Paper The following article is Free article

A transport approach to relate asymmetric protein segregation and population growth

and

Published 28 July 2021 © 2021 IOP Publishing Ltd and SISSA Medialab srl
, , Citation Jiseon Min and Ariel Amir J. Stat. Mech. (2021) 073503 DOI 10.1088/1742-5468/ac1262

1742-5468/2021/7/073503

Abstract

Many unicellular organisms allocate their key proteins asymmetrically between the mother and daughter cells, especially in a stressed environment. A recent theoretical model is able to predict when the asymmetry in segregation of key proteins enhances the population fitness, extrapolating the solution at two limits where the segregation is perfectly asymmetric (asymmetry a = 1) and when the asymmetry is small (0 ⩽ a ≪ 1) (Lin et al 2019 Phys. Rev. Lett. 122 068101). We generalize the model by introducing stochasticity and use a transport equation to obtain a self-consistent equation for the population growth rate and the distribution of the amount of key proteins. We provide two ways of solving the self-consistent equation: numerically by updating the solution for the self-consistent equation iteratively and analytically by expanding moments of the distribution. With these more powerful tools, we can extend the previous model by Lin et al (2019 Phys. Rev. Lett. 122 068101) to include stochasticity to the segregation asymmetry. We show the stochastic model is equivalent to the deterministic one with a modified effective asymmetry parameter (aeff). We discuss the biological implication of our models and compare with other theoretical models.

Export citation and abstract BibTeX RIS

1. Introduction

Genetically identical unicellular organisms may nevertheless exhibit different traits such as growth rate or generation time, a phenomenon known as phenotypic variability [1]. One cause of such diversity is the asymmetric allocation of key proteins at division. For instance, some bacteria and eukaryotic cells accumulate protein aggregates when exposed to stress and tend to segregate more of them to one of the two newborn cells at a division event. As a result, the 'rejuvenated' cells (i.e. those with less damage) grow faster than the 'senescent' cells [27]. There have been a number of theoretical works [811] trying to quantify the effect of asymmetric segregation of proteins to the overall growth of the population. In this paper, we strengthen the theory developed by Lin et al [8] to relate the asymmetric protein segregation and the population growth rate by implementing a more general method involving transport equations and moment expansions. This allows us to recover previous results rigorously, and go beyond them. We extend the model to consider stochastic asymmetric segregation. Lastly, we compare alternative models of asymmetric segregation by applying the self-consistent equations.

2. Model

We define a generalization of the model proposed by Lin et al [8]. Consider a clonal microbial population that grows exponentially with rate Λp . We assume that as a cell grows by volume ΔV, the amount of key proteins D increases by ΔD = SΔV. Each cell's inter-division time (τ) is a function of the initial concentration of the key protein (σb). For simplicity, suppose every cell is born with volume Vb = 1 and divides when the volume reaches Vd = 2. As it divides, the key proteins segregate asymmetrically according to the asymmetry parameter a: one daughter cell inherits $\frac{1+a}{2}$ of the key proteins from the mother and the other gets the rest. We call the key protein damaging if the generation time τ[σb] is an increasing function of the initial protein concentration and beneficial if it is a decreasing function (figure 1).

Figure 1.

Figure 1. Single cell level model. For simplicity, we assume that every cell has a volume Vb = 1 at birth and divides when it reaches Vd = 2. The amount of key proteins in a cell increases in proportion to the added volume with a prefactor S. The generation time τ[σb] is a function of the key protein concentration at birth (σb = Db/Vb). Asymmetry a ∈ [0, 1] tunes the ratio of key proteins inherited by two offspring cells (a = 0 being perfectly asymmetric, a = 1 being perfectly asymmetric). The plots on the right show an example of generation time as a function of initial concentration of key protein which is deleterious (top) or beneficial (bottom).

Standard image High-resolution image

While our analysis can be applied to any analytical function τ[σb], we assume the instantaneous growth rate λ[σ] is a Hill-type function of the key protein concentration of the cell, as proposed by Lin et al [8]. These assumptions by Lin et al are based on observations made in experimental studies, where the growth rate function of microbes often has an inflection point [5, 6].

Equation (1)

Equation (2)

Every time a cell divides, the asymmetry parameter a is randomly and independently drawn from a probability distribution g(a). The deterministic model by Lin et al [8] assumes a delta function for g(a), but here we allow a to fluctuate.

3. Results

3.1. Transport equation

Recently, Levien et al [12] showed that a transport equation can be used to get a self-consistent equation for the population distribution of phenotype and population growth rate even when the properties of the cells are correlated across the generations. Here we follow their derivation by setting the phenotype of interest as the key protein concentration at birth σb.

