Skip to main content
Log in

A Diffusion Approximation Based on Renewal Processes with Applications to Strongly Biased Run–Tumble Motion

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

Abstract

We consider organisms which use a renewal strategy such as run–tumble when moving in space, for example to perform chemotaxis in chemical gradients. We derive a diffusion approximation for the motion, applying a central limit theorem due to Anscombe for renewal-reward processes; this theorem has not previously been applied in this context. Our results extend previous work, which has established the mean drift but not the diffusivity. For a classical model of tumble rates applied to chemotaxis, we find that the resulting chemotactic drift saturates to the swimming velocity of the organism when the chemical gradients grow increasingly steep. The dispersal becomes anisotropic in steep gradients, with larger dispersal across the gradient than along the gradient. In contrast to one-dimensional settings, strong bias increases dispersal. We next include Brownian rotation in the model and find that, in limit of high chemotactic sensitivity, the chemotactic drift is 64 % of the swimming velocity, independent of the magnitude of the Brownian rotation. We finally derive characteristic timescales of the motion that can be used to assess whether the diffusion limit is justified in a given situation. The proposed technique for obtaining diffusion approximations is conceptually and computationally simple, and applicable also when statistics of the motion is obtained empirically or through Monte Carlo simulation of the motion.

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

Similar content being viewed by others

References

  • Adler J (1966) Chemotaxis in bacteria. Science 153(3737):708–716

    Article  Google Scholar 

  • Alt W (1980) Biased random walk models for chemotaxis and related diffusion approximations. J Math Biol 9:147–177

    Article  MathSciNet  MATH  Google Scholar 

  • Anscombe F (1952) Large-sample theory of sequential estimation. Proc Camb Philos Soc 48:600–607

    Article  MathSciNet  MATH  Google Scholar 

  • Bearon R (2001) Run-and-tumble chemotaxis in an ambient fluid flow. PhD thesis, University of Cambridge

  • Bearon R (2003) An extension of generalized Taylor dispersion in unbounded homogeneous shear flows to run-and-tumble chemotactic bacteria. Phys Fluids 15:1552

    Article  MATH  Google Scholar 

  • Bearon R (2007) A model for bacterial colonization of sinking aggregates. Bull Math Biol 69(1):417–431

    Article  MATH  Google Scholar 

  • Bearon R, Pedley T (2000) Modelling run-and-tumble chemotaxis in a shear flow. Bull Math Biol 62:775–791

    Article  MATH  Google Scholar 

  • Benhamou S (2014) Of scales and stationarity in animal movements. Ecol Lett 17(3):261–272. doi:10.1111/ele.12225

    Article  MathSciNet  Google Scholar 

  • Berg H (1993) Random walks in biology, 2nd edn. Princeton University Press, Princeton

    Google Scholar 

  • Brown D, Berg H (1974) Temporal stimulation of chemotaxis in Escherichia coli. Proc Nat Acad Sci 71(4):1388–1392

    Article  Google Scholar 

  • Cluzel P, Surette M, Leibler S (2000) An ultrasensitive bacterial motor revealed by monitoring signaling proteins in single cells. Science 287(5458):1652–1655

    Article  Google Scholar 

  • Codling E, Bearon R, Thorn G (2010) Diffusion about the mean drift location in a biased random walk. Ecology 91(10):3106–3113

    Article  Google Scholar 

  • Colin R, Zhang R, Wilson L (2014) Fast, high-throughput measurement of collective behaviour in a bacterial population. J R Soc Interface 11(98):20140,486

    Article  Google Scholar 

  • Frankel I, Brenner H (1989) On the foundations of generalized Taylor dispersion theory. J Fluid Mech 204(–1):97–119

    Article  MATH  Google Scholar 

  • Grimmett G, Stirzaker D (1992) Probability and random processes, 2nd edn. Oxford University Press, Oxford

    MATH  Google Scholar 

  • Grünbaum D (1999) Advection–diffusion equations for generalized tactic searching behaviors. J Math Biol 38:169–194

    Article  MathSciNet  MATH  Google Scholar 

  • Grünbaum D (2000) Advection–diffusion equations for internal state-mediated random walks. SIAM J Appl Math 61(1):43–73

    Article  MathSciNet  MATH  Google Scholar 

  • Hastings A, Petrovskii S, Morozov A (2011) Spatial ecology across scales. Biol Lett 7(2):163–165. doi:10.1098/rsbl.2010.0948

