Skip to main content
The Journal of Chemical Physics logoLink to The Journal of Chemical Physics
. 2012 Aug 1;137(5):054104. doi: 10.1063/1.4737662

Exact Green's function of the reversible diffusion-influenced reaction for an isolated pair in two dimensions

Thorsten Prüstel 1,a), Martin Meier-Schellersheim 1,b)
PMCID: PMC3422329  PMID: 22894329

Abstract

We derive an exact Green's function of the diffusion equation for a pair of disk-shaped interacting particles in two dimensions subject to a backreaction boundary condition. Furthermore, we use the obtained function to calculate exact expressions for the survival probability and the time-dependent rate coefficient for the initially unbound pair and the survival probability of the bound state. The derived expressions will be of particular utility for the description of reversible membrane-bound reactions in cell biology.


Diffusion, or Brownian motion, and its microscopic basis, random walks, are key concepts of non-equilibrium statistical mechanics and the theory of stochastic processes. Their ubiquity renders them applicable in numerous areas of physics and beyond, for instance, engineering, chemistry, biology, and mathematical finance, see, for example, Refs. 1 and 2 and references therein.

In the theory of diffusion-influenced reactions,3 solutions of the diffusion equation which satisfy certain boundary conditions can be used to investigate different types of chemical reactions. Among those solutions, Green's functions (GF) enjoy a privileged role because they permit calculating the solution for any given initial distribution and can be used to derive important other quantities, for instance survival probabilities and time-dependent reaction rate coefficients.4, 5, 6 In this sense, knowledge of the GF is tantamount to a “completely solved” problem. However, in most cases, an analytical representation of the GF remains elusive, a notable exception being the case of an isolated pair. Here, GFs and derived quantities have not only been important from a conceptual point of view but have also been used to fit experimental data for a diffusion model with reversible reactions7 and to investigate how the reduced parameter set could be determined from experimental data describing geminate recombination.8

Motivated by the desire to understand the influence of stochastic fluctuations and spatial heterogeneities on the behavior of biochemical networks, the last decade has witnessed an increased interest in theoretical approaches describing diffusion-influenced reactions at the molecular level. Analytical representations of GFs describing an isolated pair figure prominently in a number of proposed particle-based stochastic simulation algorithms, because a reaction network may be thought of as composed of unimolecular and bimolecular reactions A + Bproducts. In this context, GFs can be used to enhance the efficiency of Brownian dynamics simulations.9, 10, 11 Moreover, the knowledge of exact analytical expressions permits to validate newly devised stochastic simulation algorithms.12

Exact analytic expressions for the GF of an isolated pair that can undergo a reversible reaction have been derived for the one and three dimensional cases,4, 6 whereas a corresponding expression in the time domain for the two dimensional case has been lacking and analytical approaches were limited to approximations13, 14 in spite of its broad applicability to surface phenomena. The two dimensional case is of particular importance in cell biological applications, providing the basis for a better understanding of processes such as signal-induced inhomogeneities and receptor clustering on cell membranes15 where small diffusion constants govern the lateral motion of receptors and diffusion represents a less efficient mixing mechanism than in the three dimensional cytoplasm. Finally, diffusion in two dimensions (2D) is special because the steady-state solution of the diffusion equation is inconsistent with the boundary condition at infinity16 and because 2D is the critical dimension with regard to the recurrence and transience of random walks.17

To derive the GF in 2D, we consider an isolated pair of two disklike particles A and B with diffusion constants DA and DB, respectively. The particles may associate when their separation equals the encounter distance a to form a bound molecule AB. When bound, the molecules may dissociate again to form an unbound pair A + B. Such a system may be described as the diffusion of a point-like particle with diffusion constant D = DA + DB around a static disk with radius a. In this picture, reactions are introduced by imposing boundary conditions at the disk's “surface.” The irreversible association reaction is described by the radiation boundary condition (BC) that is characterized by an intrinsic association constant κa. To take into account reversible reactions, i.e., to allow for dissociations, the radiation BC has to be generalized to the backreaction BC that involves an additional intrinsic dissociation constant κd.4, 5, 6 We consider the probability density function (PDF) g(r, t|r0), which gives the probability to find the particles at a distance equal to r at time t, given that the distance was initially r0 at time t = 0. The time evolution of g(r, t|r0) is governed by the 2D diffusion equation5

tg(r,t|r0)=D2r2+1rrg(r,t|r0),ra. (1)

