06.01.2015 Views

Fractional reaction-diffusion equation for species ... - ResearchGate

Fractional reaction-diffusion equation for species ... - ResearchGate

Fractional reaction-diffusion equation for species ... - ResearchGate

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Noname manuscript No.<br />

(will be inserted by the editor)<br />

Boris Baeumer · Mihály Kovács · Mark<br />

M. Meerschaert<br />

<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong><br />

<strong>for</strong> <strong>species</strong> growth and dispersal<br />

Received: date / Revised: date<br />

Abstract Reaction-<strong>diffusion</strong> <strong>equation</strong>s are commonly used to model the<br />

growth and spreading of biological <strong>species</strong>. The classical <strong>diffusion</strong> term implies<br />

a Gaussian dispersal kernel in the corresponding integro-difference <strong>equation</strong>,<br />

which is often unrealistic in practice. In this paper, we propose a fractional<br />

<strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> where the classical second derivative <strong>diffusion</strong><br />

term is replaced by a fractional derivative of order less than two. The resulting<br />

model captures the faster spreading rates and power law invasion profiles<br />

observed in many applications, and is strongly motivated by a generalised<br />

central limit theorem <strong>for</strong> random movements with power-law probability tails.<br />

We then develop practical numerical methods to solve the fractional <strong>reaction</strong><strong>diffusion</strong><br />

<strong>equation</strong> by time discretization and operator splitting, along with<br />

Partially supported by NSF grants DMS-0139927 and DMS-0417869 and the Marsden<br />

Fund administered by the Royal Society of New Zealand.<br />

Boris Baeumer<br />

Department of Mathematics and Statistics, University of Otago,<br />

P.O. Box 56, Dunedin 9001, New Zealand<br />

Tel.: +643-479-7763<br />

Fax: +643-479-8427<br />

E-mail: bbaeumer@maths.otago.ac.nz<br />

Mihály Kovács<br />

Department of Mathematics and Statistics, University of Otago,<br />

P.O. Box 56, Dunedin 9001, New Zealand<br />

Tel.: +643-479-7889<br />

Fax: +643-479-8427<br />

E-mail: mkovacs@maths.otago.ac.nz<br />

Mark M. Meerschaert<br />

Department of Mathematics and Statistics, University of Otago,<br />

P.O. Box 56, Dunedin 9001, New Zealand<br />

Tel.: +643-479-7889<br />

Fax: +643-479-8427<br />

E-mail: mcubed@maths.otago.ac.nz


2 Baeumer, Kovács, Meerschaert<br />

some existing methods from the literature on anomalous super-<strong>diffusion</strong>. In<br />

the process, we establish the mathematical relationship between the discrete<br />

time integro-difference and continuous time <strong>reaction</strong>-<strong>diffusion</strong> analogues of<br />

the model, along with error bounds. Our general approach also applies to<br />

other alternative non-Gaussian dispersal kernels, and it identifies the analogous<br />

continuous time evolution <strong>equation</strong>s <strong>for</strong> those models.<br />

Keywords <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> · growth and dispersal · fractional<br />

derivative · anomalous <strong>diffusion</strong> · operator splitting<br />

Mathematics Subject Classification (2000) 35K57 · 26A33 · 47D03<br />

1 Introduction<br />

Classical <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong>s are useful to model the spread of invasive<br />

<strong>species</strong> [40,41]. In this model, the population density u(x, t) at location<br />

x and time t is the solution of the partial differential <strong>equation</strong><br />

∂u<br />

∂t = f(u) + D ∂2 u<br />

∂x 2 . (1)<br />

The first term on the right is the <strong>reaction</strong> term that models population<br />

growth; a typical choice is Fisher’s <strong>equation</strong> f(u) = ru(1 − u/K) where r is<br />

the intrinsic growth rate and K is the carrying capacity. The second term<br />

is the <strong>diffusion</strong> term; it models spreading/dispersion. Solutions to (1) spread<br />

at a rate proportional to t with exponential leading edges. The main failure<br />

of this model in real applications is the unrealistically slow spreading, since<br />

typical invasive <strong>species</strong> have population densities that spread faster than t,<br />

with power law leading edges [13,16,17,26,28,43]. The power law exhibits as<br />

a straight line on a log-log plot of dispersive distance, a phenomenon often<br />

seen in field data [58–60]. In this paper, we propose an alternative fractional<br />

<strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong><br />

∂u<br />

∂t = f(u) + D ∂α u<br />

∂x α (2)<br />

with 0 < α ≤ 2, which reduces to the classical <strong>equation</strong> if α = 2. Solutions to<br />

(2) spread faster than t with power law leading edges [19] and hence provide<br />

a more realistic model <strong>for</strong> invasive <strong>species</strong>.<br />

<strong>Fractional</strong> derivatives are almost as old as their more familiar integerorder<br />

counterparts [37,48]. <strong>Fractional</strong> derivatives have recently been applied<br />

to many problems in physics [5,10,11,27,32–35,47,61], finance [22,45,46,51–<br />

53], and hydrology [4,6–8,55,56]. If a function u(x) has Fourier trans<strong>for</strong>m<br />

û(λ) = ∫ e −iλx u(x)dx then the fractional derivative d α u(x)/dx α has Fourier<br />

trans<strong>for</strong>m (iλ) α û(λ), extending the familiar <strong>for</strong>mula <strong>for</strong> integer α. <strong>Fractional</strong><br />

derivatives are used to model anomalous <strong>diffusion</strong> or dispersion, where a<br />

particle plume spreads at a different rate than the classical <strong>diffusion</strong> model<br />

predicts [15,34]. Just as the fundamental solution to the classical <strong>diffusion</strong><br />

