Skip to main content
Log in

Exit Time Distribution in Spherically Symmetric Two-Dimensional Domains

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

The distribution of exit times is computed for a Brownian particle in spherically symmetric two-dimensional domains (disks, angular sectors, annuli) and in rectangles that contain an exit on their boundary. The governing partial differential equation of Helmholtz type with mixed Dirichlet–Neumann boundary conditions is solved analytically. We propose both an exact solution relying on a matrix inversion, and an approximate explicit solution. The approximate solution is shown to be exact for an exit of vanishing size and to be accurate even for large exits. For angular sectors, we also derive exact explicit formulas for the moments of the exit time. For annuli and rectangles, the approximate expression of the mean exit time is shown to be very accurate even for large exits. The analysis is also extended to biased diffusion. Since the Helmholtz equation with mixed boundary conditions is encountered in microfluidics, heat propagation, quantum billiards, and acoutics, the developed method can find numerous applications beyond exit processes.

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

Similar content being viewed by others

References

  1. Redner, S.: A Guide to First-Passage Processes. Cambridge University Press, Cambridge (2001)

    Book  MATH  Google Scholar 

  2. Condamin, S., Bénichou, O., Tejedor, V., Voituriez, R., Klafter, J.: First-passage times in complex scale-invariant media. Nature 450(7166), 77–80 (2007)

    Article  ADS  Google Scholar 

  3. Bénichou, O., Chevalier, C., Meyer, B., Voituriez, R.: Facilitated diffusion of proteins on chromatin. Phys. Rev. Lett. 106, 38102 (2011)

    Article  Google Scholar 

  4. Sheinman, M., Bénichou, O., Kafri, Y., Voituriez, R.: Classes of fast and specific search mechanisms for proteins on DNA. Rep. Prog. Phys. 75(2), 026601 (2012)

    Article  ADS  Google Scholar 

  5. Mazzolo, A.: Properties of diffusive random walks in bounded domains. Europhys. Lett. (EPL) 68(3), 350–355 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  6. Reingruber, J., Holcman, D.: Diffusion in narrow domains and application to phototransduction. Phys. Rev. E (Statistical, Nonlinear, and Soft Matter Physics) 79(3), 30904 (2009)

    Article  ADS  Google Scholar 

  7. Meyer, B., Bénichou, O., Kafri, Y., Voituriez, R.: Geometry-induced bursting dynamics in gene expression. Biophys. J. 102(9), 2186–2191 (2012)

    Article  ADS  Google Scholar 

  8. Singer, A., Schuss, Z., Holcman, D.: Narrow escape, Part II: the circular disk. J. Stat. Phys. 122(3), 465–489 (2006)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  9. Pillay, S., Ward, M.J., Peirce, A., Kolokolnikov, T.: An asymptotic analysis of the mean first passage time for narrow escape problems: part I: two-dimensional domains. Multiscale Model. Simul. 8(3), 803–835 (2009)

    Article  MathSciNet  Google Scholar 

  10. Chevalier, C., Bénichou, O., Meyer, B., Voituriez, R.: First-passage quantities of Brownian motion in a bounded domain with multiple targets: a unified approach. J. Phys. A 44, 25002 (2011)

    Article  Google Scholar 

  11. Isaacson, Samuel A., Newby, Jay: Uniform asymptotic approximation of diffusion to a small target. Phys. Rev. E 88(1), 012820 (2013)

    Article  ADS  Google Scholar 

  12. Bénichou, O., Chevalier, C., Klafter, J., Meyer, B., Voituriez, R.: Geometry-controlled kinetics. Nat. Chem. 2(6), 472–477 (2010)

    Article  Google Scholar 

  13. Meyer, B., Chevalier, C., Voituriez, R., Bénichou, O.: Universality classes of first-passage-time distribution in confined media. Phys. Rev. E 83(5), 51116 (2011)

    Article  ADS  Google Scholar 

  14. Nguyen, Binh T., Grebenkov, Denis S.: A spectral approach to survival probabilities in porous media. J. Stat. Phys. 141(3), 532–554 (2010)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  15. Sneddon, I.N.: Mixed Boundary Value Problems in Potential Theory. North-Holland, Amsterdam (1966)