The diffusion equation has to be completed by specifying boundary conditions. Together with the following initial:

2πr0g(r,0|r0)=δ(rr0), (2)

and boundary condition

limrg(r,t|r0)=0, (3)

Eq. 1 is equivalent to the free-space diffusion equation in 2D (if a = 0, r > 0) with the familiar solution18 (Ch. 14.8, Eq. 1)

gfree(r,t|r0)=14πDte(r2+r02)/4DtI0rr02Dt, (4)

which is also known as the free-space GF. I0(x) denotes the modified Bessel function of first kind and zero order19 (Sec. 9.6).

The PDF we are interested in is only defined for ra > 0 and one has to impose a BC for r = a specifying the behavior at the encounter distance. To incorporate association and dissociation on the surface of the “interaction disc,” we thus require4, 5, 6

2πaDrg(r,t|r0)|r=a=κag(a,t|r0)κd[1S(t|r0)]. (5)

Here, S(t|r0) denotes the survival probability that a pair of molecules with initial distance r0 survives until time t

S(t|r0)=2πag(r,t|r0)rdr, (6)
=12πaD0trg(r,t|r0)|r=adt. (7)

Note that, strictly speaking, the notion of a “survival probability” is somewhat misleading in the case we consider here. For the irreversible reaction, a non-vanishing likelihood to find a separated isolated pair at a time t necessarily implies that no association has occurred before. For the reversible reaction, association and subsequent dissociation events that took place possibly many times before t also contribute to the probability of finding an isolated pair unbound at a time t. Therefore, more precisely, one could refer to the quantity S(t|r0) as a separation probability.5 Nevertheless, in the following we will use survival probability interchangeably with separation probability.6 Furthermore, we would like to point out that the possibility of an initial bound state (henceforth denoted by *) requires to consider two different survival probabilities: S(t|r0) and S(t|*), corresponding to the initially unbound and bound state, respectively. Similarly, the reversible GF has four components: besides g(r, t|r0), one has to consider g(r, t|*), g(*, t|r0), and g(*, t|*) for the initial and final bound states.5, 6 For all these quantities, we will derive exact analytical expressions in the time domain.

Following the general strategy of Ref. 18, [Ch. 14.8, p. 368], we make the following ansatz for the Laplace transform of the Green's function that satisfies the backreaction BC

g˜(r,q|r0)=g˜free(r,q|r0)+g˜br(r,q|r0). (8)

Here,

g˜free(r,q|r0)=12πDI0(qr0)K0(qr),r>r0I0(qr)K0(qr0),r<r0, (9)

is the Laplace transform of the free-space GF Eq. 4, cp.18 (Ch. 14.8, Eq. 2). K0(x) refers to the modified Bessel function of second kind and zero order19 (Sec. 9.6). The variable q is defined by q:=p/D, where p denotes the Laplace domain variable.

Alternatively, one could also seek a solution in the form4, 6

g˜(r,q|r0)=g˜ref(r,q|r0)+g˜ref,br(r,q|r0), (10)

where g˜ref(r,q|r0) denotes the Laplace transform of the GF that obeys reflecting BCs.