Assume that the population size is N at time t and it grows exponentially with the rate Λp . Let us also assume that the population has reached a stable probability distribution ψ(u, σb, t), where u is time since birth, σb is the key protein concentration at birth, and t is the global time. The probability distribution then evolves during an infinitesimal time step dt as:

Equation (3)

Rearranging the terms and dividing both sides by dt, we get

Equation (4)

In a steady state population, there is no dependence on t, so ∂t ψ = 0. Therefore,

Equation (5)

Equation (6)

The boundary conditions are set by the division. The probability distribution of having a newborn cell with a protein concentration σb is

Equation (7)

The probability distribution f(σb|σb', a') and g(a') depend on the physiological model we choose. For instance, in a deterministic model, we have

Equation (8)

Equation (9)

Applying the solution given by equation (6) (in the general case), we can eliminate the dependence on u. Thus, we can rewrite to obtain a self-consistent equation:

Equation (10)

In the next section, we apply the self-consistent equation (10) to a toy model of E. coli recently published by Blitvic and Fernandez [13] as a simple example.

3.2. Toy model of the aging of E. coli [13]

Let us introduce a simple model published by Blitvic et al to demonstrate how to use our approach to connect the asymmetry between two newborn cells and population growth rate [13]. Blitvic et al assume that as an E. coli cell divides, one daughter cell is rejuvenated and the other senescent where the generation time of a rejuvenated cell (τ[R]) is shorter than a senescent cell (τ[S]). In this model, there are two phenotype x = S for a senescent cell and x = R for a rejuvenated one.

Then the transition density f(x|x') from a phenotype x' to x is

Equation (11)

The self-consistent equation is

Equation (12)

Integrating both sides with respect to x, we have

Equation (13)

Because the exponential function is convex, by Jensen's inequality,

Equation (14)

where $\left\langle \tau [x]\right\rangle =\frac{1}{2}(\tau [S]+\tau [R])$. Thus, the asymmetry of generation time of two distinct phenotypes enhances the population growth rate if the mean of the generation time, $\left\langle \tau [x]\right\rangle $ is kept constant. Note that equation (13) was also derived in [14] in the context of the asymmetric division of the budding yeast Saccharomyces cerevisiae.

In the next section, we solve equation (10) numerically and analytically more generally.

3.3. Numerically solving the self-consistent equation

The self-consistent equation (10) can be numerically solved by iterating the following procedure. First, we start with initial guesses for the solutions ${\psi }_{\text{birth}}^{(0)}[{\sigma }_{\mathrm{b}}]$ and ${{\Lambda}}_{p}^{(0)}$. We then update our guess for the distribution ψbirth[σb] using equation (10). Next, we update Λp by using the Euler–Lotka equation [12], obtained by integrating equation (10) with respect to σb. This procedure is summarized in this set of equations:

Equation (15)

Equation (16)

Equation (17)

${\psi }_{\text{birth}}^{(n)}[{\sigma }_{\mathrm{b}}]$ and ${{\Lambda}}_{p}^{(n)}$ converge to constant values as n. Note that the solution is not sensitive to the choice of ${\psi }_{\text{birth}}^{(0)}$ and ${{\Lambda}}_{p}^{(0)}$. In appendix D, we discuss the shape of the distribution ψbirth[σb].

In figure 2, we compare the numerical solution for equation (10) in the case of deterministic a (i.e. g(a') is a delta function) and the fitted exponential growth rate obtained by simulating the population directly (see appendix A for details of the simulation).

Figure 2.

Figure 2. The relative population growth rate, compared to a = 0, is plotted against the asymmetry a. The solid lines show the numerical solutions of equation (10) and the circles mark simulated population growth rates. (a) Beneficial protein (λ0 = 0.2, λ1 = 1.2, n = 2), (b) deleterious protein (λ0 = 1, λ1 = 0, n = 2). For numerical solutions, we use ${\psi }_{\text{birth}}^{(0)}[\sigma ]=\delta ({\sigma }_{\mathrm{b}}-S)$ and ${{\Lambda}}_{p}^{(0)}=\mathrm{ln}\enspace 2/\lambda [S]$. Error bars are smaller than the marker size.