    MATH  Google Scholar 

  16. Caginalp, C., Chen, X.: Analytical and numerical results for an escape problem. Arch. Ration. Mech. Anal. 203, 329–342 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  17. Benichou, O., Voituriez, R.: Narrow-escape time problem: time needed for a particle to exit a confining domain through a small window. Phys. Rev. Lett. 100(16), 168104–168105 (2008)

    Article  ADS  Google Scholar 

  18. Mattos, Thiago G., Mejía-Monasterio, Carlos, Metzler, Ralf, Oshanin, Gleb: First passages in bounded domains: when is the mean first passage time meaningful? Phys. Rev. E 86(3), 031143 (2012)

    Article  ADS  Google Scholar 

  19. Singer, A., Schuss, Z., Holcman, D.: Narrow escape, Part III: non-smooth domains and Riemann surfaces. J. Stat. Phys. 122(3), 491–509 (2006)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  20. Lagache, T., Holcman, D.: Effective motion of a virus trafficking inside a biological cell. SIAM J. Appl. Math. 68(4), 1146–1167 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  21. Carslaw, H.: Conduction of Heat in Solids. Clarendon, Oxford (1959)

    Google Scholar 

  22. Crank, J.: The Mathematics of Diffusion. Oxford Science Publications, Oxford (1975)

    Google Scholar 

  23. Sbragaglia, M., Prosperetti, A.: A note on the effective slip properties for microchannel flows with ultrahydrophobic surfaces. Phys. Fluids 19(4), 043603 (2007)

    Article  ADS  Google Scholar 

  24. Joseph, P., Cottin-Bizonne, C., Benoît, J.-M., Ybert, C., Journet, C., Tabeling, P., Bocquet, L.: Slippage of water past superhydrophobic carbon nanotube forests in microchannels. Phys. Rev. Lett. 97(15), 156104 (2006)

    Article  ADS  Google Scholar 

  25. Cottin-Bizonne, C., Barentin, C., Charlaix, E., Bocquet, L., Barrat, J.-L.: Dynamics of simple liquids at heterogeneous surfaces: molecular-dynamics simulations and hydrodynamic description. Eur. Phys. J. E Soft Matter 15(4), 427–438 (2004)

    Article  Google Scholar 

  26. Gardiner, C.W.: Handbook of Stochastic Methods for Physics, Chemistry and Natural Sciences. Springer, New York (2004)

    Book  MATH  Google Scholar 

  27. Abramowitz, Milton: Handbook of mathematical functions with formulas, graphs, and mathematical tables. Am. J. Phys. 56(10), 958 (1988)

    Article  ADS  Google Scholar 

  28. Grebenkov, D.: Efficient Monte Carlo methods for simulating diffusion-reaction processes in complex systems. arXiv:1304.7807 (2013)

  29. Berezhkovskii, A.M., Barzykin, A.V.: Extended narrow escape problem: boundary homogenization-based analysis. Phys. Rev. E 82(1), 011114 (2010)

    Article  ADS  Google Scholar 

  30. Bénichou, O., Grebenkov, D., Levitz, P., Loverdo, C., Voituriez, R.: Optimal reaction time for surface-mediated diffusion. Phys. Rev. Lett. 105, 150606 (2010)

    Article  Google Scholar 

  31. Bénichou, O., Grebenkov, D., Levitz, P., Loverdo, C., Voituriez, R.: Mean first-passage time of surface-mediated diffusion in spherical domains. J. Stat. Phys. 142(4), 657–685 (2011)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  32. Rupprecht, J.F., Bénichou, O., Grebenkov, D., Voituriez, R.: Kinetics of active surface-mediated diffusion in spherically symmetric domains. J. Stat. Phys. 147(5), 891–918 (2012)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  33. Pryor, R.W.: Multiphysics Modeling Using COMSOL: A First Principles Approach. Jones & Bartlett Learning, Sudbury (2009)