The part g˜br in Eq. 8 that takes into account the BC is a solution of the Laplace transformed 2D diffusion equation18 (Ch. 14.8, Eq. 3)

d2g˜brdr2+1rdg˜brdrq2g˜br=0. (11)

The general solution to Eq. 11 is AK0(qr) + BI0(qr). Because we require limrg˜br0, and limx → ∞I0(x) → ∞, the coefficient B has to vanish and hence,

g˜br(r,q|r0)=A(q,r0)K0(qr). (12)

A(q, r0) is determined by the requirement that the complete GF Eq. 8 satisfies the Laplace transformed backreaction BC, cp. Eq. 5,

rg˜(r,q|r0)|r=a=hg˜(a,q|r0)κdp1rg˜(r,q|r0)|r=a, (13)

where we have defined h ≔ κa/(2πaD). Using Eqs. 8, 9, 12, 13, and I0(x)=I1(x),K0(x)=K1(x)19 (Eq. (9.6.27)) and defining κD ≔ κd/D, we obtain

g˜br(r,q|r0)=12πDK0(qr)K0(qr0)×(q2+κD)I1(qa)hqI0(qa)(q2+κD)K1(qa)+hqK0(qa). (14)

We would like to point out that for the irreversible case, the GF that takes into account the reaction by satisfying the radiation BC can be rewritten in terms of the GF that corresponds to nonreactive diffusion and satisfies a reflective BC (Ref. 20)

g˜rad(r,q|r0)=g˜ref(r,q|r0)κag˜ref(r,q|a)g˜ref(a,q|r0)1+κag˜ref(a,q|a). (15)

It was also shown that Eq. 15 relating the GFs for reactive and nonreactive diffusion takes the form of a Dyson integral equation in the time domain.20 Furthermore, in Ref. 21, Eq. 15 was extended to the reversible case in 3D. This relationship is valid for the reversible case in 2D also. To obtain the GF satisfying reflective BCs, we set h = κD = 0 in Eq. 14 and find

g˜ref,bc(r,q|r0):=g˜ref(r,q|r0)g˜free(r,q|r0),=12πDK0(qr)K0(qr0)I1(qa)K1(qa). (16)

Using I0(x)K1(x) + I1(x)K0(x) = x−1, cp.19 (Eq. (9.6.15)) it follows that

g˜br(r,q|r0)=g˜ref,bc(r,q|r0)κaq2g˜ref(r,q|a)g˜ref(a,q|r0)q2+κD+κaq2g˜ref(a,q|a). (17)

The inversion theorem for the Laplace transformation can be applied to find the corresponding expression of g˜br in the time domain

gbr(r,t|r0)=12πiγiγ+ieptg˜br(r,q|r0)dp. (18)

To calculate the Bromwich contour integral, we first note that g˜br has a branch point at p = 0. Therefore, we use the contour of Fig. 1 with a branch cut along the negative real axis, cp.18 (Ch. 12.3, Fig. 40). Furthermore, we note that the integrand has no poles within and on the contour22 and that the contribution from the small circle around the origin vanishes, which can be seen by the limiting forms of the modified Bessel functions for small arguments19 (Eqs. (9.6.7)–(9.6.9)). Therefore, we obtain

0=eptg˜br(r,q|r0)dp=γiγ+ieptg˜br(r,q|r0)dp+C2eptg˜br(r,q|r0)dp+C4eptg˜br(r,q|r0)dp. (19)

Thus, it remains to calculate the integrals C2,C4. To this end, we choose p = Dx2eiπ and use18 (Append. 3, Eqs. 25, 26)

In(xe±πi/2)=e±nπi/2Jn(x), (20)
Kn(xe±πi/2)=±12πienπi/2[Jn(x)±iYn(x)]. (21)

Jn(x), Yn(x) denote the Bessel functions of first and second kind, respectively19 (Sec. 9.1). It follows that

C2eptg˜br(r,q|r0)dp=i20eDx2tH0(2)(xr)H0(2)(xr0)×α(x)[α(x)+iβ(x)]α(x)2+β(x)2xdx. (22)

Here, Hn(2)(x):=Jn(x)iYn(x) denotes the Bessel function of third kind (also referred to as Hankel function)19 (Eq. 9.1.4) and we have defined

α:=α(x):=(x2κD)J1(xa)+hxJ0(xa), (23)
β:=β(x):=(x2κD)Y1(xa)+hxY0(xa). (24)

To evaluate the integral along the contour C4, we choose p = Dx2eiπ and after an analogous calculation, one finds that C2eptg˜br(r,q|r0)dp=C4eptg˜br(r,q|r0)dp*, where * means complex conjugation. Thus, one arrives at

gbr(r,t|r0)=1πC2eptg˜br(r,q|r0)dp,=12π0eDx2tααΩ+βΠα2+β2xdx, (25)

where we have defined

Ω:=Ω(x,r,r0):=J0(xr)J0(xr0)Y0(xr)Y0(xr0),Π:=Π(x,r,r0):=Y0(xr)J0(xr0)+J0(xr)Y0(xr0).