<strong>equation</strong> is provided by the normal or Gaussian probability density functions<br />

of a Brownian motion, the α-stable probability density functions [21,49] of


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 3<br />

a Lévy motion [9,50] solve the fractional <strong>diffusion</strong> <strong>equation</strong> [32,57]. This is<br />

because fractional derivatives arise from sums of random movements with<br />

power law probability tails [56,31], <strong>for</strong> which the usual central limit theorem<br />

is replaced by its heavy tail analogue [21,36].<br />

An alternative discrete time model <strong>for</strong> population growth and dispersal<br />

uses an integro-difference <strong>equation</strong><br />

u(x, t + τ) =<br />

∫ ∞<br />

−∞<br />

k τ (x, y)g τ (u(y, t))dy (3)<br />

where the dispersal kernel k τ (x, y) is the probability density of the location<br />

x to which an individual disperses at time t + τ, given that the individual<br />

is located at spatial coordinate y at time t [30,41]. Equation (3) can be<br />

understood as a two-step process: First we apply the iteration <strong>equation</strong> u →<br />

g τ (u) to grow the population over one time step, and then we integrate<br />

against the kernel k τ (x, y) to disperse the population over the same time step.<br />

The classical <strong>reaction</strong>-<strong>diffusion</strong> model (1) can be approximated (in a way we<br />

will make precise later in this paper) with the discrete time model (3) where<br />

g(u 0 ) is the solution to the ordinary differential <strong>equation</strong> ˙u = f(u); u(0) = u 0<br />

at time t = τ and<br />

k τ (x, y) = √ 1<br />

4πDτ<br />

e − (x−y)2<br />

4Dτ (4)<br />

is a Gaussian dispersal kernel, the fundamental solution to the classical <strong>diffusion</strong><br />

<strong>equation</strong><br />

∂u<br />

∂t = D ∂2 u<br />

; u(x, t = 0) = δ(x − y)<br />

∂x2 at time t = τ. In the same manner, we will show in this paper that solutions to<br />

the fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (2) can be usefully approximated by<br />

the same iteration <strong>equation</strong> (3) with the only change being that the dispersal<br />

kernel k τ (x, y) is the fundamental solution to the fractional <strong>diffusion</strong> <strong>equation</strong><br />

∂u<br />

∂t = D ∂α u<br />

; u(x, t = 0) = δ(x − y)<br />

∂xα at time t = τ. This dispersal kernel can be explicitly computed [14,42,49].<br />

Ecologists have also experimented with other alternative kernels incorporating<br />

skewness and heavier probability tails, such as the Gamma, Laplace,<br />

Weibull, and Student-t distribution [13,16,17,30,39]. This paper will also<br />

show how those models relate to a variant of the fractional <strong>reaction</strong>-<strong>diffusion</strong><br />

<strong>equation</strong> (2), and the exact manner in which the discrete time evolution<br />

<strong>equation</strong> (3) can be used to approximate the solution of the continuous time<br />

<strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> in this case. Our methods also provide explicit<br />

error bounds on the approximation, which are useful in numerical simulations.<br />

These methods also extend to the case where the <strong>reaction</strong> term f(u, x)<br />

explicitly depends on the space variable x. This includes, <strong>for</strong> example, the<br />

Fisher <strong>equation</strong> with space-variable intrinsic growth rate r(x) and carrying<br />

capacity K(x). Finally we note that, since <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong>s are<br />

also important in other areas where anomalous <strong>diffusion</strong> is widely observed,<br />

including cell biology, chemistry, physics, geophysics, and finance, it seems<br />

likely that the results reported here will also be useful in a broader context.


4 Baeumer, Kovács, Meerschaert<br />

2 Problem outline<br />

The classical <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (1) and its fractional analogue (2)<br />

are both special cases of a general <strong>for</strong>m<br />

∂u<br />

∂t (x, t) = (Au)(x, t) + f(u(x, t)), u(x, 0) = u 0(x) (5)<br />

where A is a pseudo-differential operator [25] that appears as the generator<br />

of some continuous convolution semigroup [1,44]. Our goal in this section is<br />

to explore the connection between this continuous time evolution <strong>equation</strong><br />

and its discrete time analogue, the integro-difference <strong>equation</strong><br />

u n+1 (x) =<br />

∫ ∞<br />

−∞<br />

k τ (x, y)g τ (u n (y)) dy (6)<br />

where u n (x) = u(x, nτ) with some τ > 0 fixed. Our general approach is based<br />

operator theory <strong>for</strong> abstract differential <strong>equation</strong>s [1,20,44] and infinitely<br />

divisible probability distributions [21,36].<br />

Example 1 To demonstrate the basic idea let us consider the classical <strong>reaction</strong><strong>diffusion</strong><br />

<strong>equation</strong> (1) with a growth term f(u) = ru <strong>for</strong> some real r ≥ 0. It<br />

is well know that if r = 0 then the solution is given by<br />

u(x, t) =<br />

∫ ∞<br />

−∞<br />

1<br />

√ e − (x−y)2<br />

4Dt u 0 (y) dy := [T (t)u 0 ](x)<br />

4πDt<br />

where the dispersal operator T (t) maps the initial condition at time t = 0,<br />

the function u 0 (x), to the solution of the <strong>diffusion</strong> <strong>equation</strong> at time t > 0.<br />

Likewise, if D = 0 then the solution is given by<br />

u(x, t) = e rt u 0 (x) := [S(t)u 0 ](x).<br />

where the growth operator S(t) maps the initial condition to the solution of<br />

the growth <strong>equation</strong> at time t > 0. Clearly, [S(t+s)u 0 ](x) = [S(t)S(s)u 0 ](x) =<br />

[S(s)S(t)u 0 ](x) and [T (t + s)u 0 ](x) = [T (t)T (s)u 0 ](x) = [T (s)T (t)u 0 ](x) <strong>for</strong><br />

all t, s ≥ 0 (the semigroup property). Using the product rule its is easy to see<br />

that the solution to (1) is of the <strong>for</strong>m<br />

u(x, t) =<br />

∫ ∞<br />

−∞<br />

√ 1<br />

4πDt<br />

e − (x−y)2<br />

4Dt e rt u 0 (y) dy = [T (t)S(t)u 0 ](x) (7)<br />

= [S(t)T (t)u 0 ](x) = e rt ∫ ∞<br />

−∞<br />

√ 1<br />

4πDt<br />

e − (x−y)2<br />

4Dt u 0 (y) dy, (8)<br />

that is; the solution of the complex problem (1) can be written using the<br />

solution of the simpler sub-problems. Now if we fix τ > 0, then by (7) and<br />

(8) we have<br />

u n+1 (x) = u(x, (n + 1)τ) = [T ((n + 1)τ)S((n + 1)τ)u 0 ](x)<br />

= [(T (τ)) n+1 (S(τ)) n+1 u 0 ](x) = [T (τ)S(τ)(T (τ)) n (S(τ)) n u 0 ](x)<br />

= [T (τ)S(τ)T (nτ)S(nτ)u 0 ](x) = [T (τ)S(τ)u n ](x)<br />

=<br />

∫ ∞<br />

−∞<br />

√ 1<br />

4πDτ<br />

e − (x−y)2<br />

4Dτ e rτ u n (y) dy :=<br />

∫ ∞<br />

−∞<br />

k τ (x, y)g τ (u n (y)) dy.