    Google Scholar 

  34. Castro, L.P., Speck, F.O., Teixeira, F.S.: Mixed boundary value problems for the helmholtz equation in a quadrant. Integral Equ. Oper. Theory 56(1), 1–44 (2005)

    Article  MathSciNet  Google Scholar 

  35. Grebenkov, D.: Efficient Monte Carlo methods for simulating diffusion-reaction processes in complex systems. In: First-Passage Phenomena and their Applications. World Scientific Publishing Company, Singapore (2013)

  36. Temkin, S.: Elements of Acoustics. American Institute of Physics, Woodbury (2001)

    Google Scholar 

  37. Duffy, D.G.: Mixed Boundary Value Problems. Chapman & Hall, London (2007)

  38. Valsa, J.: Invlap package (from Matlabcentral/fileexchange), (2011)

Download references

Acknowledgments

O.B. is supported by the ERC Starting Grant No. FPTOpt-277998. D.G. is supported by an ANR project “INADILIC”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J.-F. Rupprecht.

Appendices

Appendix 1: Simplification of \(\alpha _{n}\) and \(M_{nm}\)

1.1 Simplified Expressions for \(\alpha _{0}\)

Herewith we prove the following identity for all \(0 \le t < \pi \)

$$\begin{aligned} \frac{\sqrt{2}}{\pi } \int _{0}^{t} \mathrm {d}x \frac{x \sin (x/2)}{\sqrt{\cos x - \cos t }} = - 2 \ln \left( \cos \left( \frac{t}{2}\right) \right) . \end{aligned}$$
(111)

We proceed by a change of variable \(z = \cos x\) in the left-hand side term of Eq. (111) and we denote \(T = \cos (t)\):

$$\begin{aligned} \int _{0}^{t} \mathrm {d}x \frac{x \sin (x/2)}{\sqrt{\cos x - \cos t }} = \int _{T}^{1} \mathrm {d}z \left[ \frac{\arccos z}{\sqrt{2\left( 1+z\right) }} \right] \frac{1}{\sqrt{z - T }}. \end{aligned}$$
(112)

We write the right-hand side of Eq. (111) in the form

$$\begin{aligned} - 2 \log \left( \cos \left( \frac{t}{2} \right) \right) = \log \left( \frac{2}{1+ T} \right) . \end{aligned}$$
(113)

From Ref. [15], the Abel’s equation

$$\begin{aligned} \int ^{1}_{T} \frac{y(z) \mathrm {d}z}{\sqrt{ z - T}} = \frac{\pi }{\sqrt{2}} \log \left( \frac{2}{1+ T} \right) \end{aligned}$$
(114)

has an unique solution for all \( -1 < X < 1\)

$$\begin{aligned} y(z) = \frac{1}{\pi } \frac{\pi }{\sqrt{2}} \int ^{z}_{1} \frac{\mathrm {d}u}{\sqrt{u -z}(1+z)} = \frac{\arccos {z}}{\sqrt{2 (1 + z)}}. \end{aligned}$$
(115)

Identification of the kernels of Eqs. (112) and (115) proves the identity (111).

The expression for \(\alpha _{0}\) from Eq. (27a) is found by setting \(t = \pi - \epsilon \) in Eq. (112). Note that the obtained expression for \(\alpha _{0}\) from Eq. (27a) could also be deduced from the expression of the MFPT from Ref. [16].

1.2 Simplified Expressions for \(\alpha _{n}, n \ge 1\)

The solution of Eqs. (24a), (24b) is given in [8] in the form:

$$\begin{aligned} \alpha _{n} = \frac{1}{\sqrt{2} \pi } \int _{0}^{\pi - \epsilon }\! \! \! \mathrm {d}t \left( \frac{\partial }{\partial t} \int _{0}^{t} \mathrm {d}x \frac{x \sin (x/2)}{\sqrt{\cos x - \cos t }} \right) \bigl [ P_n(\cos t) + P_{n-1}(\cos t) \bigr ], \qquad n \ge 1. \end{aligned}$$
(116)