Standard image High-resolution image

Figure 2 confirms that equation (10) has sufficient information for calculating the population growth rate. The iterative procedure is more accurate, faster, and less memory-intensive than a brute force simulation which involves exponentially growing population size. However, the main shortcoming of the numerical solution is that it does not give any insight about how Λp depends on S and a. It is possible infer the relation by finding Λp for various values of S and a, but for every parameter value, we have to go through equations (15)–(17), which is highly inefficient. For those reasons, the moment expansion method we discuss in the following section is a good alternative way of calculating Λp for small a.

3.4. Moment expansion

Equation (10) can be approximately solved by expanding the moments of ψbirth[σb]. For simplicity, we start with the deterministic case. As we will explain in the following sections, it is straightforward to generalize this moment expansion method to the stochastic case.

When the key protein segregation is perfectly symmetric (i.e. a = 0), every cell has a concentration of key protein of σb = S at cell birth. Therefore, for a small a, the distribution ψbirth[σb] should be sharply peaked around S. Based on this, we are motivated to expand the moment of the distribution around σb = S by defining Δσb = σbS. In order to calculate the nth moment, we multiply both sides of equation (10) by ${\Delta}{\sigma }_{\mathrm{b}}^{n}$ and integrate them with respect to σb. We use equations (8) and (9) to specify f(σb|σb', a') and g(a'). Invoking Newton's binomial theorem, we obtain:

Equation (18)

Note that by convention, $\left(\genfrac{}{}{0pt}{}{n}{0}\right)=1$ and ⌊n/2⌋ is the greatest integer that is less than or equal to n/2. From the right-hand side of equation (18), we can infer that $\langle {\Delta}{\sigma }_{\mathrm{b}}^{n}\rangle =O({a}^{2\lfloor n/2\rfloor })$. This means that by solving a system of 2k + 1 equations (n = 0, ..., 2k), we will find Λp to O(a2k ) accuracy. For each equation, we need to expand the exponential function ${\mathrm{e}}^{-{{\Lambda}}_{p}\tau [{\Delta}{\sigma }_{\mathrm{b}}+S]}$ around Δσb = 0 to order a2⌊n/2⌋. To illustrate the process more clearly, let us calculate Λp to O(a2) accuracy.

If a = 0, every cell has σb = S when the population is stabilized, so ${{\Lambda}}_{p}=\frac{\mathrm{ln}\enspace 2}{\tau [S]}$. Due to symmetry, Λp [a] = Λp [−a]. Thus, to solve Λp to O(a2) is equivalent to finding a coefficient C2 defined as below,

Equation (19)

Taylor expanding τσb + S] around Δσb = 0 to O(a2) we find

Equation (20)

Expanding ${\mathrm{e}}^{-{{\Lambda}}_{p}\tau [{\Delta}{\sigma }_{\mathrm{b}}+S]}$ to O(a2) leads to:

Equation (21)

Applying these to equation (18) while setting n = 0, 1, or 2, we can set up a system of three linear equations with the unknowns being ${C}_{2},\left\langle {\Delta}{\sigma }_{\mathrm{b}}\right\rangle $ and $\left\langle {\Delta}{\sigma }_{\mathrm{b}}^{2}\right\rangle $. (Note that here τ[S] is a known function specified in the model.)

Equation (22)

Equation (23)

Equation (24)

Solving these, we get

Equation (25)

Equation (26)

Equation (27)

Therefore, the population growth rate is

Equation (28)

This is general to any choice of τ[σb], but in particular, if the generation time is described by equation (2), we can rewrite the population growth rate in terms of λ[S].

Equation (29)

This second order solution for Λp is consistent with the result by Lin et al [8] but is derived here more rigorously and more generally since it applies to any functional form of τ[σb]. Furthermore, we can calculate Λp to an arbitrarily high order as desired by solving for ${{\Lambda}}_{p}=\frac{\mathrm{ln}\enspace 2}{\tau [S]}\left(1+{C}_{2}{a}^{2}+{C}_{4}{a}^{4}+\cdots \enspace \right)$, as we elaborate in appendix C.

4. Phase transition of ac

In the previous sections, we explained how to solve equation (10) and predict the population growth rate Λp as a function of growth rate of a single cell λ[σ], protein accumulation rate S, and asymmetry a. In the context of evolution, we can imagine that a single-celled organism may adapt to environmental stress (characterized by S and λ[σ] in our model) by tuning the asymmetry a. In population genetics, fitness of a population is often defined as an exponential growth rate of the population size, and traits that increase the fitness tend to survive longer due to natural selection [15]. In fact, there are several experimental studies reporting that segregation of protein aggregates tend to become more asymmetric as temperature increases above the optimal temperature [25]. In this section, we present how to find the optimal asymmetry ac, the value of a that maximizes the population growth rate Λp , as a function of population accumulation rate S. As Lin et al [8] did, we find a sharp phase transition of ac if the key protein is deleterious and a smooth transition in the case of beneficial key protein.