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 5<br />

This means that the exact solution at time t = nτ, <strong>for</strong> any positive integer<br />

n = 1, 2, . . . satisfies an iteration of the <strong>for</strong>m (6) with k τ given by (4) and<br />

g τ (u n (y)) = e rτ u n (y). Iterations of the <strong>for</strong>m u n+1 = S(τ)T (τ)(u n ) are called<br />

sequential splitting.<br />

We now try to follow the same basic idea in the case when both the differential<br />

operator and the growth term are more complicated. Assume that<br />

the continuous model is well posed in the sense that given u 0 we obtain a<br />

unique solution on R × [0, T ]. In this case, at least <strong>for</strong>mally, the solution can<br />

be written as u(x, t) = [W (t)u 0 ](x) and we call W (t) the solution operator<br />

of (5). In Example 1 the solution operator is given by (7) and since T (t)<br />

and S(t) commute W (t) = T (t)S(t). If we would be able to construct W (t)<br />

as in Example 1, then the solution of (5) would satisfy the natural iteration<br />

u n+1 (x) = [W (τ)u n ](x). Un<strong>for</strong>tunately if the function f is not just a multiplication<br />

by a constant, then it is usually impossible to obtain a closed <strong>for</strong>m<br />

of W (t). There<strong>for</strong>e, we might want to follow the idea of sequential splitting<br />

here, too, that is; instead of looking at (5) we look at two separate differential<br />

<strong>equation</strong>s which are easier to solve and construct an iteration based on<br />

their solutions. Of course we then have to investigate what this has to do<br />

with W (t), the solution operator of (5). Let us denote by T (t) the solution<br />

operator to the pseudo-differential <strong>equation</strong><br />

and by S(t) the solution operator to<br />

∂u<br />

(x, t) = (Au)(x, t) (9)<br />

∂t<br />

∂u<br />

(x, t) = f(u(x, t)). (10)<br />

∂t<br />

The idea behind sequential splitting is that <strong>for</strong> τ small, we solve first (10)<br />

on [0, τ] and then using the obtained solution at time t = τ as initial data<br />

<strong>for</strong> (9) we solve that on [0, τ]. Then hope that this will be close to the result<br />

if we would do both at the same time. Mathematically, instead of computing<br />

[W (τ)u 0 ](x) we compute [T (τ)S(τ)u 0 ](x). Now if S(t) and T (t) commute,<br />

then this is the same thing. If not, then it is usually different. Assume<br />

that we are interested in the solution of (5) at time t = nτ. Then, by the<br />

above described idea, we set up an iteration u n+1 (x) = [T (τ)S(τ)u n ](x) with<br />

starting value u 0 , or equivalently, u n+1 (x) = [(T (τ)S(τ)) n u 0 ] (x). Usually,<br />

u n (x) ≠ u(x, nτ) (except <strong>for</strong> the commuting case), but is it true that u n (x)<br />

is close to u(x, nτ) in some sense For example, if we fix t and choose finer<br />

and finer time steps τ = t n , then does u n(x) converge to u(x, t) in some<br />

sense as n → ∞, in other words, does [ T ( t n )S( t n )] n<br />

u0 converge to W (t)u 0<br />

as n → ∞<br />

Example 2 Let us consider the fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> with<br />

Fisher’s growth term<br />

( )<br />

∂u<br />

∂t (x, t) = ru(x, t) u(x, t)<br />

1 − + D ∂α u<br />

K ∂x α (x, t); u(x, 0) = u 0(x) (11)


6 Baeumer, Kovács, Meerschaert<br />

where 1 < α ≤ 2. To apply the sequential operator splitting procedure, we<br />

first consider the fractional <strong>diffusion</strong> <strong>equation</strong><br />

∂u<br />

∂t (x, t) = D ∂α u<br />

∂x α (x, t); u(x, 0) = u 0(x) (12)<br />

which ∫ can be solved by Fourier trans<strong>for</strong>m methods [15,32]: Write û(λ, t) =<br />

e −iλx u(x, t)dx and recall that ∂ α u/∂x α has Fourier trans<strong>for</strong>m (iλ) α û. Taking<br />

Fourier trans<strong>for</strong>ms on both sides of (12) yields the ordinary differential<br />

<strong>equation</strong><br />

d<br />

dtû(λ, t) = D(iλ)α û(λ, t); û(λ, 0) = û 0 (λ)<br />

whose solution is evidently of the <strong>for</strong>m<br />

û(λ, t) = û 0 (λ) exp(tD(iλ) α )<br />

and now we use the fact from probability theory that exp(C(iλ) α ) is known<br />

to be the Fourier trans<strong>for</strong>m of an α-stable probability density, which appears<br />

in the generalised central limit theorem [21,36,49]. Then we may write<br />

u(x, t) = [T (t)u 0 ](x) =<br />

∫ ∞<br />

−∞<br />

k t (x − y)u 0 (y) dy (13)<br />

where k t (x) is an α-stable probability density. Next consider the growth<br />

<strong>equation</strong><br />

( )<br />

∂u<br />

∂t (x, t) = ru(x, t) u(x, t)<br />

1 − ; u(x, 0) = u 0 (x)<br />

K<br />

with solution<br />

u(x, t) = [S(t)u 0 ](x)<br />

= [ ˜S(t)](u 0 (x)) = K<br />

(<br />

)<br />

K − u 0 (x)<br />

1 −<br />

K + u 0 (x)(exp(rt) − 1)<br />

(14)<br />

where we emphasise that the nonlinear operator S(t) maps the function u 0 to<br />