Using the identity (111), we show that

$$\begin{aligned} \alpha _{n} = \frac{1}{2} \int _{0}^{\pi - \epsilon } \! \! \! \mathrm {d}t \tan \left( \frac{t}{2}\right) \bigl [ P_n(\cos t) + P_{n-1}(\cos t) \bigr ], \qquad n \ge 1. \end{aligned}$$
(117)

After the change of variable \( u = \cos t\), the latter identity leads to

$$\begin{aligned} \alpha _{n} = \frac{1}{2}\int ^{1}_{-\cos \epsilon } \frac{\mathrm {d}u}{1+u} \bigl [ P_n(u) + P_{n-1}(u)\bigr ] , \qquad n \ge 1. \end{aligned}$$
(118)

We now use the identity

$$\begin{aligned} \frac{\partial }{\partial x} \left( \frac{P_{m}(x)-P_{m-1}(x)}{m}\right) _{|x = X} = \frac{P_{m}(X) + P_{m-1}(X)}{1 + X}, \qquad m \ge 1, \end{aligned}$$
(119)

which is valid for all \(X \in \left[ - 1 , 1 \right] \), to obtain the announced result:

$$\begin{aligned} \alpha _{n}&= \frac{(-1)^{n-1}}{2 n} \bigl [ P_{n}(\cos \epsilon ) + P_{n-1}(\cos \epsilon ) \bigr ] ,\qquad n \ge 1. \end{aligned}$$
(120)

1.3 Simplified Expression for \(M_{nm}\)

The expression for \(M_{nm}\) from Eq. (46) can be simplified using Mehler’s integral representation (35):

$$\begin{aligned} M_{nm}&= \frac{1}{2} \int _{0}^{\pi - \epsilon } \! \! \! \mathrm {d}t \left\{ \frac{\partial }{\partial t} \left[ P_{m}(\cos t)-P_{m-1}(\cos t)\right] \right\} \bigl [ P_n(\cos t) + P_{n-1}(\cos t) \bigr ],\nonumber \\&\quad n \ge 1, \quad m \ge 1. \end{aligned}$$
(121)

The identity Eq. (119) then leads to the announced expression (47). Notice that the following matrix

$$\begin{aligned} S_{nm} = \sqrt{\frac{n}{m}}M_{nm}, \qquad n \ge 1, \quad m \ge 1, \end{aligned}$$
(122)

is symmetric. The eigenvalues of \(S_{nm}\) are real, and so are the eigenvalues of \(M\). In Sect. 5.5 we show that the coefficients \(\alpha _n\) are given by an eigenvector of the infinite-dimensional matrix \(M\).

1.4 Perturbative Expansion of \(M_{nm}\)

We first derive an alternative identity to Eq. (119). Let us define

$$\begin{aligned} A_n(x) \equiv \frac{\partial }{\partial x} (P_{n}(x) - P_{n-1}(x)), \qquad n \ge 1. \end{aligned}$$
(123)

Using the recurrence formulas for Legendre polynomials, we obtain

$$\begin{aligned} A_n(x)&= (P'_{n}(x) -P'_{n-2}(x))- (P'_{n-1}(x) -P'_{n-2}(x))\nonumber \\&= (2n-1)P_{n-1}(x) - A_{n-1}(x)= \sum ^{n}_{k=1} (-1)^{n-k} (2k-1)P_{k-1}(x), \end{aligned}$$
(124)

where we used \(P_{0}(x) = 1\), \(P_{1}(x) = x\), and \(A_1(x) = 1 = (2-1) P_{0}(x)\).