Next, we use the fact that the free-space Green's function may be written as

gfree(r,t|r0)=12π0eDx2tJ0(xr)J0(xr0)xdx

to arrive at the exact Green's function in the time domain

g(r,t|r0)=12π0eDx2tT(x,r)T(x,r0)xdx, (26)

where we introduced the function23

T(x,r):=J0(rx)β(x)Y0(rx)α(x)[α(x)2+β(x)2]1/2. (27)

Note that in the limit κd → 0, one recovers the known GF for the irreversible case with radiation BC18 (Chap. 14.8, Eqs. 12, 13). We compared the obtained analytical expression for the GF to solutions that were numerically constructed using the Spherically Symmetric Diffusion Problem (SSDP) software,24 version 2.66. Results are shown in the left panel of Fig. 2. We find excellent agreement.

Figure 1.

Figure 1

Integration contour used in Eq. 19.

Figure 2.

Figure 2

The reversible Green's function and survival probability for an initially unbound isolated pair in 2D. The parameters are: a = 1, D = 2, r0 = 1.1, κa = 2πD. The left panel shows the r dependence of the reversible GF 2πrg(r, t|r0) at time t = 10η for various values of the dissociation constant κd = 1 η−1, 0.5 η−1, 0.1 η−1, 0.01 η−1, where η = a2D−1 denotes the natural time scale of the system. The solid lines correspond to the analytical expression Eq. 26. The various markers refer to numerical solutions that were obtained using the SSDP software, ver. 2.66. The right panel shows the time dependence of the survival probability S(t|r0). The dissociation constant is κd = 0.5 η−1. The solid line refers to the analytical expression in Eq. 29. The markers indicate the SSDP results.

With the exact GF at our disposal, further important quantities can be derived.25 Notably, the survival probability can be calculated using Eq. 7. To this end, we first introduce the function

P(x,r):=1xrT(x,r)=J1(xr)βY1(xr)α[α2+β2]1/2, (28)

where we used J0(x)=J1(x) and Y0(x)=Y1(x)19 (Eq. 9.1.28). Then the survival probability can be written as

S(t|r0)=1a0eDtx2P(x,a)T(x,r0)dx. (29)

To obtain this result, we note that, as discussed in Appendix A, the integral 0P(x,a)T(x,r0)dx has to vanish. Clearly, the reversible survival probability approaches unity for large times, limt → ∞S(t|r0) = 1. Therefore, it can explicitly be shown that the fate of an isolated pair for the reversible reaction is always dissociation in 2D, as in the 1D and 3D cases.4, 6

Let g(*, t|r0) and g(r, t|*), g(*, t|*) denote the Green's functions for the final and initial bound states, respectively.5, 6 Note that g(*, t|r0) and g(*, t|*) are probabilities instead of probability densities and that g(*, t|r0) = 1 − S(t|r0). Because of the detailed balance condition5, 6

κdg(*,t|r)=κag(r,t|*), (30)

the GF of the initial bound state becomes

g(r,t|*)=κdκaa0eDtx2P(x,a)T(x,r)dx. (31)

Thus, the probability that an initially bound pair is unbound at time t > 0 becomes

S(t|*)=12πκdκaa20eDtx2P2(x,a)1xdx. (32)

Note that S(0|*) = 0, due to Eq. 38. In Ref. 5, it was demonstrated that S(t|*) relates to S(t|r0) in the time domain as

S(t|*)=κd0teκdtS(tt|a)dt. (33)