another real-valued function of x, and ˜S(t) is a function that maps the real<br />

number u 0 (x) to another real number. According to the discussion above,<br />

define an iteration based on the sequential splitting by<br />

u n+1 (x) = [T (τ)S(τ)u n ](x)<br />

=<br />

∫ ∞<br />

−∞<br />

k τ (x − y)K<br />

(<br />

)<br />

K − u n (y)<br />

1 −<br />

dy.<br />

K + u n (y)(exp(rτ) − 1)<br />

(15)<br />

This is a discrete time evolution <strong>equation</strong> of the <strong>for</strong>m (6), an approximate<br />

solution u(x, nτ) of the fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (11). Our goal<br />

is to show that this approximate solution converges to the actual solution as<br />

the time step τ tends to zero.


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 7<br />

Remark 1 The splitting method (15) has a useful biological interpretation.<br />

Given a typical time scale τ, the <strong>for</strong>mula (15) expresses that the population<br />

first increases via an application of the growth operator S(τ), and then<br />

spreads out via an application of the dispersal operator T (τ). Similarly, if we<br />

model first dispersion and then growth, we obtain the alternative sequential<br />

splitting<br />

u n+1 (x) := [S(τ)T (τ)u n ](x)<br />

⎛<br />

(∫ ∞<br />

K −<br />

= K ⎜<br />

⎝ 1 −<br />

−∞<br />

)<br />

k τ (x − y)u n (y) dy<br />

)<br />

(∫ ∞<br />

K + k τ (x − y)u n (y) dy<br />

−∞<br />

⎞<br />

⎟<br />

⎠ . (16)<br />

(exp(rτ) − 1)<br />

In this paper, we will show that both approaches lead to the same continuous<br />

time limit.<br />

In the next section, we will prove that solutions to the discrete time<br />

integro-difference <strong>equation</strong> with an α-stable dispersal kernel converge to solutions<br />

of the fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> as the time step tends to<br />

zero. We will do this by proving a more general result, that also applies to<br />

other useful dispersal kernels. If the growth term f(u) = ru(1 − u/K) as in<br />

Example 2, we will also show that these splitting procedures provide lower<br />

and upper bounds <strong>for</strong> solutions to the continuous time model. These bounds<br />

will be useful in the numerical methods discussed in Section 4.<br />

3 Mathematical details<br />

Let X be a Banach space of functions u : R → R with associated norm<br />

‖u‖, and consider the abstract <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (5) which we will<br />

rewrite here in operator theory notation as<br />

˙u(t) = Au(t) + f(u(t)), t > 0, u(0) = u 0 (17)<br />

to emphasise our point of view that u : [0, ∞) → X and f : X → X. Here<br />

A is the generator of a strongly continuous semigroup {T (t)} t≥0 on X, a one<br />

parameter family of linear operators T (t) : X → X such that: T (0) = I the<br />

identity operator (Iu = u); each T (t) is bounded, meaning that there exists<br />

a real number M > 0 depending on t > 0 such that ‖T (t)u‖ ≤ M‖u‖ <strong>for</strong> all<br />

u ∈ X; T (t+s) = T (t)T (s) <strong>for</strong> t, s ≥ 0; t ↦→ T (t)u is continuous in the Banach<br />

space norm <strong>for</strong> all u ∈ X; and the generator Au = lim h→0+ h −1 (T (h)u − u)<br />

exists <strong>for</strong> at least some nonzero u ∈ X. We call the set D(A) ⊂ X <strong>for</strong><br />

which this limit exists the domain of the linear operator A, and we say that<br />

the semigroup T (t) is generated by A. We say that u : [0, δ) → X is a<br />

local classical/strong solution of (17) if u is continuous on [0, δ), continuously<br />

differentiable on (0, δ), u(t) ∈ D(A) <strong>for</strong> t ∈ (0, δ) and u satisfies (17) on<br />

(0, δ). If δ can be chosen arbitrarily large then u is a global classical/strong


8 Baeumer, Kovács, Meerschaert<br />

solution of (17). A function u : [0, δ) → X is a local mild solution of (17) if u<br />

is continuous and satisfies the corresponding integral <strong>equation</strong><br />

u(t) = T (t)u 0 +<br />

∫ t<br />

0<br />

T (t − s)f(u(s)) ds, 0 ≤ t < δ. (18)<br />

We note that the integral in (18) is a Bochner integral [1,24,44], an extension<br />

of the Lebesgue integral to the Banach space setting which coincides with a<br />

Riemann integral if the integrand is continuous in the Banach space norm.<br />

If δ can be chosen arbitrarily large then u is a global mild solution of (17).<br />

If f : X → X is Lipschitz continuous, that is; ||f(u) − f(v)|| ≤ M||u − v||<br />

<strong>for</strong> all u, v ∈ X, then <strong>for</strong> all u 0 ∈ X there is a unique global mild solution<br />

u(t) := W (t)u 0 of (17) with ||W (t)u 0 − W (t)v 0 || ≤ M T ||u 0 − v 0 ||, t ∈ [0, T ]<br />

(see, <strong>for</strong> example, [44, Section 6.1]). Denote the unique mild solution of (17)<br />

in the special case A ≡ 0 by u(t) := S(t)u 0 . In this case T (t)u = u <strong>for</strong> all<br />

t ≥ 0, and hence it follows from (18) that<br />

u(t) = u 0 +<br />

∫ t<br />

0<br />

f(u(s)) ds (19)<br />

<strong>for</strong> all t > 0. Then u is also a strong solution, since if u and f are continuous,<br />

then t ↦→ f(u(t)) is also continuous and t ↦→ ∫ t<br />

f(u(s)) ds is differentiable,<br />

∫<br />

0<br />

and d t<br />

dt<br />

f(u(s)) ds = f(u(t)) (see, e.g., [24, page 67]). There<strong>for</strong>e t → u(t) is<br />