Previously, Lin et al [8] inferred ac[S] by interpolating the solutions near a = 0 and at a = 1 with a minimal number of extrema. Lin et al used a Landau approach to find the population growth rate Λp to O(a2) accuracy for a small a. In summary, the Landau approach investigates the shape of the fitness curve f = Λp [a] − Λp [a = 0] as a function of a at a given key protein accumulation rate S (i.e. the fitness is analogous to a free energy with an order parameter a, and S to temperature). They derived the relationship between the fitness f and asymmetry a from an equation for the total cell volume, V = ∑i vi (t).

Equation (30)

Then they Taylor expand λ[σ] around σ = S.

Equation (31)

Using the fact that odd power terms of a should vanish due to symmetric cell division in terms of volume and an ansatz that the variance of the distribution of σb scales as a2, Lin et al derived equation (29) without any knowledge about the distribution ψ[σb].

Equation (29) breaks down as a approaches 1, but when a is exactly 1, there exists a closed formula for Λp due to the self-similarity in a population tree as illustrated in figure 3. Lin et al showed that

Equation (32)

Figure 3.

Figure 3. The lineage tree for a population with a = 1 is self-similar. Specifically, a population starting with a cell with no key protein at birth (σb = 0, marked by a green circle) should all have the same population size as a function of time since birth of the ancestor. Using this property Lin et al derived equation (32) [8].

Standard image High-resolution image

Interpolating the solutions at a ≈ 0 and a = 1 with the minimal number of extrema, Lin et al infers the value of ac (i.e. what value of a maximizes Λp for a given S) as described in figure 4. They found that the location of ac depends on the relative magnitude of two critical values of S: Sc and S*. Each of these two critical values arises from the solutions at a ≈ 0 and a = 1. Sc is where the second derivative of λ[S] is zero, which means that the fitness near a = 0 changes its sign when S = Sc. S* is where Λp [a = 1] = Λp [a = 0]. See figure 4 for the inferred shape of the fitness curve at various values of S relative to Sc and S*. Note that ac is between 0 and 1 (marked by star in figure 4) in the case of beneficial key protein when S is between S* and Sc.

Figure 4.

Figure 4. Finding the optimal asymmetry ac by graphical interpolation [8]. Lin et al find the second derivative of the fitness (Λp [a] − Λp [0]) near a = 0 (solid line) and numerically solve the self-consistent equation for a = 1 (equation (32), filled circles). Interpolating between the two limits with a minimal number of extrema, the authors find what value of a maximizes the fitness (empty stars). Unlike in the damage case, in the benefit case, there is a large range of S where the fitness curve is convex near a = 0 and Λp [a = 1] < Λp [a = 0] (i.e. S* < S < Sc). This means that the population growth rate is maximized when a is between 0 and 1.

Standard image High-resolution image

Similar to the previous paper by Lin et al, we use an interpolation to infer where the phase transition of ac is. In addition, we are equipped with a more rigorous way of finding the population growth rate near a = 0 as described in section 3.4. We also have a transcendental equation for the slope of Λp [a] at a = 1, which we discuss in depth in appendix B. We denote the solution to equation (32) as Λp1 and expand Λp [a] = Λp1(1 + C1(1 − a) + O((1 − a)2)). Finding the sign of C1 makes it possible to predict whether Λp [a = 1] is a local minimum (C1 > 0) or a local maximum (C1 < 0). Using a similar moment expansion approach as we do for a ≈ 0, we find the first coefficient C1 in appendix B as

Equation (33)

Here we can define another critical value ${S}_{\mathrm{c}}^{a=1}$ which is the value of S that makes C1 = 0. Because τ'[kS] > 0  (τ'[kS] < 0) for the damage (benefit) case, C1 starts out positive (negative) for a small S and becomes negative (positive) as S exceeds ${S}_{\mathrm{c}}^{a=1}$. In other words, when $S{< }{S}_{\mathrm{c}}^{a=1}$, a = 1 is a local minimum (maximum) of Λp [a], whereas it becomes a local maximum (minimum) when $S{ >}{S}_{\mathrm{c}}^{a=1}$ in the damage (benefit) case. This provides us an additional piece of information that we can use to infer whether ac = 1 for a given S. For instance, in the case of damage (λ0 = 1, λ1 = 0), ${S}_{\mathrm{c}}^{a=1}\approx 0.592$ is greater than S* ≈ 0.558. This means that if S is between S* and ${S}_{\mathrm{c}}^{a=1}$, Λp [a] should have a local maximum below a = 1. When S is above both S* and ${S}_{\mathrm{c}}^{a=1}$, the population growth rate is maximized at a = 1 (i.e. ac = 1). This means that the transition of ac is not a single discrete jump from ac = 0 to ac = 1, but rather a jump to an intermediate ac ∈ (0, 1) and then a smooth increase to ac = 1. This process is illustrated in figure 5.