    Article  Google Scholar 

  • Jackson G (1987) Simulating chemosensory response of marine microorganisms. Limnol Oceanogr 32:1253–1266

    Article  Google Scholar 

  • Jackson G (1989) Simulation of bacterial attraction and adhesion to falling particles in an aquatic environment. Limnol Oceanogr 34(3):514–530

    Article  Google Scholar 

  • Kay R, Langridge P, Traynor D, Hoeller O (2008) Changing directions in the study of chemotaxis. Nat Rev Mol Cell Biol 9(6):455–463

    Google Scholar 

  • Keller E, Segel L (1971) Model for chemotaxis. J Theor Biol 30:225–234

    Article  MATH  Google Scholar 

  • Kiørboe T, Jackson G (2001) Marine snow, organic solute plumes, and optimal chemosensory behavior of bacteria. Limnol Oceanogr 46(6):1309–1318

    Article  Google Scholar 

  • Kiørboe T, Ploug H, Thygesen UH (2001) Fluid motion and solute distribution around sinking aggregates. I. Small scale fluxes and heterogeneity of nutrients in the pelagic environment. Mar Ecol Prog Ser 211:1–13

    Article  Google Scholar 

  • Locsei J (2007) Persistence of direction increases the drift velocity of run and tumble chemotaxis. J Math Biol 55(1):41–60

    Article  MathSciNet  MATH  Google Scholar 

  • Nielsen BF, Nilsson LF, Thygesen UH, Beyer JE (2007) Higher order moments and conditional asymptotics of the batch markovian arrival process. Stoch Models 23(1):1–26

    Article  MathSciNet  MATH  Google Scholar 

  • Øksendal B (2010) Stochastic differential equations: an introduction with applications, sixth edn. Springer, Berlin

    MATH  Google Scholar 

  • Okubo A, Levin S (2001) Diffusion and ecological problems: modern perspectives. Springer, Berlin

    Book  MATH  Google Scholar 

  • Othmer H, Hillen T (2002) The diffusion limit of transport equations II: chemotaxis equations. SIAM J Appl Math 62(4):1222–1250

    Article  MathSciNet  MATH  Google Scholar 

  • Othmer H, Dunbar S, Alt W (1988) Models of dispersal in biological systems. J Math Biol 26:263–298

    Article  MathSciNet  MATH  Google Scholar 

  • Patlak C (1953) Random walk with persistence and external bias. Bull Math Biophys 15:311–338

    Article  MathSciNet  MATH  Google Scholar 

  • Rivero M, Tranquillo R, Buettner H, Lauffenburger D (1989) Transport models for chemotactic cell populations based on individual cell behavior. Chem Eng Sci 44(12):2881–2897

    Article  Google Scholar 

  • Sandoval M, Marath NK, Subramanian G, Lauga E (2014) Stochastic dynamics of active swimmers in linear flows. J Fluid Mech 742:50–70

    Article  MATH  Google Scholar 

  • Taylor G (1921) Diffusion by continuous movements. Proc Lond Math Soc 20:196–211

    MathSciNet  MATH  Google Scholar 

  • Thygesen U, Nilsson A, Andersen K (2007) Eulerian techniques for individual-based models based on additive processes. J Mar Syst 67:179–188. doi:10.1016/j.marsys.2006.10.005

    Article  Google Scholar 

  • Xue C (2015) Macroscopic equations for bacterial chemotaxis: integration of detailed biochemistry of cell signaling. J Math Biol 70(1–2):1–44

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Uffe Høgsbro Thygesen.

Appendices

Appendix 1: Proof of Proposition 1

Define the scalar stochastic process

$$\begin{aligned} z_i = c' x_i - \frac{c'\mu }{\tau } t_i \end{aligned}$$

where c is an arbitrary column vector and \(c'\) denotes transpose. \(z_i\) is a scalar auxiliary process which has been constructed so as to have mean \(\mathbf {E}z_i = 0\), thus enabling us to apply Anscombe’s theorem. Its variance is

$$\begin{aligned} \mathbf {V}z_i = \sigma ^2 = c ' \left( {\varSigma }_x - \frac{1}{\tau } \left( \mu \rho ' + \rho \mu '\right) + \frac{\mu \mu ' }{ \tau ^2} \sigma ^2_t \right) c = c'{\varSigma }c~~. \end{aligned}$$