0<br />

differentiable in view of (19), and ˙u(t) = f(u(t)). It also follows from (19) that<br />

the collection of nonlinear operators {S(t)} t≥0 <strong>for</strong>ms a semigroup, sometimes<br />

called the flow of the abstract differential <strong>equation</strong> ˙u = f(u). We say that the<br />

collection S(·) is generated by f. It is well known that the solution operator<br />

W (t)u 0 to the abstract <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (17) can be computed by<br />

the Trotter Product Formula<br />

[<br />

W (t)u 0 = lim T (<br />

t<br />

n→∞ n )S( t n )] n [<br />

u0 = lim S(<br />

t<br />

n→∞ n )T ( t n )] n<br />

u0 , u 0 ∈ X, (20)<br />

see, <strong>for</strong> example, [12,18,38]. This result means that the mild solution to the<br />

abstract <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (17) can be computed as an approximation<br />

using the mild solutions of the two sub-problems, the abstract <strong>reaction</strong><br />

<strong>equation</strong> ˙u = f(u) and the abstract <strong>diffusion</strong> <strong>equation</strong> ˙u = Au.<br />

The following proposition is a simplified variant of [18, Theorem 15]. We<br />

call a Banach space X an ordered Banach space if it is a real Banach space<br />

endowed with a partial ordering ≤ such that<br />

1. u ≤ v implies u + w ≤ v + w <strong>for</strong> all u, v, w ∈ X.<br />

2. u ≥ 0 implies λu ≥ 0 <strong>for</strong> all u ∈ X and λ ≥ 0.<br />

3. 0 ≤ u ≤ v implies ||u|| ≤ ||v|| <strong>for</strong> all u, v ∈ X.<br />

4. The positive cone X + := {x ∈ X : x ≥ 0} is closed.<br />

A typical example of an ordered Banach space is C 0 (R), the space of continuous<br />

functions u : R → R such that u(x) → 0 as |x| → ∞, endowed with the<br />

supremum norm ‖u‖ = sup{|u(x)| : x ∈ R}, and endowed with the partial<br />

ordering u ≤ v whenever u(x) ≤ v(x) <strong>for</strong> all x ∈ R. Another example is L p (R)<br />

(1 ≤ p ≤ ∞) endowed with the partial ordering u ≤ v whenever u(x) ≤ v(x)


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 9<br />

<strong>for</strong> x ∈ R almost everywhere. An operator A on an ordered Banach space is<br />

called positive if 0 ≤ u ≤ v implies 0 ≤ Au ≤ Av. We also write B ≤ A if<br />

0 ≤ Bu ≤ Au <strong>for</strong> any u ≥ 0.<br />

Proposition 1 Let X be an ordered Banach space, and assume that the<br />

strongly continuous semigroup T (·) generated by the linear operator A and<br />

the nonlinear semigroup S(·) generated by the Lipschitz continuous function<br />

f are positive. If<br />

T (t)S(t)u 0 ≤ S(t)T (t)u 0 (21)<br />

holds <strong>for</strong> all t ∈ [0, T ] and u 0 ≥ 0, then the unique mild solution W (t)u 0 of<br />

the abstract <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (17) satisfies<br />

[<br />

T (<br />

t<br />

n )S( t n )] n<br />

u0 ≤ [ T ( t<br />

2n )S( t<br />

2n )] 2n<br />

u0 ≤ W (t)u 0<br />

<strong>for</strong> all u 0 ≥ 0, n ∈ N and t ∈ [0, T ].<br />

≤ [ S( t<br />

2n )T ( t<br />

2n )] 2n<br />

u0 ≤ [ S( t n )T ( t n )] n<br />

u0 (22)<br />

Proof It follows from our assumption that W (t)u 0 is given by (20). Next we<br />

show that<br />

[<br />

T (<br />

t<br />

n )S( t n )] n<br />

u0 ≤ [ T ( t<br />

2n )S( t<br />

2n )] 2n<br />

u0<br />

≤ [ S( t<br />

2n )T ( t<br />

2n )] 2n<br />

u0 ≤ [ S( t n )T ( t n )] n<br />

u0 , (23)<br />

u 0 ≥ 0, n ∈ N and t ∈ [0, T ]. For u 0 ≥ 0 using (21) repeatedly we have<br />

T ( t n )S( t n )u 0 = T ( t<br />

2n )T ( t<br />

2n )S( t<br />

2n )S( t<br />

2n )u 0<br />

≤ T ( t<br />

2n )S( t<br />

2n )T ( t<br />

2n )S( t<br />

2n )u 0 ≤ S( t<br />

2n )T ( t<br />

2n )T ( t<br />

2n )S( t<br />

2n )u 0<br />

≤ S( t<br />

2n )T ( t<br />

2n )S( t<br />

2n )T ( t<br />

2n )u 0 ≤ S( t<br />

2n )S( t<br />

2n )T ( t<br />

2n )T ( t<br />

2n )u 0 = S( t n )T ( t n )u 0.<br />

For any operators A, B : X → X, it is not hard to check that the inequality<br />

0 ≤ B ≤ A implies 0 ≤ B n ≤ A n (n ∈ N) which finishes the proof of (23). Fix<br />

n ∈ N and let x k := [ T ( t )S( t ) ] n2 k u<br />

n2 k n2 k 0 . By (23), x 0 ≤ x 1 ≤ ... ≤ x k ≤ ...<br />

and x k → W (t)u 0 as k → ∞ by (20). There<strong>for</strong>e, it follows from the closedness<br />

of the positive cone X + that x k ≤ W (t)u 0 <strong>for</strong> all k = 0, 1, .... This shows the<br />

second inequality in (3). A similar argument yields the third inequality and<br />

the proof is complete.<br />

⊓⊔<br />

One useful class of strongly continuous semigroups are the infinitely divisible<br />

semigroups, which are associated with certain families of probability<br />

densities. Suppose that Y is a random variable on R with probability density<br />

k(y) and Fourier trans<strong>for</strong>m ˆk(λ) = ∫ e −iλy k(y)dy. Let k n = k ∗ · · · ∗ k denote<br />

the n−fold convolution of k with itself. We say that Y (or k) is infinitely<br />

divisible if <strong>for</strong> each n = 1, 2, 3, . . . there exist independent random variables<br />