Figure 5.

Figure 5. Increase of Λp relative to a = 0 against ac calculated by numerically solving equation (10). S increases as the line gets lighter (S = 0.554 to 0.592). The stars mark ac for each S. If S < S*, then automatically, S < Sc and $S{< }{S}_{\mathrm{c}}^{a=1}$. This means that a = 0 is a local maximum and there is also another peak between a = 0 and a = 1. At S = S*, $S{< }{S}_{\mathrm{c}}^{a=1}$, which means that there is a < 1 that has a higher Λp than a = 0 and a = 1. Therefore, there should be a first-order phase transition at S < S*.

Standard image High-resolution image

In the case of beneficial key proteins, Lin et al found that ac = 1 for small S and decreases smoothly to ac = 0 as S increases, but the start of that transition was not specified. Based on appendix B, we find that ac starts deviating from one at ${S}_{\mathrm{c}}^{a=1}=0.214$ (blue dotted line in figure 6(a)), consistent with the numerical solution of equation (10). Comparing the solid blue line (ac predicted by the moment expansion to the sixth order) and the orange dots (numerical simulation), note that the moment expansion method does not successfully predict ac when the true ac value is close to 1. However, as S increases, and ac becomes closer to 0, the answers agree better, since a2 ≈ 0 our approximation holds.

Figure 6.

Figure 6.  ac, the value of a that maximizes the population growth rate, plotted against S. The solid blue lines are the prediction made from the sixth order solution of Λp . The orange points, each of which has an error bar, show ac found by solving equation (10) numerically as described in section 3.3. (a) Beneficial protein (λ0 = 0.2, λ1 = 1.2, n = 2), (b) deleterious protein (λ0 = 1, λ1 = 0, n = 2). The inset zooms in on the region around S = 0.6. The dotted blue line shows where the slope of Λp at a = 1 changes its sign. The red dashed line is at S = S* where Λp [a = 0] = Λp [a = 1] and the black dashed line is at S = Sc where λ''[S] = 0 (i.e. the second derivative of Λp [a] changes its sign at a = 0). As anticipated, for (a) the exact numerics show a second-order phase transition coinciding with the blue dotted line, and a constant ac = 0 for S larger than the dashed black line. For (b), the exact numerics show a first-order phase transition coinciding with the red dashed line, and constant ac = 1 for S above the black dashed line.

Standard image High-resolution image

5. Introducing noise

The moment expansion method can be easily generalized to the stochastic model where a is drawn from a distribution g(a) every time a cell divides. Equation (18) becomes

Equation (34)

This implies that in the small a limit, we can define an effective a, ${a}_{\text{eff}}=\sqrt{\langle {a}^{2}\rangle }$, for which the population growth rate Λp has the same formula independent of g(a).

Equation (35)

In figure 7, we numerically solve equation (10) where g(a) is a uniform distribution ($g(a)=U(\overline{a}-{\sigma }_{a},\overline{a}+{\sigma }_{\mathrm{b}})$ with $\overline{a}=0$ and σa = 0, 0.1, 0.2, ..., 1) and plot the population growth rate against ${a}_{\text{eff}}=\sqrt{\left\langle {a}^{2}\right\rangle }$. As predicted, aeff behaves exactly as a does in the deterministic model for a broad range of σa . For a relatively large S and large σa , the population growth rate obtained by simulating the exponentially growing population starts to deviate from the second order prediction as shown in figure 7. This conclusion that stochastic segregation of key proteins increases effective asymmetry is analogous to the finding made by Chao et al [11] about damage partitioning in E. coli that larger fluctuations result in similar increase of the population growth as a fixed asymmetric ratio.

Figure 7.