Define

$$\begin{aligned} S_n = \sum _{i=1}^n z_i \end{aligned}$$

and note that by the law of large numbers, \(t^{-1} N(t) \rightarrow \tau ^{-1}\) almost surely. We may thus apply Anscombe’s theorem to conclude that \(S_{N(t)}\) is asymptotically normally distributed. Now, \(S_n\) is closely related to the cumulated reward \(\bar{X}_T\) which is our object of investigation:

$$\begin{aligned} S_{N(t)} = c' \bar{X}_t - \frac{c' \mu }{\tau } T_{N(t)} = c' \bar{X}_t - \frac{c' \mu }{\tau } t + \frac{c'\mu }{\tau } A_t \end{aligned}$$

where

$$\begin{aligned} A_t = t - T_{N(t)} \end{aligned}$$

is the age, i.e., the time that has passed since last tumble. The process \(A_t\) is bounded in probability, and hence

$$\begin{aligned} \frac{A_t}{\sqrt{t}} \rightarrow 0 \hbox { in probability as } t\rightarrow \infty . \end{aligned}$$

Thus, the asymptotic distribution of the sum \(S_{N(t)}\) implies a similar property of \(\bar{X}_t\):

$$\begin{aligned} c ' \frac{\bar{X}_t - \frac{\mu }{\tau } t}{\sqrt{t/\tau }} = \frac{S_{N(t)}}{\sqrt{t/\tau }} - \frac{\mu }{\sqrt{\tau }} \frac{A_t}{\sqrt{t}} ~~ \rightarrow N(0,\sigma ) \hbox { in distribution. } \end{aligned}$$

Since this holds for any c, \(\bar{X}_t\) itself must be asymptotically normally distributed:

$$\begin{aligned} \frac{1}{\sqrt{t}} \left( \bar{X}_t - \frac{\mu }{\tau } t \right) \rightarrow N(0,{\varSigma }/\tau ) \text{ in } \text{ distribution. } \end{aligned}$$

\(\square \)

Appendix 2: Proof of Theorem 2

We aim to apply Proposition 1 and therefore pursue the statistics of a single run. For simplicity, we rescale time and length so that the swimming speed is 1 and the base tumble rate is 1, and express the tumble rate as a function of the angle \(\theta \) between v and \(e_z\). Then, the tumble rate is

$$\begin{aligned} \lambda (\theta ) = e^{-\beta \cos \theta } \end{aligned}$$
(4)

At each tumble, a new run direction is chosen, uniformally with respect to solid angle, i.e., the angle \(\theta _i\) between the z-axis and the velocity during run i is distributed between 0 and \(\pi \) with density:

$$\begin{aligned} \phi (\theta ) = \frac{1}{2} \sin \theta ~~\hbox { for } 0\le \theta \le \pi . \end{aligned}$$

The duration \(t_i\) of a run is, when conditioning on the angle \(\theta _i\), exponentially distributed with mean \(1/\lambda (\theta _i)\). We then obtain the unconditional mean:

$$\begin{aligned} \tau = \mathbf {E}t_i = \mathbf {E}\left( \mathbf {E}\{ t_i | \theta _i \} \right) = \mathbf {E}\frac{1}{\lambda (\theta _i)} = \int _0^\pi \frac{\phi (\theta )}{\lambda (\theta )} ~d\theta = \frac{1}{2\beta } \left( e^\beta - e^{-\beta } \right) = \frac{\sinh \beta }{\beta }~~. \end{aligned}$$

This mean duration is a timescale which characterizes the process. It is examined in Sect. 4 and compared to other timescales, and graphed in Fig. 2.

Likewise, conditional on the angle \(\theta _i\), the second moment of the run duration is \(\mathbf {E}\{t_i^2 | \theta _i\} = 2/\lambda ^2(\theta _i)\). We thus find the unconditional second moment:

$$\begin{aligned} \mathbf {E}t_i^2 = \mathbf {E}\left( \mathbf {E}\left\{ t_i^2 | \theta _i \right\} \right) = \mathbf {E}\frac{2}{\lambda ^2(\theta _i)} = \int _0^\pi \frac{2\phi (\theta )}{\lambda ^2(\theta )} ~d\theta = \frac{1}{2\beta } \left( e^{2\beta }- e^{-2\beta } \right) = \frac{\sinh 2 \beta }{\beta } \end{aligned}$$

from which we obtain the variance of the run duration:

$$\begin{aligned} \sigma ^2_t = \mathbf {V}t_i = \mathbf {E}t_i^2 - (\mathbf {E}t_i)^2 = \frac{\sinh 2\beta }{\beta } - \frac{\sinh ^2 \beta }{\beta ^2} \end{aligned}$$