Using the obtained expression for the GF, Eqs. 26, 27, and the survival probabilities, Eq. 29, 32, as well as

Y0(x)J0(x)J0(x)Y0(x)=2/(πx), (34)

Ref. 19 (Eq. (9.1.16)) and Eq. 38, we can explicitly verify that Eq. 33 is satisfied in 2D. Finally, we obtain for g(*, t|*) = 1 − S(t|*),

g(*,t|*)=2πκdκaa20eDtx2P2(x,a)1xdx, (35)

and g(*, 0|*) = 1. Using the SSDP software, we compared the numerically calculated survival probability S(t|r0) of the initially unbound as well as the GF and survival probability S(t|*) of the initially bound state to our analytical expressions Eq. 29, Eq. 31, and Eq. 32. The right panel of Fig. 2 and both panels of Fig. 3 show the results. Again, we find excellent agreement.

Figure 3.

Figure 3

The reversible Green's function and survival probability for an initially bound isolated pair in 2D. The parameters are: a = 1, D = 2, κa = 2πD. The left panel shows the r dependence of the reversible GF 2πrg(r, t|*) at time t = 10 η for various values of the dissociation constant κd = 1 η−1, 0.5 η−1, 0.1 η−1, 0.025 η−1, where η = a2D−1 denotes the natural time scale of the system. The solid lines are obtained from the analytical expression Eq. 31. The various markers indicate the corresponding numerical solutions generated by the SSDP software. The right panel shows the time dependence of the survival probability S(t|*). The dissociation constant is κd = 0.5 η−1. The solid line refers to the analytical expression Eq. 32. The markers indicate the SSDP results.

Finally, the time-dependent reaction rate coefficient can be defined as5, 6

k(t):=2πaDr0S(t|r0)|r0=a. (36)

The resulting exact expression is

k(t)=2πa2D0eDtx2P2(x,a)xdx. (37)

As explained in Appendix A, one can show that

0P2(x,a)1xdx=κaκd2πa2. (38)

We thus recover the correct expression for the equilibrium constant Keq,

0k(t)dt=κaκd=:Keq. (39)

Alternatively, Eq. 39 may be deduced in the following way. According to Ref. 5, the time-dependent reaction rate can also be calculated by using the survival probability S(t|*),

k(t)=KeqS(t|*)t. (40)

Now, given Eqs. 29, 32, one can explicitly show that Eqs. 36, 40 yield the same result. Then, taking into account that S(∞|*) = 1 and S(0|*) = 0, Eq. 39 is immediately implied by Eq. 40.

ACKNOWLEDGMENTS

This research was supported by the Intramural Research Program of the NIH, National Institute of Allergy and Infectious Diseases. We would like to thank Bastian R. Angermann and Frederick Klauschen for helpful discussions.

APPENDIX A: INTEGRAL IDENTITIES

In this appendix, we will show how the derived GF Eq. 26 and the backreaction BC Eq. 5 can be used to calculate seemingly intractable integrals that are needed in the main text. Substituting the found expression of the GF into the backreaction BC and taking the limit t → ∞ or, alternatively, using T(x, a) = −(x2 − κD)/(hx)P(x, a), yields

0P(x,a)T(x,r0)dx=0. (A1)

By definition, the Green' s function has to satisfy the initial condition Eq. 2, i.e., in the case considered here we have

12π0T(x,r)T(x,r0)xdx=δ(rr0)2πr0. (A2)

It follows by direct integration over ardrr

0P(x,r)T(x,r0)dx={0,r<r0,r1,r>r0, (A3)

where we have used Eq. A1. Note also that to perform the integral, we used d/dx[xνJν(x)] = xνJν−1 and d/dx[xνYν(x)] = xνYν − 119 (Eq. (9.1.30)). Next, we apply the derivative ∂/∂r0 to the BC Eq. 5 and subsequently integrate from 0 to t. Finally, we again consider the limit t → ∞ and find

0P(x,a)P(x,r0)xdx=κa2πaD0T(x,a)P(x,r0)dx+κdD0P(x,a)P(x,r0)dxx. (A4)