Combining Eq. (121) and the identity (124) we obtain the announced result:

$$\begin{aligned} M_{nm}&= \int _{0}^{\pi - \epsilon } \! \mathrm {d}t \ \frac{\sin t}{2} A_m (\cos t) \bigl [ P_n(\cos t) + P_{n-1}(\cos t) \bigr ], \nonumber \\&= \sum ^{m}_{k=1} (-1)^{n-k} (2k - 1) \left( K_{k-1,n-1} + K_{k-1,n} \right) , \end{aligned}$$
(125)

where the coefficients \(K_{k,n}\) are defined by

$$\begin{aligned} K_{k,n} \equiv \int _{0}^{\pi - \epsilon } \! \! \! \mathrm {d}t \ P_k(\cos t) P_n(\cos t) \frac{\sin t}{2}. \end{aligned}$$
(126)

In the leading order in \(\epsilon \ll 1\), Eq. (125) reads

$$\begin{aligned} M_{nm} = \sum ^{m}_{k=1} (-1)^{n-k} (2k - 1) \left( \frac{\delta _{k-1,n-1}}{2n-1} + \frac{\delta _{k-1,n}}{2n+1} \right) + \mathcal {O}(\epsilon ). \end{aligned}$$
(127)

If \(m < n\), it is straightforward to show that \(M_{nm} = \mathcal {O}(\epsilon )\). If \(m > n\), \(M_{nm} = \mathcal {O}(\epsilon )\) as successive terms with \(k = n\) and \( k = n +1\) cancel each other. The matrix \(M_{nm}\) (\(n,m \ge 1\)) is thus diagonal at the first order in \(\epsilon \):

$$\begin{aligned} M_{nm} = \delta _{nm} + \mathcal {O}(\epsilon ). \end{aligned}$$
(128)

In order to get the next term in the series expansion in \(\epsilon \ll 1\), we write

$$\begin{aligned} M_{nm} - \delta _{nm}&= - \frac{m}{2} \int ^{-\cos \epsilon }_{-1} \frac{1}{1+x} \bigl [ P_m(x)+P_{m-1}(x) \bigr ] \bigl [ P_n(x)+P_{n-1}(x) \bigr ] \mathrm {d}x,\nonumber \\&\quad n \ge 1, \quad m \ge 1. \end{aligned}$$
(129)

We now focus on the term in the right-hand side of Eq. (129). In the vicinity of \(x = -1\), the integrand of Eq. (129) expands into

$$\begin{aligned} \frac{1}{1+x} \bigl [ P_m(x)+P_{m-1}(x) \bigr ] \bigl [ P_n(x)+P_{n-1}(x) \bigr ] = \frac{nm (-1)^{n+m}}{8} (1 +x)+ \mathcal {O}(1+x). \end{aligned}$$
(130)

Notice that

$$\begin{aligned} \int ^{-\cos \epsilon }_{-1} (1 + x) \mathrm {d}x = \frac{\epsilon ^4}{8} + \mathcal {O}(\epsilon ^5). \end{aligned}$$
(131)

Substituting Eqs. (130) and (131) into Eq. (129) leads to

$$\begin{aligned} M_{nm} = \delta _{mn} + \frac{nm^2 (-1)^{n+m}}{8} \epsilon ^4 + \mathcal {O}(\epsilon ^5). \end{aligned}$$
(132)

1.5 Summation Identities

Using the identities for sums of Legendre polynomials from Ref. [37], we derive the following equation

$$\begin{aligned} S&\equiv \sum ^{\infty }_{m = 1} (-1)^{m-1} \bigl [ P_{m}(\cos x) + P_{m-1}(\cos x) \bigr ] \bigl [ P_{m}(\cos \epsilon ) + P_{m-1}(\cos \epsilon ) \bigr ] = 2,\nonumber \\&\quad 0 < x < \pi - \epsilon . \end{aligned}$$
(133)

To prove this identity, we first use the Mehler’s representation (35) for Legendre polynomials \(P_{m}(\cos \epsilon )\) and \(P_{m-1}(\cos \epsilon )\) to obtain

$$\begin{aligned} S&= \frac{2}{\pi } \int ^{\pi - \epsilon }_{0} \frac{ \sum ^{\infty }_{m = 1} \bigl [ P_{m}(\cos x) + P_{m-1}(\cos x) \bigr ] \left[ \cos (\left( m + \frac{1}{2}\right) t)- \cos (\left( m - \frac{1}{2}\right) t)\right] }{ \sqrt{(\cos (t)+\cos \epsilon )(\cos x- \cos (t))} },\nonumber \\&\quad 0 < x < \pi . \end{aligned}$$
(134)