Y n1 , . . . , Y nn with the same density k n such that Y n1 + · · · + Y nn is identically<br />

distributed with Y . The normal, Cauchy, double-Gamma, Laplace, α-stable,<br />

and Student-t densities are all infinitely divisible. The Lévy representation


10 Baeumer, Kovács, Meerschaert<br />

(see, e.g., Theorem 3.1.11 in [36]) states that if k is infinitely divisible then<br />

ˆk(λ) = e ψ(λ) where<br />

ψ(λ) = −iλa − 1 ∫ (<br />

2 λ2 b 2 + e −iλy − 1 +<br />

iλy )<br />

y≠0<br />

1 + y 2 φ(y)dy, (24)<br />

where a ∈ R, b ≥ 0 , and the jump intensity φ satisfies<br />

∫<br />

min{1, y 2 } φ(y)dy < ∞.<br />

y≠0<br />

The triple [a, b, φ] is unique, and we call this the Lévy representation of<br />

the infinitely divisible density k. It follows that we can define the convolution<br />

power k t to be the infinitely divisible density with Lévy representation<br />

[ta, tb, tφ], so that k t has Fourier trans<strong>for</strong>m e tψ(k) <strong>for</strong> any t ≥ 0. It is well<br />

known (e.g., see [25, Example 4.1.3]) that every infinitely divisible density is<br />

associated with a strongly continuous semigroup on C 0 (R) defined by<br />

[T (t)u](x) :=<br />

∫ ∞<br />

−∞<br />

k t (x − y) u(y)dy. (25)<br />

The following result allows us to compute the generator of this semigroup<br />

from the Lévy representation.<br />

Proposition 2 Every function u ∈ C 0 (R) with u ′ , u ′′ ∈ C 0 (R) belongs to the<br />

domain of the generator A of the semigroup (25), and <strong>for</strong> such functions we<br />

have<br />

Au(x) = − au ′ (x) + 1 2 b2 u ′′ (x)<br />

∫ (<br />

)<br />

+ u(x − y) − u(x) + u′ (x)y<br />

1 + y 2 φ(y)dy.<br />

y≠0<br />

(26)<br />

Proof Example 4.1.12 in [25] shows that, <strong>for</strong> any u ∈ Cc<br />

∞ (R), the space of<br />

infinitely differentiable functions u : R → R that vanish off a compact set, u<br />

belongs to the domain of A and (26) holds. Then it follows from Corollary<br />

2.4 of [54] that the same holds <strong>for</strong> all u ∈ C 0 (R) with u ′ , u ′′ ∈ C 0 (R). ⊓⊔<br />

Remark 2 The integral <strong>for</strong>mula in the Lévy representation (24) is a slight<br />

generalisation of the <strong>for</strong>mula <strong>for</strong> the Fourier trans<strong>for</strong>m of a compound Poisson<br />

density, and indeed any infinitely divisible density is in some sense the<br />

limiting case of a compound Poisson [36, Remark 3.1.18], a random sum with<br />

a Poisson distributed number of summands where the relative frequency of<br />

jumps of size y is given by the jump intensity φ(y). The generator <strong>for</strong>mula<br />

comes from the fact that T (t)u is a convolution k t ∗ u so that its Fourier<br />

trans<strong>for</strong>m is e tψ(λ) û(λ), and then the generator can be computed in Fourier<br />

space by t −1 [e tψ(λ) − 1]û(λ) → ψ(λ)û(λ) so that ψ(λ)û(λ) is the Fourier<br />

trans<strong>for</strong>m of Au(x). Then (26) comes from inverting this Fourier trans<strong>for</strong>m<br />

using (24) and the fact that multiplication by iλ, e −iλy corresponds with<br />

taking the derivative or shifting in real space, respectively.


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 11<br />

In what follows we discuss a general abstract <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong><br />

<strong>for</strong> Fisher’s <strong>equation</strong> (11) <strong>for</strong> non-negative initial data on X := C 0 (R):<br />

˙u(t) = Au(t) + f(u(t)), u(0) = u 0 ≥ 0, (27)<br />

where f : X → X is defined via the function ˜f : R → R as<br />

(<br />

[f(u)](x) = ˜f(u(x)) := ru(x) 1 − u(x) )<br />

K<br />

We introduce a cut off function via the function f ˜ N : R → R as<br />

⎧<br />

0<br />

⎪⎨ (<br />

if u(x) < 0<br />

[f N (u)](x) = ˜f N (u(x)) : ru(x) 1 − u(x) )<br />

if 0 ≤ u(x) ≤ NK<br />

K<br />

⎪⎩<br />

rNK(1 − N) if u(x) > NK.<br />

(28)<br />

(29)<br />

Proposition 3 Let X := C 0 (R) and let f be given by (28). Then the abstract<br />

differential <strong>equation</strong><br />

˙u(t) = f(u(t)), u(0) = u 0 ≥ 0 (30)<br />

has a unique strong global solution given by u(t) = S(t)u 0 <strong>for</strong> each nonnegative<br />

u 0 ∈ X, and in fact we have [S(t)u 0 ](x) = [ ˜S(t)](u 0 (x)) where ˜S(t) is<br />

defined by (14). For any positive integer N ≥ 2, the abstract differential <strong>equation</strong><br />

˙u = f N (u), u(0) = u 0 ≥ 0 <strong>for</strong> the Lipschitz continuous function f N defined<br />

in (29) also has a unique strong global solution given by u(t) = S N (t)u 0<br />