The integral on the lhs has to vanish, as can be seen by taking into account Eq. A1 and

0P(x,a)P(x,r0)xdx=r00P(x,a)T(x,r0)dx. (A5)

Furthermore, we just calculated the first integral on the rhs, cp. Eq. A3. Hence, we arrive at

0P(x,a)P(x,r0)dxx=κaκd2πar0. (A6)

APPENDIX B: ASYMPTOTIC AND APPROXIMATE EXPRESSIONS

In the following, we present some approximate solutions for small and long times. Note that approximate solutions have been discussed before5, 13, 14 and at least for the survival probabilities, we re-derive already known results.

Generally, two alternative ways exist to obtain approximate solutions for long times. One method starts in the Laplace domain and is described in Ref. 18 (Ch. 13.6). One starts from the expressions Eq. 9 and Eq. 14 in the Laplace domain and expands them in the ascending powers of q. The obtained series can be integrated term by term using the integrals given in Ref. 18 (Ch. 13.6, Eqs. 4, 5, 6, 7, 8, 9). Alternatively, one might use the exact expressions in the time domain to generate approximations. To demonstrate the general idea, we start with the survival probability S(t|r0). First, we note that

numeratorofP(x,a)=2hπa, (B1)

because of Eqs. 28, 34. We make a variable substitution Dtxu to obtain

0eDtx2P(x,a)T(x,r0)dx1Dt0eu2P(uDt1,a)T(uDt1,r0)du. (B2)

Because we are interested in long times, we use the expansions18 (Append. 3, Eqs. 1, 5) of the Bessel functions suitable for small arguments to expand the integrand in ascending powers of ξ:=uDt1. To leading order, we find

1α(ξ)2+β(ξ)2=ξ2[2(πa)1κD]2[1+O(ξ2lnξ)]2. (B3)

Similarly, we arrive at

numeratorofξT(ξ,r0)=2κDπa(1+O(ξ2lnξ)). (B4)

Combining Eqs. 29, B1, B2, B3, B4, we arrive at

S(t|r0)t1Keq4πDt+ (B5)

In the same way, we can find an approximate expression for S(t|*). Again performing the variable substitution Dtxu and using Eqs. 32, B1, B3, we obtain

S(t|*)t1Keq4πDt+, (B6)

so we find that the survival probabilities S(t|r0) and S(t|*) show the same long time behavior. This is what one expects, because the long time characteristics should be independent of the initial state. Finally, to derive a long time expansion for the GF Eq. 26, we first expand the free GF Eq. 4 as follows:

gfree(r,t|r0)=14πDt1r2+r024Dt+O(t2), (B7)

where we have used the expansion of I0(x) suitable for small arguments18 (Append. 3, Eqs. 7). We obtain for Eq. 25

gbr(r,t|r0)=a24πD2t212Keq2πa2lnCrr0Dt+O(t3), (B8)

where C ≔ 1/4eγ + 1 and γ = .5772156649… denotes Euler's constant19 (Eq. (6.1.3)). To derive Eq. B8, we used

0emxln(x)xdx=1m2(1γln(m)), (B9)

which follows from Ref. 26 (Eq. (4.331.1))

0emxln(x)dx=γ+ln(m)m, (B10)

by differentiation with respect to the parameter m.

To obtain a small time expansion, it is convenient to start from the Laplace domain and to use the expressions for the free GF and the boundary component, given by Eq. 9 and Eq. 14, respectively. Furthermore, we switch to dimensionless variables R=r/a,R0=r0/a,h˜=ha,κ˜D=κDa2,q˜=qa, and τ = tDa−2. To obtain expansions in powers of q˜1, we exploit the asymptotic expansions of the modified Bessel functions for large arguments19 (Eqs. (9.7.1), (9.7.2)),