We then use trigonometric identities and the series identity (39) to obtain the following integral representation

$$\begin{aligned} S = -\frac{2}{\pi } \int ^{\pi - \epsilon }_{x} \! \! \! \mathrm {d}t \frac{2 \sin \left( \frac{t}{2}\right) \cos \left( \frac{t}{2}\right) }{\sqrt{(\cos (t)+\cos \epsilon )(\cos x- \cos (t))}}, \qquad 0 < x < \pi - \epsilon . \end{aligned}$$
(135)

The consecutive change of variables \(z = \cos (t)\) and \(U = \sqrt{(z+\cos \epsilon )(\cos x- z)}\) leads to

$$\begin{aligned} S = \frac{2}{\pi } \int ^{\cos x}_{-\cos x} \frac{\mathrm {d}z}{\sqrt{(z+\cos \epsilon )(\cos x- z)}} = 2, \end{aligned}$$
(136)

which proves the identity (133). We will use this identity in the following form:

$$\begin{aligned} \sum ^{\infty }_{m = 1} \bigl [ P_{m}(\cos x) + P_{m-1}(\cos x) \bigr ] m \ \alpha _m = 1, \qquad 0 < x < \pi - \epsilon . \end{aligned}$$
(137)

where \(\alpha _m\) are given by Eq. (27b).

1.5.1 Proof of the Identity (83)

We write \(\alpha _m\) from Eq. (25b) and exchange the sum and the integral to obtain:

$$\begin{aligned} \sum ^{\infty }_{m=1} 2 m \alpha ^2_m&= \frac{\sqrt{2}}{\pi } \int _{0}^{\pi - \epsilon } \! \! \! \mathrm {d}t \left( \frac{\partial }{\partial t} \int _{0}^{t} \mathrm {d}x \frac{x \sin (x/2)}{\sqrt{\cos x - \cos t }} \right) \nonumber \\&\quad \times \left( \sum ^{\infty }_{m=1} \bigl [ P_m(\cos t) + P_{m-1}(\cos t) \bigr ] m \alpha _m \right) . \end{aligned}$$
(138)

Using the identity (133) in the right-hand side of Eq. (138) and the representation (25a) of \(\alpha _0\) leads to the result of Eq. (83).

1.5.2 An Eigenvector of the Matrix \(M\)

We show that the coefficients \(\alpha _n\) (with \(n \ge 1\)) form an eigenvector of the matrix \(M\):

$$\begin{aligned} \sum ^{\infty }_{m=1} M_{nm} \alpha _m= \alpha _n. \end{aligned}$$
(139)

We express \(M_{nm}\) through Eq. (47) and exchange the sum and the integral:

$$\begin{aligned} \sum ^{\infty }_{m=1} M_{nm} \alpha _m&= \frac{1}{2} \int ^{1}_{-\cos \left( \epsilon \right) } \frac{1}{1+x} \left( P_n(x)+P_{n-1}(x)\right) \nonumber \\&\quad \times \left( \sum ^{\infty }_{m=1} \bigl [ P_m(\cos t) + P_{m-1}(\cos t) \bigr ] m \alpha _m \right) \mathrm {d}x, \end{aligned}$$
(140)

from which the identity (133) leads to the announced identity (139).

Appendix 2: Spatially Averaged Variances

We denote by \(\overline{\mathbb {E}\left[ \tau ^n\right] }\) the spatial average of the \(n\)th moment of the exit time:

$$\begin{aligned} \overline{\mathbb {E}\left[ \tau ^n\right] } \equiv \dfrac{1}{\pi } \int _{\vec {r}_0 \in \Omega } \! \! \! \mathrm {d} \vec {r}_0 \ \mathbb {E}\left[ \tau ^n_{\vec {r}_0}\right] , \end{aligned}$$
(141)

where \(\mathrm {d} \vec {r}_0\) is the uniform measure over \(\Omega \). Let us now consider the random variable \(\tau _{\Omega }\), defined in Sect. 2.4.1 as the exit time of a particle started at a random starting position \(X\). The \(n\)th moment \(\tau _{\Omega }\) reads