<strong>for</strong> each u 0 ≥ 0 in X. Furthermore, in this case, if 0 ≤ u 0 (x) ≤ NK <strong>for</strong> all<br />

x ∈ R, then we also have that 0 ≤ [S(t)u 0 ](x) = [S N (t)u 0 ](x) ≤ NK <strong>for</strong> all<br />

x ∈ R and all t ≥ 0.<br />

Proof Consider the abstract initial value problem<br />

˙u(t) = f N (u(t)), u(0) = u 0 ≥ 0 ∈ X<br />

which has, by the Lipschitz continuity of f N , a unique global strong solution<br />

u(t) = S N (t)u 0 (see the discussion after (19)), where S N (·) is the nonlinear<br />

semigroup generated by f N . Hence, since the operator norm in this space is<br />

the supremum norm, it follows that the function u x (t) := [S N (t)u 0 ](x) is <strong>for</strong><br />

each fixed x ∈ R the unique solution of the ordinary differential <strong>equation</strong><br />

d<br />

dt u x(t) = ˜f N (u x (t)), u x (0) = u 0 (x) ≥ 0 ∈ R.<br />

Then it follows easily, using the uniqueness of solutions, that S N (·) is positive,<br />

i.e., if u 0 (x) ≥ 0 <strong>for</strong> all x ∈ R then u x (t) = [u(t)](x) = [S N (t)u 0 ](x) ≥<br />

0 <strong>for</strong> all x ∈ R, and also if u 0 (x) ≥ v 0 (x) <strong>for</strong> all x ∈ R then u x (t) =<br />

[u(t)](x) = [S N (t)u 0 ](x) ≥ v x (t) = [v(t)](x) = [S N (t)v 0 ](x) <strong>for</strong> all x ∈ R.<br />

Since ˜f N (y) < 0 <strong>for</strong> all y > K it also follows from uniqueness of solutions<br />

that, if u 0 (x) ≤ NK <strong>for</strong> all x ∈ R, then [S N (t)u 0 ](x) ≤ NK <strong>for</strong> all x ∈ R.


12 Baeumer, Kovács, Meerschaert<br />

Moreover, since ˜f N (u) = ˜f(u) <strong>for</strong> 0 ≤ u ≤ NK it follows that S N (t)u 0 also<br />

solves<br />

˙u(t) = f(u(t)), u(0) = u 0 . (31)<br />

Since f is locally Lipschitz, Theorem [24, Chapter 3, Theorem 3.4.1] also implies<br />

that (31) has a unique strong local solution. The function t ↦→ S N (t)u 0<br />

is defined <strong>for</strong> all t ≥ 0 and hence S N (t)u 0 is the unique strong global solution<br />

S(t)u 0 of (31) which is necessarily given by (14) since that solves (31)<br />

evaluated pointwise.<br />

⊓⊔<br />

Now we come to the main result of this paper. It connects an abstract<br />

<strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (5) in continuous time with its discrete time counterpart<br />

(6), and shows that in the case of a Fisher’s <strong>equation</strong> growth term,<br />

the discrete integro-difference <strong>equation</strong> yields bounds on the solution of the<br />

continuous <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> that converge to the unique solution<br />

of the continuous <strong>equation</strong> as the time step tends to zero.<br />

Theorem 1 Let k be infinitely divisible and let A denote the generator of the<br />

associated strongly continuous convolution semigroup T (t) defined by (25) on<br />

X := C 0 (R). Then (27,28) has a unique mild solution u(t) = W (t)u 0 <strong>for</strong> all<br />

u 0 ≥ 0 in X given by the Trotter Product Formula<br />

W (t)u 0 = lim<br />

n→∞<br />

Moreover, <strong>for</strong> all n ∈ N,<br />

[<br />

T (<br />

t<br />

n )S( t n )] n<br />

u0 = lim<br />

n→∞<br />

[<br />

T (<br />

t<br />

n )S( t n )] n<br />

u0 ≤ [ T ( t<br />

2n )S( t<br />

2n )] 2n<br />

u0 ≤ W (t)u 0<br />

where S(·) is defined in (14).<br />

[<br />

S(<br />

t<br />

n )T ( t n )] n<br />

u0 . (32)<br />

≤ [ S( t<br />

2n )T ( t<br />

2n )] 2n<br />

u0 ≤ [ S( t n )T ( t n )] n<br />

u0 , (33)<br />

Proof Choose N ≥ 2 such that 0 ≤ u 0 (x) ≤ NK <strong>for</strong> all x and consider the<br />

abstract <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong><br />

˙u(t) = Au(t) + f N (u(t)), u(0) = u 0 ≥ 0. (34)<br />

Since f N : X → X is globally Lipschitz continuous and A is a generator,<br />

there is a unique mild solution u N (t) = W N (t)u 0 of (34) given by the Trotter<br />

Product Formula<br />

[<br />

u N (t) = W N (t)u 0 = lim T (<br />

t<br />

n→∞ n )S N ( t n )] n<br />

u0<br />

[<br />

= lim SN ( t<br />

n→∞ n )T ( t n )] n (35)<br />

u0 ,<br />

see, e.g., [12,18,38]. The semigroup T (·) satisfies 0 ≤ T (t)u 0 ≤ T (t)v 0 <strong>for</strong><br />

0 ≤ u 0 ≤ v 0 and t ≥ 0 since k t is a probability density function. If 0 ≤<br />

u 0 (x) ≤ NK <strong>for</strong> all x, then<br />

0 ≤ [T (t)u 0 ](x) ≤ NK<br />

∫ ∞<br />

−∞<br />

k t (x − y)dy = NK (36)


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 13<br />

There<strong>for</strong>e, by (36) and Proposition 3,<br />

<strong>for</strong> all x and<br />

0 ≤ [T ( t n )S N ( t n )]n u 0 (x) = [T ( t n )S( t n )]n u 0 (x) ≤ NK (37)<br />

0 ≤ [S N ( t n )T ( t n )]n u 0 (x) = [S( t n )T ( t n )]n u 0 (x) ≤ NK (38)<br />

<strong>for</strong> all x. This also shows that 0 ≤ [u N (t)](x) ≤ NK <strong>for</strong> all x in view of (35).<br />

There<strong>for</strong>e u N (t) is a mild solution of (27), too, since f N (u(x)) = f(u(x)) <strong>for</strong><br />

0 ≤ u(x) ≤ NK. Since f is locally Lipschitz, a well known result [44, Chapter<br />

6, Theorem 1.4] implies that (27) has a unique local mild solution and since<br />

u N (t) is defined <strong>for</strong> all t > 0 it follows that u N (t) is the unique global mild<br />

solution of (27) and is given by the Trotter product <strong>for</strong>mula (32) in view of<br />