Figure 7. The change of population growth rate (Λp [aeff] − Λp [a = 0]) plotted against the effective value of a where $g(a)=U(\overline{a}-{\sigma }_{a},\overline{a}+{\sigma }_{a})$, a uniform distribution. The solid lines show the numerical solutions of equation (10) for σa = 0 and the circles correspond to simulated values where $\overline{a}=0$ and σa = 0, 0.1, 0.2, ..., 1. (a) Beneficial protein (λ0 = 0.2, λ1 = 1.2, n = 2), (b) deleterious protein (λ0 = 1, λ1 = 0, n = 2).

Standard image High-resolution image

6. Results when asymmetry a depends on the concentration of key proteins

So far, we have considered a model where the segregation ratio of key proteins is independent from the amount of key proteins each cell has. However, there is experimental evidence that the degree of asymmetry may depend on how much key proteins a cell has [2, 7]. Vedel et al [5] studied how protein aggregates in E. coli created in responds to heatshock are segregated differently between two daughters. One of their results is that the asymmetric segregation parameter a is a function of the amount of protein aggregates D, and if a is assumed to be constant, the fitness is underestimated. We can test the effect of dependence of a to D by replacing a with a[D] in equation (10).

Equation (36)

Here we replaced D with σb' + S since we assume the cell volume always is 1 at birth and 2 at division. In figure 8, we show the numerical value of population growth rate increases due to asymmetry (Λp [a] − Λp [a = 0]), while a is a function of the amount of protein at division. Specifically, we set a[D] = tanh(kD) and tune k. In other words, we assume a to increase from zero initially with the slope k and asymptotically approach a = 1. Qualitatively, we expect the effective a to be small when S is small and to become closer to 1 as S gets larger. We can check this intuition by writing down the equation for the moments of ψbirth analogous to equation (18).

Equation (37)

Figure 8.

Figure 8. The change of population growth rate (Λp [a = tanh(kD)] − Λp [a = 0]) plotted against accumulation rate S. Here the key proteins are deleterious (λ0 = 1, λ1 = 0, n = 2). The blue cross show the values when a is fixed at 1 (Λp [a = 1] − Λp [a = 0]) and the red circles show the change of the population growth rate when a is equal to the optimal value ac at a given S.

Standard image High-resolution image

If the slope of a[D] is small compared to a2, we can approximate $a{[{\Delta}{\sigma }_{\mathrm{b}}+2S]}^{2m}\left.\right]\approx a[2S]$. Note that the population growth rate can be higher with an adaptive a (e.g. with k = 0.9, 0.55 < S < 0.65) than when a is fixed at ac[S] (orange dots in figure 8). Qualitatively, this suggests that there is an advantage if each cell tunes the asymmetry of segregation according to its level of damage.

7. Discussion and outlook

In this paper, we investigate how asymmetric segregation of key proteins in microbial organisms impact the population growth rate with a self-consistent equation (equation (10)) solved analytically and numerically. We adapt a physiological model proposed by Lin et al [8] and apply the transport equation to set up an equation for the distribution of protein concentration at birth ψ[σb] and population growth rate Λp (equation (10)). The equation can be solved by updating the solutions iteratively as elaborated in section 3.3. When the asymmetry is small (a ≪ 1), Λp can be approximated to a2n precision by solving a set of linear equations with respect to the moments of ψ[σb] and Λp . Similar approximations can be made from the opposite limit, a ≈ 1 (see appendix B). Changing the accumulation rate of key proteins, S, we confirm the sharp phase transition of ac for a deleterious protein and a smooth transition for a beneficial protein as Lin et al reported [8]. In addition, since numerically solving equation (10) provides more precise prediction of Λp than a direct simulation of an exponentially growing population, we are able to find that the phase transition in the case of damage is composed of a discrete jump to ac between 0 and 1 and a smooth increase to ac = 1. We also explore different physiological models motivated by experimental studies. We discuss the case where the cell tunes the asymmetry with response to the amount of key proteins it has at division and how the adaptive strategy may minimize the disadvantage of asymmetry when the accumulation rate of damage is low. Finally, we show by generalizing our model that even if the segregation ratio is on average symmetric, the noise can have a similar effect as an asymmetric segregation.

There are many other aspects of asymmetric division in biology where our theoretical framework might be useful for building a theoretical model [1, 16]. First, there are many other ways that microbes develop phenotypic diversity rather than due to asymmetric key protein segregation. For instance, budding yeast (Saccharomyces cerevisiae) has cell size difference between a daughter (smaller cell) and mother cell (bigger cell), whereas when a Caulobacter crescentus cell divides, one is non-motile and reproductive and the other swarms and does not divide until maturation [17, 18]. In the former case, introduction of volume asymmetry turns out to create interesting correlations between the noise in growth and the population growth rate [19]. Second, mortality rate rather than volumetric growth rate may depend on the key protein concentration [20]. Lastly, connecting back to multicellular organisms, where the theory of aging (soma theory) was first proposed [21], can be interesting both mathematically and biologically. Major difference from microbial system is that the cells interact with the neighbors and use the cues to make decisions [22]. Considering the geometry of the tissue might complicate the model beyond the level of a self-consistent equation can capture, but plants which have rigid tissue and simpler geometry may be a good starting point for a theoretical investigation [23].