We now turn to the statistics of the displacement during a run, which we—with a slight abuse of notation—represent with coordinates \((x_i,y_i,z_i)\). We have the geometric relationship \(z_i =t_i \cdot \cos \theta _i\) which yields the mean displacement in the z-direction:

$$\begin{aligned} \mu _z = \mathbf {E}z_i = \mathbf {E}( \mathbf {E}\{ z_i | \theta _i \} ) =\mathbf {E}\frac{\cos \theta _i}{\lambda (\theta _i)} = \int _0^\pi \frac{\phi (\theta )}{\lambda (\theta )} \cos \theta ~d\theta = \frac{ \cosh \beta }{\beta } - \frac{\sinh \beta }{\beta ^2} \end{aligned}$$

Rotational symmetry yields \(\mu _x=\mu _y=0\). Proposition 1 now gives the drift in the diffusion approximation:

$$\begin{aligned} v_x = v_y = 0, \quad v_z = \frac{\mu _z}{\tau } = \coth \beta - \frac{1}{\beta } \end{aligned}$$

as shown in Fig. 1.

Next, we need the second-order properties of the displacement during a run. Rotational symmetry and Pythagoras’ theorem yield

$$\begin{aligned} \mathbf {E}x_i^2 = \mathbf {E}y_i^2 \hbox { and } x_i^2 + y_i^2 = t_i^2\sin ^2 \theta _i. \end{aligned}$$

Combining these two, we obtain the variance of displacement in the direction of the x-axis (or the y-axis):

$$\begin{aligned} \mathbf {E}x_i^2= & {} \frac{1}{2} \mathbf {E}\left( \mathbf {E}\left\{ \left( t_i^2 \sin ^2\theta _i\right) | \theta _i \right\} \right) \\= & {} \mathbf {E}\frac{\sin ^2 \theta _i}{\lambda ^2(\theta _i)} \\= & {} \int _0^\pi \frac{\phi (\theta ) \sin ^2 \theta }{\lambda ^2(\theta )} ~d\theta \\= & {} \frac{\cosh 2 \beta }{2\beta ^2} - \frac{\sinh 2 \beta }{4 \beta ^3} \end{aligned}$$

Correspondingly, for the z-direction:

$$\begin{aligned} \mathbf {E}z_i^2= & {} \mathbf {E}\left( \mathbf {E}\left\{ t_i^2 \cos ^2\theta _i | \theta _i \right\} \right) \\= & {} \mathbf {E}\frac{2 \cos ^2\theta _i }{\lambda ^2(\theta _i)} \\= & {} \int _0^\pi \frac{2 \phi (\theta ) ~\cos ^2\theta }{\lambda ^2(\theta )}~d\theta \\= & {} \frac{2\beta ^2 +1}{2\beta ^3} \sinh 2 \beta - \frac{\cosh 2 \beta }{\beta ^2} \end{aligned}$$

which yields the variance

$$\begin{aligned} \mathbf {V}z_i = \mathbf {E}z_i^2 - (\mathbf {E}z_i)^2 = \frac{\left( 4 \beta ^3 + 6 \beta \right) \sinh 2 \beta - \left( 6\beta ^2 + 2\right) \cosh 2 \beta + 2 - 2\beta ^2}{4\beta ^4}. \end{aligned}$$

Rotational symmetry dictates that the run displacement \((x_i,y_i,z_i)\) has diagonal covariance matrix \( {\varSigma }_x = \hbox {diag}(\mathbf {V}x_i,\mathbf {V}y_i,\mathbf {V}z_i ) \) and that the covariance between run duration and displacement is 0 in the two directions \(e_x\) and \(e_y\), so

$$\begin{aligned} \rho = (0,0,\mathbf {Cov}(z_i,t_i))'. \end{aligned}$$

Here, the covariance between run duration \(t_i\) and displacement \(z_i\) in z-direction is obtained with similar calculations:

$$\begin{aligned} \mathbf {E}t_i z_i= & {} \mathbf {E}\left( \mathbf {E}\left\{ t_iz_i | \theta _i \right\} \right) \\= & {} \mathbf {E}\left( \mathbf {E}\left\{ t_i^2 \cos \theta _i | \theta _i \right\} \right) \\= & {} \mathbf {E}\frac{2\cos \theta _i}{\lambda ^2(\theta _i)} \\= & {} \int _0^\pi \frac{2 \phi (\theta ) \cos \theta }{ \lambda ^2(\theta ) } ~d\theta \\= & {} \frac{\cosh \beta }{\beta } - \frac{\sinh 2 \beta }{2 \beta ^2} \end{aligned}$$

yielding

$$\begin{aligned} \mathbf {Cov}(t_i,z_i) = \mathbf {E}t_i z_i - \tau \mu _z = \frac{\left( 2\beta ^2 + 1\right) \cosh 2\beta - 2 \beta \sinh 2 \beta -1}{2\beta ^3} ~~. \end{aligned}$$

We now have the statistics which are required to apply Proposition 1. In the x- and y-directions, we get

$$\begin{aligned} D_x =D_y = \frac{\mathbf {E}x_i^2}{2 \tau } = \frac{ 2 \beta \cosh 2 \beta - \sinh 2 \beta }{ 8 \beta ^2 \sinh \beta } \end{aligned}$$

while in the z direction, we get

$$\begin{aligned} D_z = \frac{\mathbf {V}z_i + \frac{(\mathbf {E}z_i)^2}{\tau ^2}\mathbf {V}t_i -2\frac{\mathbf {E}z_i}{\tau } \mathbf {Cov}(t_i,z_i)}{2\tau } = \frac{1}{4} \frac{\sinh 2\beta - 2 \beta }{\beta ^2 \sinh \beta }. \end{aligned}$$

Proposition 1 now establishes the limiting Gaussian distribution of the position at tumbles, and it remains only to show that the same limiting distribution applies during a run: The displacement during a run from the position at last tumble will be bounded in probability. Therefore, \((\bar{X}_t - X_t)/\sqrt{t}\) converges to 0 in probability, so that \((X_t - v t)/\sqrt{t}\) and \((\bar{X}_t - vt)/\sqrt{t}\) has the same limiting distribution, i.e., the limiting Gaussian distribution applies also to the position at any point in time, whether at a tumble or during a run. \(\square \)

Appendix 3: Monte Carlo Simulation

Here, we describe the Monte Carlo simulation performed to verify the results of Theorem 2, i.e., the estimates and associated confidence intervals displayed in Fig. 1. The run–tumble model with constant speed \(U=1\), base tumble rate \(\lambda _0=1\), and different values of the parameter \(\beta \) is simulated stochastically. The initial distribution for the velocity is the stationary distribution which scales with \(1/\lambda (v)\); this is a truncated exponential distribution for \(v_z=v\cdot e_z\) and conditional on this, a uniform distribution of the two other components on the circle \(\{ (v_x,v_y) : v_x^2 + v_y^2 = 1- v_z^2\}\). The motion is simulated one tumble at a time until the time of last tumble exceeds the simulation time \(T=10000\); after this, the position at time T is obtained by interpolation in the last run and reported. This simulation time is orders of magnitude larger than the decorrelation time of the velocity process (compare Fig. 2) which ensures that the diffusion approximation is reasonable. \(N=1000\) independent replicates of the motion are simulated for each value of \(\beta \). Estimates of the drift and diffusivity are obtained as sample means computed from mean displacement and its variance, respectively. 95 % confidence intervals for the mean are obtained using \(\pm 1.96\) times the standard error between replicates; these intervals are smaller than the plotting symbols so practically invisible in the plot. 95 % confidence intervals for the diffusivity are obtained from percentiles in \(M=1,000\) bootstrap replicates.

Appendix 4: Proof of Theorem 3

We first nondimensionalize the model; this amounts to taking \(U=1\) and \(\kappa =1\). We aim to apply Proposition 1 and therefore examine the statistics of time and displacement during a single run. A run starts at time 0 with a random velocity \(V_0\) sampled uniformly on the half-sphere \(B_+\) with respect to solid angle and ends when the Brownian motion on the half-sphere hits the plane \(\{ v: e_z v = 0\}\).

