of cell cycles.
We tested the robustness of the transition properties to inaccuracy of the inheritance of IDTs by allowing small fluctuations of the offspring’s IDT when it inherits its parent’s IDT with probability h. To do so, we modified the transition kernel to be
| (12) |
where pnorm (x, σ) is a normal distribution density function with standard deviation σ, truncated at τ0. As a result, the population can reach equilibrium over a reasonable time scale, that is, the long-term behaviors coincide between the two distinct initial conditions. Collective divisions are weaker but still detectable through the decaying oscillations of the autocorrelation function of divisions in the population. Additionally, we analyzed the behavior of a general Gaussian kernel, truncated at τ0, with parameters specifying the mean, variance and parent-offspring correlation of IDTs (the latter can be negative or positive). In all cases, for sufficiently high heritability, we found that as the population size increases from N = 1 to N = 1000, Λ exhibits a pronounced increase (see Fig. 4A and additional results in [6]). In contrast, for low heritability or for negative IDT correlations, Λ varies little with N. Similar behavior is observed for , indicating that even without exact inheritance of parental IDTs, signatures of the phase transition are observed when varying the population size [6].
FIG. 4.

(A) Dependence of Λ on population size N. Simulations use Eq. 12 with σ = 0.1τ0 and parameters as in Fig. 2. (B) Fitting the Lebowitz-Rubinow model using single cell lineage data (F3 rpsL-gfp, glucose, 37°C from [3]). Empirical (shaded) and fitted (solid) cumulative distribution of plin (τ). Inset: empirical pdiv (τ). Fitting yields h = 0.23; for comparison, the Pearson correlation of mother-daughter IDTs is 0.26. See [6] for methods and additional data.
Discussion.
In this Letter, we analyzed how the strength of IDT heritability affects age-structured population dynamics. In a model with heritable cell cycle durations first introduced in [10], we demonstrate the existence of a localization phase transition with a heritability threshold above which the population’s distribution of IDTs localizes at the minimal IDT, corresponding to the fastest possible single-cell growth rate. We show that the ancestral mean IDT provides an order parameter of the transition, and above the heritability threshold predict the emergence of single cell lineages that maintain perfect inheritance of a minimal IDT.
A similar transition in exponentially growing populations is known in mutation-selection models of theoretical population genetics [9, 15, 16]. For example, Kingman’s house-of-cards model [16], which describes population dynamics on a specific type of fitness landscape, exhibits localization at maximal fitness below a critical mutation rate. The results can be generalized to cases in which fitness is correlated between parent and offspring using theory of positive linear operators [17].
The Lebowitz-Rubinow model is an idealization in which IDTs are either inherited precisely (with probability h) or not at all, and a key property of the model which underlies localization is the existence of a positive, minimal IDT, τ0. In [6], using experimental measurements in E. coli we show that by fitting empirical lineage IDT distributions (using Eqs. 4 and 11), the model correctly infers h (see Fig. 4B). Additionally, the data support the existence of τ0 > 0. Thus, the Lebowitz-Rubinow kernel captures key features of real biological data, and correctly predicts its underlying parameters.
We showed that the signature of the localization transition can be experimentally observed by measuring the population growth rate Λ at different population sizes in the range N = 1 − 1000, which is feasible in microfluidic experiments [3, 18, 19]. Varying the population size enables one to control the strength of selection in the experiments, as the size of heritable fitness differences that selection can act on efficiently scales with N−1. Negative IDT correlations, as often observed in bacteria (Table S1 in [6]), correspond to a lack of heritability and do not yield transition-like behavior for increasing N [6]. Above the IDT heritability threshold, population growth rates exhibit aging dynamics similar to evolutionary dynamics in a random, unbounded fitness landscape in the low mutation rate limit [20, 21]. While aging dynamics occur for precise IDT heritability, transition-like behavior is present in less accurate IDT inheritance systems typical of biological systems [6].
Remarkably, the Lebowitz-Rubinow model does not include cell-cell interactions, which are thought to be a major mechanism for cell-cycle synchronization, yet our analysis predicts that strong but imperfect IDT heritability may be sufficient to cause self-synchronization of cell cycles in finite populations. This finding provides a starting point for the design of new types of synthetic biological oscillators that leverage population-level selective forces to establish robust cell cycle synchronization with sustained oscillations.
Supplementary Material
Acknowledgments
We thank A. Skanata and Y. Wakamoto for discussions. This work was supported by NIH R01-120231.
References
- [1].Taheri-Araghi S, Bradde S, Sauls JT, Hill NS, Levin PA, Paulsson J, Vergassola M, and Jun S, Curr. Biol 25, 385 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [2].Susman L, Kohram M, Vashistha H, Nechleba JT, Salman H, and Brenner N, Proc. Natl. Acad. Sci. U. S. A 115, 201615526 (2018), arXiv:1805.05058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [3].Hashimoto M, Nozoe T, Nakaoka H, Okura R, Akiyoshi S, Kaneko K, Kussell E, and Wakamoto Y, Proc. Natl. Acad. Sci 113, 3251 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Amir A, Phys. Rev. Lett 112, 208102 (2014). [Google Scholar]
- [5].Grilli J, Cadart C, Micali G, Osella M, and Cosentino Lagomarsino M, Front. Microbiol 9, 1541 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [6].Supplemental Material at [URL will be inserted by publisher], which includes Refs. [3, 9, 17, 22–44].