Acknowledgments

The authors thank the Amir group, the Desai lab, Andrew Murray, Jane Kondev and Doeke Hekstra for helpful discussions. J M acknowledges Yubo Su for helping improve numerical simulations. A A was supported by the NSF CAREER Award Number 1752024.

Appendix A.: Simulation details

When we explicitly simulate the exponentially growing population, we start with 100 cells with uniform distribution of age. We run until the number of cells reaches 5 × 106 and find the slope of ln(∑V(t)) after t = 5. This use of sum of volume rather than the number of cells for finding the population growth rate is because the volume increases continuously and the growth rate should be equivalent if the cell size is controlled. To solve equation (10), we discretize the range of σb and repeat the process described in section 3.3 until the difference between 1 and $2\int \int {\psi }_{\text{birth}}^{(n)}[{\sigma }_{\mathrm{b}}^{\prime }]g({a}^{\prime }){\mathrm{e}}^{-{{\Lambda}}_{p}^{(n-1)}\tau [{\sigma }_{\mathrm{b}}^{\prime }]}\mathrm{d}{a}^{\prime }\enspace \mathrm{d}{\sigma }_{\mathrm{b}}^{\prime }$ is smaller than some tolerance value. This implies that ${{\Lambda}}_{p}^{(n-1)}\approx {{\Lambda}}_{p}^{(n)}$ and our solution for Λp has converged. Python scripts for the simulations are on our Github (https://github.com/jiseonmin/asymmetric_segregation).

Appendix B.: Λp when a  1

The challenge for a highly asymmetric case is that the distribution ψbirth[σb] is not symmetric around S. Because one of two newborn cells is rejuvenated at each division, we would expect the distribution to be skewed to the right and maximized near σb = 0. This means that we should expand ψbirth[σb] around different values depending on when the cell has been rejuvenated for the last time (i.e. generational age), in order to solve the self-consistent equation. Categorizing cells based on their generational age, we can rewrite the self-consistent equation (10) as,

Equation (B.1)

Here, ψi [σb] is the probability that a cell has protein concentration σb at birth and the generational age i. The transition function for the rejuvenated cell is $f(\sigma ,\mathrm{R}\vert {\sigma }_{\mathrm{b}}^{\prime })=\delta \left(\sigma -\frac{1-a}{2}({\sigma }_{\mathrm{b}}^{\prime }+S)\right)$, and the senescent is $f(\sigma ,\mathrm{S}\vert {\sigma }_{\mathrm{b}}^{\prime })=\delta \left(\sigma -\frac{1+a}{2}({\sigma }_{\mathrm{b}}^{\prime }+S)\right)$. Note that the rejuvenating transition function does not depend on σb' (parent cell) if a = 1. This is equivalent to the property used by Lin et al [8] to find a closed formula for Λp [a = 1]: there is a self-symmetry in a population tree in a stable exponential growth.

Equation (B.2)

Here, we can use the fact that every ψi [σb] is a delta function if a = 1.

Equation (B.3)

Equation (B.4)

By integrating ψ0[σb] with respect to σb, we get

Equation (B.5)

which is consistent with the previous paper [8]. Let us call the unique solution for equation (B.5), Λp,1. Near a = 1, we can find the approximate population growth rate by expanding Λp = Λp1(1 + C1(1 − a) + C2(1 − a)2 + ⋯) where a < 1. As a deviates from 1, we expect each delta function to broaden near σb = iS. Similar to the case of a small a, we will find the moments of ψi [σb] starting from the lowest order. Here, we will investigate the first order coefficient C1. Importantly, finding the sign of C1 will allow us to predict whether Λp [a = 1] is a local minimum (C1 > 0) or a local maximum (C1 < 0).