Iν(x)ex2πxk=0(1)kak(ν)xk, (B11)
Kν(x)π2xexk=0ak(ν)xk, (B12)

where the coefficients are defined by

ak(ν)=(4ν212)(4ν232)(4ν2(2k1)2)k!8k (B13)

to arrive at the following asymptotic expansion in powers of q−1:

g˜free(R,q˜|R0)eq˜(RR0)4πDRR0q˜k=0ψk(R,R0)q˜k, (B14)
g˜br(R,q˜|R0)eq˜(R+R02)4πDRR0q˜k=0ϕk(R,R0)q˜k. (B15)

Note that Eq. B14 assumes R > R0. For convenience, we give the first expansion coefficients (ψ0 = ϕ0 = 1),

ψ1=18(R1R01), (B16)
ϕ1=18(R1+R01)342h˜. (B17)

The explicit form of the coefficients quickly becomes cumbersome, so that for instance in a simulation that uses small time expansions of the GF, all the required coefficients should be calculated by iterative use of B13. To proceed, we use the (inverse) Laplace transforms18 (Append. 5, Eqs. 7, 11)

L1eqxq=Dπt1/2ex24Dt, (B18)
L1eqxp1+n/2=(4t)n/2inerfcx2Dt. (B19)

The functions inerfc(x) are defined by18 (Append. 2, Eqs. 9, 10, 11)

inerfc(x):=xin-1erfc(ξ)dξ, (B20)

and i0erfc(x) ≔ erfc(x). None of the integrals in Eq. B20 has to be calculated, because the inerfc(x) functions satisfy the recursion relation18 (Append. 2, Eq. 14)

2ninerfc(x)=in-2erfc(x)2xin-1erfc(x), (B21)

which allows to calculate Eq. B20 swiftly.

We finally arrive at the expressions for small times in the time domain

g(R,τ|R0)14πa2τRR0{π1/2e(RR0)24τ+e(R+R02)24τ+12n=1[ψnin-1erfcRR02τ+ϕnin-1erfcR+R022τ](2τ)n}. (B22)

Note that again R > R0 is assumed.

Small-time expansions for the other quantities considered here can be derived in a similar way. The survival probability becomes

S(t|r0)1n=2φnin-1erfcR012τ(2τ)n4R0τ. (B23)

The first two coefficients are φ2=2h˜ and φ3=h˜(3/4+2h˜+1/(4R0)).

Using the Laplace transform18 (Append. 5, Eq. 2) L1(pν+1)=tνΓ(ν+1)1, we finally obtain

S(t|*)κ˜Dτn=0ωnτn/2, (B24)

where ω0=1,ω1=4/(3π)h˜.

Finally, we would like to comment on how to numerically evaluate the analytical expressions for the GF Eq. 26 and the derived quantities. The time dependence of all the quantities is generically given by 0eDtx2...dx. The quadratic exponential strongly dampens the value of the whole integrand and limits the infinite integration range to a finite range in the sense that contributions to the integral coming from x values larger than a cut-off can be neglected. For the actual integration, we used the Quadrature Adaptive General integrand (QAG) integration algorithm of the GNU Scientific Library (GSL),27 which employs the Gauss-Kronrod scheme28 (Ch. 2.7.1.1). If the accuracy is still not satisfactory, one might divide the integration range in three intervals corresponding to small, intermediate, and large x values. Then the integrand can be approximated corresponding to the x range using the expansions of the Bessel functions that are suitable for small and large arguments,18 (Append. 3, Eqs. 1, 5) and19 (Eqs. (9.2.5), (9.2.6)). Alternatively, for small and large times one may directly use the expressions Eqs. B22, B7, B8, respectively.