$$\begin{aligned} \mathbb {E}\left[ \tau _{\Omega }^n\right] \equiv \int _{\vec {r}_0 \in \Omega } \! \! \! \mathrm {d}\mu (X = \vec {r}_0) \mathbb {E}\left[ \tau _{X}^n\right] , \end{aligned}$$
(142)

where \(\mathrm {d} \mu (X =\vec {r}_0)\) is the probability density for \(X\) to be started at the position \(\vec {r}_0\). If \( \mathrm {d} \mu (X =\vec {r}_0) =\mathrm {d} \vec {r}_0\) is the uniform probability distribution, we can identify Eqs. (141) and (142), \(\mathbb {E}\left[ \tau _{\Omega }^n\right] = \overline{\mathbb {E}\left[ \tau ^n\right] }\), and the variance of the random variable \(\tau _{\Omega }\) is

$$\begin{aligned} \mathrm {Var}\left[ \tau _{\Omega } \right] \equiv \left( \int _{\vec {r}_0 \in \Omega } \! \mathrm {d}\mu (X = \vec {r}_0) \mathbb {E}\left[ \tau _X^2\right] \right) - \left( \int _{\vec {r}_0 \in \Omega } \! \mathrm {d}\mu (X= \vec {r}_0) \mathbb {E}\left[ \tau _X\right] \right) ^2 = \overline{\mathbb {E}\left[ \tau ^2\right] } - \overline{\mathbb {E}\left[ \tau \right] }^2. \end{aligned}$$
(143)

Note that \(\mathrm {Var}\left[ \tau _{\Omega } \right] \) differs from the spatial average of the variance: \(\mathrm {Var}\left[ \tau _{\Omega } \right] \ne \overline{\mathrm {Var}\ \tau } = \overline{\mathbb {E}\left[ \tau ^2\right] } - \overline{\mathbb {E}\left[ \tau \right] ^2}\), because \(\overline{\mathbb {E}\left[ \tau \right] }^2 \ne \overline{\mathbb {E}\left[ \tau \right] ^2}\).

Appendix 3: Convergence to an Exponential Distribution in the Narrow-Escape Limit

1.1 From the Expression for the Survival Distribution

We recall that the expressions for \((\alpha _n), n \ge 0,\) in the limit \(\epsilon \ll 1\) are provided in Eqs. (59a)–(59b). We denote by \(\tilde{S}_{e}(t)\) the normalized single exponential distribution whose mean is equal to the GMFPT defined in Eq. (13). The Laplace transform of the distribution \(\tilde{S}^{(t)}_{e}\) is

$$\begin{aligned} S^{(p)}_{e} = \dfrac{\overline{\mathbb {E}\left[ \tau \right] }}{1 + p \ \overline{\mathbb {E}\left[ \tau \right] }}. \end{aligned}$$
(144)

We show that in the narrow-escape limit (\(\epsilon \ll 1\)), the averaged exit time distribution \(\overline{S^{(p)}} \approx a^{(p)}_0/2\) converges to \(S_{e}^{(p)}\), as expected from Ref. [12]. Due to the divergence of the coefficient \(\alpha _0\) from Eq. (59a), the asymptotic expansion of Eq. (52a) at the first order in \(\epsilon \ll 1\) reads

$$\begin{aligned} \frac{a_{0}^{(p)}}{2} = \alpha _0 \left[ \partial _r f^{[1]}_{0}\right] _{|r=1} \left[ \sum ^{\infty }_{k = 0} \left( p \left[ \partial _r f^{[1]}_{0}\right] _{|r=1} \alpha _0\right) ^k \right] + \mathcal {O}(\epsilon ). \end{aligned}$$
(145)

where \(\left[ \partial _r f^{[1]}_{0}\right] _{|r=1}\) is the first-order expansion in \(p \ll 1\) of \(\left[ \partial _r f^{(p)}_{0}\right] _{|r=1}\). At the leading order in \(\epsilon \), the averaged survival probability over \(\Omega \) is

$$\begin{aligned} \overline{S^{(p)}} = \frac{\alpha _0 \left[ \partial _r f^{[1]}_{0}\right] _{|r=1} }{1 + \alpha _0 \left[ \partial _r f^{[1]}_{0}\right] _{|r=1} } + \mathcal {O}(\epsilon ). \end{aligned}$$
(146)