Let us define δa = 1 − a. To find C1 we need to solve equation (B.1) to O(δa). Thus, we can rewrite the equations as

Equation (B.6)

For brevity, from now on, we shall write Ψi = ∫ψi [σb]dσb and ${\left\langle F[\delta {\sigma }_{\mathrm{b}}]\right\rangle }_{i}=\int {\psi }_{i}[{\sigma }_{\mathrm{b}}]F[{\sigma }_{\mathrm{b}}-iS]\mathrm{d}{\sigma }_{\mathrm{b}}/{{\Psi}}_{i}$. First, let us integrate the ith equation (i > 0),

Equation (B.7)

For i = 0 we have

Equation (B.8)

to O(δa) accuracy. Using the definition of Λp1$1={\sum }_{i=0}^{\infty }{\mathrm{e}}^{-{{\Lambda}}_{p1}{\sum }_{j=0}^{i}\tau [jS]}$, we can extract an expression for the coefficient C1.

Equation (B.9)

Let us find ${\left\langle \delta {\sigma }_{\mathrm{b}}\right\rangle }_{i}$ for i > 0.

Equation (B.10)

If i = 0,

Equation (B.11)

Equation (B.12)

This shows that ${\left\langle \delta {\sigma }_{\mathrm{b}}\right\rangle }_{i}$ is positive for small i and becomes more and more negative as i increases. Applying this to equation (B.9), we can find the slope near a = 1.

Equation (B.13)

Appendix C.: Finding Λp [a] to O(a2n ) (n>  1)

In order to calculate Λp [a] to O(a2n ) accuracy, we write Λp and τ[σb] as

Equation (C.1)

Equation (C.2)

Using these, we write

Equation (C.3)

Next, we should expand equation (18) and truncate any terms with higher order than a2n . We can also rewrite $\langle {\sigma }_{\mathrm{b}}^{m}\rangle \enspace (1{\leqslant}m{\leqslant}2n)$ as

Equation (C.4)

Equation (C.5)

Maintaining only terms to O(a2n ) and matching the coefficients of a2j (1 ⩽ jn), we can solve the system of equations for D and C's. We provide the SymPy code to find the coefficients on our Github repository (https://github.com/jiseonmin/asymmetric_segregation).

Appendix D.: The shape of ψbirth[σb] changes depending on asymmetry a

Here we present some examples of the distribution of protein concentration at birth ψbirth[σb], which we find by numerically solving equation (10) (red circles). In each simulation, we compare the numerically found distribution to a log-normal distribution and a normal distribution that have the same mean and variance as the numerical solution. In other words, we test whether it is possible to approximate the distribution of key protein concentration to a log-normal or a normal distribution and simplify equation (10) (figures D1D3).

Figure D1.

Figure D1. Distribution ψbirth[σb] can be approximated as a Gaussian or a log-normal distribution if $\vert \left\langle a\right\rangle \vert \ll {\sigma }_{a}$). (a) Beneficial protein (λ0 = 0.2, λ1 = 1.2, n = 2), (b) deleterious protein (λ0 = 1, λ1 = 0, n = 2). In both simulations, σa = 0.2 and $\left\langle a\right\rangle =0.1$.

Standard image High-resolution image
Figure D2.

Figure D2. Distribution ψbirth[σb] is in general skewed and multi-modal. (a) Beneficial protein (λ0 = 0.2, λ1 = 1.2, n = 2), (b) deleterious protein (λ0 = 1, λ1 = 0, n = 2). In both simulations, σa = 0.2 and $\left\langle a\right\rangle =0.4$.

Standard image High-resolution image
Figure D3.

Figure D3. The Gaussian approximation only holds for the small asymmetry case because the skewness of distribution grows as a function of a. To illustrate this effect, we plot the skewness of ψbirth σb against the average of a for the damage and benefit case. To visualize the change of distribution, we show distribution with different values of a. The distributions are found numerically from the iterative solution of the self-consistent equations. (a) Beneficial protein (λ0 = 0.2, λ1 = 1.2, n = 2), (b) deleterious protein (λ0 = 1, λ1 = 0, n = 2). In both simulations, there is very small noise in a (σa = 0.02). In both simulations, we use S = 0.5 and assume deterministic segregation (σa = 0).

Standard image High-resolution image

When a is small and the segregation is noisy (σa > 0), the distribution can be approximated as a Gaussian distribution or a log-normal distribution. However, as a increases, the skewness of the distribution grows and the distribution becomes multi-modal. This implies that with small average asymmetry, the first two moments of the distribution have sufficient information for approximately solving equation (10) for Λp , but as the segregation becomes more and more asymmetric, higher order terms of a are necessary to accurately calculate the population growth rate. The exact value of a where the Gaussian approximation breaks down depends on the shape of λ[σ] and S.

Please wait… references are loading.
10.1088/1742-5468/ac1262