We first transform the problem to a one-dimensional one by focusing on the velocity in the z-direction. During a run, the instantaneous velocity vector \(V_t\) is Brownian motion on the surface of the unit sphere, which can be written as the solution to an Itô stochastic differential equation (Øksendal 2010)

$$\begin{aligned} {\text {d}}V_t = - V_t ~{\text {d}}t + \left( I - V_tV_t' \right) ~ {\text {d}}B_t \end{aligned}$$
(5)

where \(B_t\) is three-dimensional Brownian motion. Let \(U_t = e_z V_t\) denote the velocity in the direction of the gradient. Using Itô’s lemma, we obtain

$$\begin{aligned} {\text {d}}U_t = - U_t ~{\text {d}}t + \sqrt{ 1 - |U_t|^2 } ~ {\text {d}}W_t \end{aligned}$$

Here,

$$\begin{aligned} {\text {d}}W_t = \frac{1}{\sqrt{1-|U_t|^2}} e_z ( I - V_t V_t') ~ {\text {d}}B_t \end{aligned}$$

and, noting that

$$\begin{aligned} \left| \frac{1}{\sqrt{1-|U_t|^2}} e_z ( I - V_t V_t') \right| ^2 = 1 \end{aligned}$$

we recognize \(W_t\) as standard Brownian motion (compare Øksendal (2010)). Thus, the process \(U_t\) is itself a Markov process with generator L given by

$$\begin{aligned} (L f)(u) = - u \frac{\partial f}{\partial u} + \frac{1}{2} \left( 1-u^2\right) \frac{\partial ^2f}{\partial u^2} = \frac{\partial }{\partial u} \left( \frac{1}{2} \left( 1-u^2\right) \frac{\partial f}{\partial u} \right) \end{aligned}$$

Up to the factor 1 / 2, this is Legendre’s operator of order 0.

From the point of view of the individual run, the boundary \(u=0\) is absorbing in that the run terminates when \(U_t\) reaches this boundary. Let \(T=\inf \{ t : U_t \le 0 \}\) denote this stopping time. We seek the statistics of \(T\) and \(Z_T\). During the run, the time t and up-gradient displacement \(Z_t\) are additive processes, studied in an oceanographic context by Thygesen et al. (2007). Define the moment functions

$$\begin{aligned} \psi _{ij}(u) = \mathbf {E}^u T^i Z_T^j \end{aligned}$$

then it follows that these functions \(\psi _{ij}\) for \(i,j\in \mathbf {N}_0\), \((i,j)\ne (0,0)\) are governed by boundary value problems

$$\begin{aligned} L\psi _{10} + 1= & {} 0 \\ L\psi _{01} + u= & {} 0 \\ L\psi _{ij} + i \psi _{(i-1)j} + j u \psi _{i(j-1)}= & {} 0 \hbox { for } i,j \in \mathbf {N}\end{aligned}$$

on the domain \(u\in (0,1)\). The boundary condition at \(u=0\) is \(\psi _{ij}(0)=0\) while the boundary condition at \(u=1\) is that the limit \(\lim _{u\rightarrow 1} \psi _{ij}(u)\) exists, since the operator L is singular at that point (recall that \(u=1\) does not correspond to a boundary point for the original Brownian motion on the half-sphere \(B_+\)).

For the low-order moments, we find the solutions:

$$\begin{aligned} \psi _{10}(u)= & {} \mathbf {E}^u T= 2 \log (1+u) \\ \psi _{01}(u)= & {} \mathbf {E}^u Z_T= u \\ \psi _{20}(u)= & {} \mathbf {E}^u T^2 = 8(\log 2 -1 ) \log (u + 1) - 4 (\log (2))^2 - 8 \hbox {dilog}\left( \frac{u}{2} + \frac{1}{2}\right) + 2 \frac{\pi ^2}{3} \\ \psi _{11}(u)= & {} \mathbf {E}^u TZ_T= (4+2z)\log (u+1) -2u \\ \psi _{02}(u)= & {} \mathbf {E}^u Z_T^2 = \frac{2}{3} u^2 +\frac{4}{3} \log (u+1) \end{aligned}$$

Here, the dilog function is \(\hbox {dilog} (x) = \int _1^x (1-t)^{-1} \log t ~{\text {d}}t\).