- [7].Charlesworth B, Evolution in Age-Structured Populations, 2nd ed., Cambridge Studies in Mathematical Biology (Cambridge University Press, 1994). [Google Scholar]
- [8].Eigen M, McCaskill J, and Schuster P, Adv. Chem. Phys 75, 149 (1989). [Google Scholar]
- [9].Hermisson J, Redner O, Wagner H, and Baake E, Theor. Popul. Biol 62, 9 (2002). [DOI] [PubMed] [Google Scholar]
- [10].Lebowitz JL and Rubinow SI, J. Math. Biol 1, 17 (1974). [DOI] [PubMed] [Google Scholar]
- [11].IDTs of sibling cells may be correlated, which does not change Eq. (2), as long as the number of offspring is independent of the parent’s IDT.
- [12].Fisher R, The Genetical Theory of Natural Selection (Clarendon, 1930). [Google Scholar]
- [13].Wakamoto Y, Grosberg AY, and Kussell E, Evolution 66, 115 (2012). [DOI] [PubMed] [Google Scholar]
- [14].Jafarpour F, Phys. Rev. Lett 122, 118101 (2019), arXiv:1809.10217. [DOI] [PubMed] [Google Scholar]
- [15].Eigen M, Naturwissenschaften 58, 465 (1971). [DOI] [PubMed] [Google Scholar]
- [16].Kingman JFC, J. Appl. Probab 15, 1 (1978). [Google Scholar]
- [17].Bürger R and Bomze IM, Adv. Appl. Probab 28, 227 (1996). [Google Scholar]
- [18].Wallden M and Elf J, Curr. Opin. Biotechnol 22, 81 (2011). [DOI] [PubMed] [Google Scholar]
- [19].Lambert G and Kussell E, PLoS Genet. 10, e1004556 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Park SC and Krug J, J. Stat. Mech , P04014 (2008), arXiv:0711.1989. [Google Scholar]
- [21].Brotto T, Bunin G, and Kurchan J, J. Stat. Mech , 033302 (2016), arXiv:1407.4669. [Google Scholar]
- [22].Nozoe T, Kussell E, and Wakamoto Y, PLOS Genet. 13, e1006653 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [23].Mogensen PK and Riseth AN, Journal of Open Source Software 3, 615 (2018). [Google Scholar]
- [24].Seita A, Nakaoka H, Okura R, and Wakamoto Y, bioRxiv 10.1101/2020.07.10.196881 (2020). [DOI] [Google Scholar]
- [25].Mosheiff N, Martins BMC, Pearl-Mizrahi S, Grünberger A, Helfrich S, Mihalcescu I, Kohlheyer D, Locke JCW, Glass L, and Balaban NQ, Phys. Rev. X 8, 021035 (2018). [Google Scholar]
- [26].Kubitschek HE, Exp. Cell Res 26, 439 (1962). [DOI] [PubMed] [Google Scholar]
- [27].Yu FB, Willis L, Chau RMW, Zambon A, Horowitz M, Bhaya D, Huang KC, and Quake SR, BMC Biol. 15, 1 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [28].Hawkins ED, Markham JF, McGuinness LP, and Hodgkin PD, Proc. Natl. Acad. Sci 106, 13457 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [29].Sasaki H, Yoshinaga H, and Kawano K, Radiat. Res 72, 364 (1977). [PubMed] [Google Scholar]
- [30].Staudte RG, Guiguet M, and D’Hooghe MC, J. Theor. Biol 109, 127 (1984). [DOI] [PubMed] [Google Scholar]
- [31].Hola M and Riley PA, J. Cell Sci 88 (Pt 1), 73 (1987). [DOI] [PubMed] [Google Scholar]
- [32].Chakrabarti S, Paek AL, Reyes J, Lasick KA, Lahav G, and Michor F, Nat. Commun 9, 1 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [33].Froese G, Exp. Cell Res 35, 415 (1964). [DOI] [PubMed] [Google Scholar]
- [34].Van Wijk R and Van De Poll KW, Cell Prolif. 12, 659 (1979). [Google Scholar]
- [35].Absher M, Sylwester D, and Hart BA, Environ. Res 30, 34 (1983). [DOI] [PubMed] [Google Scholar]
- [36].Absher M and Cristofalo VJ, J. Cell. Physiol 119, 315 (1984). [DOI] [PubMed] [Google Scholar]
- [37].Sandler O, Mizrahi SP, Weiss N, Agam O, Simon I, and Balaban NQ, Nature 519, 468 (2015). [DOI] [PubMed] [Google Scholar]
- [38].Miyamoto H, Zeuthen E, and Rasmussen L, J. Cell Sci 13, 879 (1973). [DOI] [PubMed] [Google Scholar]
- [39].van Meeteren A, van Wijk R, Stap J, and Deys BF, Cell Prolif. 17, 105 (1984). [DOI] [PubMed] [Google Scholar]
- [40].Grasman J, Bull. Math. Biol 52, 535 (1990). [DOI] [PubMed] [Google Scholar]
- [41].Shields R and Smith JA, J. Cell. Physiol 91, 345 (1977). [DOI] [PubMed] [Google Scholar]
- [42].Staudte RG, Huggins RM, Zhang J, Axelrod DE, and Kimmel M, Math. Biosci 143, 103 (1997). [DOI] [PubMed] [Google Scholar]
- [43].Sennerstam R and Strömberg JO, J. Theor. Biol 131, 151 (1988). [DOI] [PubMed] [Google Scholar]
- [44].Dawson KB, Madoc-Jones H, and Field EO, Exp. Cell Res 38, 75 (1965). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