Combining Eqs. (21) and (52a), the GMFPT at the leading order in \(\epsilon \ll 1\) is

$$\begin{aligned} \overline{\mathbb {E}\left[ \tau \right] } = \alpha _{0} \left[ \partial _r f^{[1]}_{0}\right] _{|r=1} + \mathcal {O}(\epsilon ). \end{aligned}$$
(147)

Combining Eqs. (144) and (146) leads to

$$\begin{aligned} \overline{S^{(p)}} = \overline{S^{(p)}_{e}} + \mathcal {O}(\epsilon ). \end{aligned}$$
(148)

This shows the convergence in law of the FPT distribution to an exponential distribution whose mean is the GMFPT as expected for the narrow-espace limit [12].

1.2 From the Expression for the Moments

Let us consider the random variable \(\tau _{\Omega }\), defined in Sect. 2.4.1 as the exit time of a particle started at a random starting position. We provide a positive answer to the following question: does the distribution of \(\tau _{\Omega }\) converge to an exponential distribution in the limit \(\epsilon \ll 1\), even though the starting positions within the boundary layer contribute to the statistics of \(\tau _{\Omega }\)? Using the recurrence scheme of Sect. 2.4.1, we verify that at the leading order in \(\epsilon \ll 1\) the moments of \(\tau _{\Omega }\) are

$$\begin{aligned} \mathbb {E}\left[ \tau _{\Omega }^n\right] = n! \left( \frac{\alpha _0}{2}\right) ^n + \mathcal {O}(\ln (\epsilon )^{n-1}), \end{aligned}$$
(149)

which leads to \(\mathbb {E}\left[ \tau _{\Omega }^n\right] = n! \ \mathbb {E}\left[ \tau _{\Omega }\right] ~ (n \ge 1)\) at the leading order in \(\epsilon \). The latter identity indicates that the FPT distribution of \(\tau _{\Omega }\) converges to an exponential distribution whose mean is the GMFPT defined by Eq. (29), as expected from Sect. 7.1. We emphasize that the relation \(\mathbb {E}\left[ \tau _{\Omega }^n\right] = n! \ \mathbb {E}\left[ \tau _{\Omega }\right] ~ (n\ge 1)\), implies that the GMFPT characterizes the whole distribution of the exit time, in contrast to the statement of Eq. (18) from Ref. [18].

Appendix 4: Computational Aspects

We summarize the numerical methods used to compute the FPT distribution.

  1. (i)

    The Monte Carlo simulations rely on a sample of \(2 \cdot 10^{6}\) of random walks. This sample is obtained through \(2\) hours of computation on \(200\) CPUs (\(3.20\) GHz Intel Core\(^\mathrm{TM}\) i7). The home-built C++ program uses an adaptive time step method so that the time steps are given by a decreasing function with the distance to the exit.

  2. (ii)

    A finite element method realized in COMSOL Multiphysics v4.2 [33] allowed to greatly reduce the computational time. For instance, the FPT probability density shown in Fig.2a, b required 10–15 min on a single CPU (\(2,66\) GHz Intel Core \(^\mathrm{TM}\) i5).

  3. (iii)

    The exact and approximate analytical solutions were computed in MATLAB and using the numerical Laplace inversion package INVLAP [38]. The series were truncated at \(N = 100\) terms and the computational time is of the order of a few minutes on a single CPU (\(2,66\) GHz Intel Core \(^\mathrm{TM}\) i5).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rupprecht, JF., Bénichou, O., Grebenkov, D.S. et al. Exit Time Distribution in Spherically Symmetric Two-Dimensional Domains. J Stat Phys 158, 192–230 (2015). https://doi.org/10.1007/s10955-014-1116-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-014-1116-6

Keywords

Navigation