References

  1. Hänggi P. and Marchesoni F., Chaos 15, 26101 (2005). 10.1063/1.1895505 [DOI] [PubMed] [Google Scholar]
  2. Frey E. and Kroy K., Ann. Phys. 14, 20 (2005). 10.1002/andp.200410132 [DOI] [Google Scholar]
  3. Rice S. A., Diffusion Limited Reactions (Elsevier, New York, 1985). [Google Scholar]
  4. Agmon N., J. Chem. Phys. 81, 2811 (1984). 10.1063/1.447954 [DOI] [Google Scholar]
  5. Agmon N. and Szabo A., J. Chem. Phys. 92, 5270 (1990). 10.1063/1.458533 [DOI] [Google Scholar]
  6. Kim H. and Shin K., Phys. Rev. Lett. 82, 1578 (1999). 10.1103/PhysRevLett.82.1578 [DOI] [Google Scholar]
  7. Agmon N., J. Phys. Chem. A 109, 13 (2005). 10.1021/jp047465m [DOI] [PubMed] [Google Scholar]
  8. Agmon N., Chem. Phys. Lett. 417, 530 (2006). 10.1016/j.cplett.2005.10.058 [DOI] [Google Scholar]
  9. Edelstein A. and Agmon N., J. Chem. Phys. 99, 5396 (1993). 10.1063/1.465983 [DOI] [Google Scholar]
  10. Kim H., Yang M., and Shin K., J. Chem. Phys. 111, 1068 (1999). 10.1063/1.479297 [DOI] [Google Scholar]
  11. van Zon J. and ten Wolde P., Phys. Rev. Lett. 94, 128103 (2005). 10.1103/PhysRevLett.94.128103 [DOI] [PubMed] [Google Scholar]
  12. van Zon J. and ten Wolde P., J. Chem. Phys. 123, 234910 (2005). 10.1063/1.2137716 [DOI] [PubMed] [Google Scholar]
  13. Park K., Shin K., and Kim H., J. Chem. Phys. 131, 154105 (2009). 10.1063/1.3242273 [DOI] [PubMed] [Google Scholar]
  14. Kim H., Chem. Phys. Lett. 507, 265 (2011). 10.1016/j.cplett.2011.03.079 [DOI] [Google Scholar]
  15. Bethani I., Skanland S., Dikic I., and Acker-Palmer A., EMBO J. 29, 2677 (2010). 10.1038/emboj.2010.175 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Emeis C. and Fehder P., J. Am. Chem. Soc. 92, 2246 (1970). 10.1021/ja00711a009 [DOI] [Google Scholar]
  17. Toussaint D. and Wilczek F., J. Chem. Phys. 78, 2642 (1983). 10.1063/1.445022 [DOI] [Google Scholar]
  18. Carslaw H. and Jaeger J., Conduction of Heat in Solids (Clarendon, New York, 1986). [Google Scholar]
  19. Abramowitz M. and Stegun I., Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables (Dover, New York, 1965). [Google Scholar]
  20. Szabo A., Lamm G., and Weiss G., J. Stat. Phys. 34, 225 (1984). 10.1007/BF01770356 [DOI] [Google Scholar]
  21. Gopich I., Solntsev K., and Agmon N., J. Chem. Phys. 110, 2164 (1999). 10.1063/1.477827 [DOI] [Google Scholar]
  22. Erdelyi A. and Kermack W., Proc. Camb. Phil. Soc. 41, 74 (1945). 10.1017/S0305004100022404 [DOI] [Google Scholar]
  23. Prüstel T. and Meier-Schellersheim M., e-print arXiv:1109.4465v1 [q-bio.QM] (2011).
  24. Krissinel' E. and Agmon N., J. Comput. Chem. 17, 1085 (1996). [DOI] [Google Scholar]
  25. Prüstel T. and Meier-Schellersheim M., e-print arXiv:1112.4010v1 [math-ph] (2011).
  26. Gradshteyn I., and Ryzhik I., Table of Integrals, Series, and Products, edited by Jeffrey A., and Zwillinger D. (Academic, New York, 2007). [Google Scholar]
  27. Galassi M., Davies J., Theiler J., Gough B., Jungman G., Alken P., Booth M., and Rossi F., GNU Scientific Library Reference Manual, 3rd ed. (Network Theory, 2009). [Google Scholar]
  28. Davis P. and Rabinowitz P., Methods of Numerical Integration (Dover, Mineola, 2007). [Google Scholar]

Articles from The Journal of Chemical Physics are provided here courtesy of American Institute of Physics

RESOURCES