These statistics are conditional on the up-gradient velocity \(u=U_0\) at the beginning of the run. To find the unconditional statistics, we take expectation with respect to the distribution of \(U_0\). Since \(V_0\) is uniformly distributed on the half-sphere \(B_+\) with respect to solid angle, \(U_0\) is uniformly distributed on [0, 1]. Hence,

$$\begin{aligned} \mathbf {E}T^i Z_T^j = \int _0^1 \psi _{ij}(u)~{\text {d}}u \end{aligned}$$

We find:

$$\begin{aligned} \mathbf {E}T= & {} 4 \log 2 - 2 \approx 0.77 \\ \mathbf {E}Z_T= & {} \frac{1}{2} \\ \mathbf {E}T^2= & {} 16 (1- \log 2 )^2 \approx 1.51 \\ \mathbf {E}TZ_T= & {} -9/2 + 8 \log 2 \approx 1.05 \\ \mathbf {E}Z^2_T= & {} -\frac{10}{9} + \frac{8}{3} \log 2 \approx 0.74 \end{aligned}$$

According to Proposition 1, we can now compute the advective velocity as

$$\begin{aligned} \frac{\mathbf {E}Z_T}{\mathbf {E}T} = \frac{1}{8 \log 2 - 4} \approx 0.64 \end{aligned}$$

To determine the diffusivity, we need the variance of the run duration

$$\begin{aligned} \mathbf {V}T= \mathbf {E}T^2 - (\mathbf {E}T)^2 = 12 - 16 \log 2 \approx 0.91 \quad , \end{aligned}$$

the variance of the up-gradient displacement during a run

$$\begin{aligned} \mathbf {V}Z_T= \mathbf {E}Z_T^2 - (\mathbf {E}Z_T)^2 = - \frac{49}{36} + \frac{8}{3} \log 2 \approx 0.49 \quad , \end{aligned}$$

and the covariance

$$\begin{aligned} \mathbf {Cov}(T, Z_T) = \mathbf {E}TZ_T- (\mathbf {E}Z_T)(\mathbf {E}T) = -7/2 + 6\log 2 \approx 0.66 \quad . \end{aligned}$$

With this, we can employ Proposition 1 to obtain the diffusivity in the z-direction:

$$\begin{aligned} D_z= & {} \frac{1}{2} \frac{\mathbf {V}Z_T+ \frac{(\mathbf {E}Z_T)^2}{(\mathbf {E}T)^2} \mathbf {V}T- 2 \mathbf {E}Z_T\frac{\rho }{\mathbf {E}T} }{\mathbf {E}T}\\= & {} {\frac{1}{144}}\,{\frac{-85+490\,\ln \left( 2 \right) -796\, \left( \ln \left( 2 \right) \right) ^{2}+384\, \left( \ln \left( 2 \right) \right) ^{3}}{ \left( -1+2\,\ln \left( 2 \right) \right) ^{3}}}\\\approx & {} 1.00 \cdot 10 ^{-2} \end{aligned}$$

By rotational symmetry, the chemotactic drift in the x- (or y)-direction is zero, and the diffusivity matrix is diagonal. It remains to find the diffusivity \(D_x\) in the x-direction, and to this end, we must find \(\mathbf {V}X_T\). Conditional on \(V_0\), this is governed by the two boundary value problems on the half-sphere \(B_+\):

$$\begin{aligned} \nabla ^2 \gamma _1 + x = 0,\quad \nabla ^2 \gamma _2 + 2x\gamma _1 = 0 \end{aligned}$$

with Dirichlet boundary condition on the boundary \(\{ (x,y,z): z=0, x^2+y^2 = 1 \}\). These equations do not appear to admit analytical solution in closed form. Instead, two numerical approaches were used: first, numerical solution of the two equations using the finite element software Comsol and next, a Monte Carlo method where the stochastic differential Eq. (5) is simulated using a modified Euler method until absorption. The two numerical methods both estimate

$$\begin{aligned} \mathbf {E}X_T^2 = 0.131 \pm 0.001 \end{aligned}$$

With this, we find

$$\begin{aligned} D_x = \frac{\mathbf {E}X_T^2}{2 \mathbf {E}T} \approx 8.54 \cdot 10^{-2} \quad . \end{aligned}$$

Finally, according to the general result (3), the expected age in stationarity is

$$\begin{aligned} \mathbf {E}A_t = \frac{1}{2} \frac{\mathbf {V}T}{\mathbf {E}T} + \frac{1}{2} \mathbf {E}T= \frac{3 - 4 \log 2}{2 \log 2 - 1} + 2 \log 2 - 1 \approx 0.975 \quad . \end{aligned}$$

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Thygesen, U.H. A Diffusion Approximation Based on Renewal Processes with Applications to Strongly Biased Run–Tumble Motion. Bull Math Biol 78, 556–579 (2016). https://doi.org/10.1007/s11538-016-0155-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-016-0155-3

Keywords

Navigation