(35), (37) and (38). Finally, an easy computation shows that the function<br />

y ↦→ [ ˜S(t)](y) in (14) is concave down on y > 0 <strong>for</strong> any t > 0. There<strong>for</strong>e, by<br />

Jensen’s inequality [21, pp. 153–154],<br />

[∫ ∞<br />

]<br />

[S(t)T (t)u 0 ](x) = ˜S(t) k t (y)u 0 (x − y) dy<br />

−∞<br />

Thus,<br />

≥<br />

=<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

k t (y) ˜S(t) [u 0 (x − y)] dy<br />

k t (y) [S(t)u 0 ] (x − y) dy = [T (t)S(t)u 0 ](x).<br />

S N (t)T (t)u 0 = S(t)T (t)u 0 ≥ T (t)S(t)u 0 = T (t)S N (t)u 0<br />

which finishes the proof by Proposition 1, (37) and (38).<br />

Corollary 1 Under the assumptions of Theorem 1, if u 0 and its first two<br />

derivatives in x exist and and belong to C 0 (R), then (27,28) has a unique<br />

strong solution u(t) = W (t)u 0 <strong>for</strong> all u 0 ≥ 0 in X given by the Trotter<br />

Product Formula (32).<br />

Proof In this case Proposition 2 shows that u 0 is in the domain of the operator<br />

A, and since f : X → X is continuously differentiable, u is also a strong<br />

solution by [44, Chapter 6, Theorem 1.5].<br />

⊓⊔<br />

Theorem 1 shows that the fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (11) in<br />

continuous time can be solved numerically by computing solutions to one of<br />

its discrete time counterparts (15) or (16) with τ = t/n. The approximate<br />

solutions u n (x, t) converge to the unique solution u(x, t) of the fractional<br />

<strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> at any time t > 0 <strong>for</strong> any smooth initial population<br />

density u 0 (x) as n → ∞. This result links the continuous time partial<br />

differential <strong>equation</strong> model with the corresponding discrete time integrodifference<br />

<strong>equation</strong> model. Furthermore, the approximation (15) gives a lower<br />

bound to the exact solution while the approximation (16) gives an upper<br />

bound. Hence these two approximation can be compared to yield exact error<br />

bounds. See Section 4 <strong>for</strong> an illustration.<br />

⊓⊔


14 Baeumer, Kovács, Meerschaert<br />

Remark 3 Theorem 1 extends immediately to any abstract <strong>reaction</strong>-<strong>diffusion</strong><br />

<strong>equation</strong> of the <strong>for</strong>m (27) as long as u ↦→ [ ˜S(t)](u) is concave down on some<br />

interval u ∈ [0, M] <strong>for</strong> each t > 0, and solutions to the differential <strong>equation</strong><br />

˙u = f(u) remain in this interval <strong>for</strong> all time whenever u(0) ∈ [0, M]. It also<br />

extends to the case where f(u, x) depends on the point x in space, and/or<br />

the multidimensional case x ∈ R d , since all the proofs are point-wise. Hence<br />

the results of this paper can also be adapted to model patchy populations in<br />

d-dimensional space, where the growth rate and carrying capacity vary with<br />

spatial location.<br />

The most familiar choice <strong>for</strong> the infinitely divisible semigroup T (t) in<br />

Theorem 1 is the Gaussian semigroup, where k t is the normal density with<br />

mean zero and variance 2Dt <strong>for</strong> some D > 0. This is the semigroup introduced<br />

in Example 1, and since the Lévy representation of k is [0, 2D, 0] (i.e., we have<br />

a = 0, φ ≡ 0, and b 2 = 2D in (24)), <strong>equation</strong> (26) shows that its generator<br />

is Au = D ∂ 2 u/∂x 2 . In applications to biology, the probability density k t is<br />

the dispersal kernel that describes the dispersion of population over a time<br />

interval of length t. The choice of a Gaussian dispersal kernel is justified by<br />

the central limit theorem [21,36] which states that the accumulation of a<br />

series of independent random movements will converge to a Gaussian <strong>for</strong>m<br />

as the number of movements increases to infinity. This is the same argument<br />

that justifies the Brownian motion model in cell and molecular biology.<br />

There are several reasons why the Gaussian dispersal kernel might not<br />

apply as a model <strong>for</strong> the spread of a population density. First of all, the time<br />

interval may be too short <strong>for</strong> the limit theorem to apply. Second, correlation<br />

in movements can invalidate the Gaussian limit. Finally, and most important<br />

<strong>for</strong> the purposes of this paper, heavy tailed movements can also invalidate<br />

the Gaussian limit. Recent empirical research [13,16,17,26,28,43,58–60] has<br />

produced a body of evidence indicating a power law distributions of movements,<br />

so that if R is the magnitude (radius) of the movement of a randomly<br />

selected member of the population, then we have P (R > r) ≈ Cr −α <strong>for</strong> r<br />

large. A log-log plot of such movements will exhibit as a straight line with<br />

slope −α (power law tail), see <strong>for</strong> example Figure 2 in [60] or Figure 3 in<br />

[58] or Figure 2 in [13]. This leads to a density function k(r) ≈ Cαr −α−1<br />

<strong>for</strong> r large, and so if 0 < α < 2, then an easy calculation shows that the<br />

second moment ∫ r 2 k(r)dr = ∞. The non-existence of the second moment<br />

invalidates the Gaussian central limit theorem, and in this case an alternative<br />

limit theorem applies [21,36] in which the Gaussian limit distribution is<br />

replaced by the more general Lévy stable distribution. The stable probability<br />

density function cannot be written in closed <strong>for</strong>m, except in a few special<br />

cases, so it is common to describe these distributions in terms of their Fourier<br />

trans<strong>for</strong>ms ˆk(λ) = e ψ(λ) where<br />

⎧<br />

⎨ −iaλ − σ α |λ| α (1 − iβ(signλ) tan πα 2 ) <strong>for</strong> α ≠ 1<br />

ψ(λ) =<br />

⎩<br />

−iaλ − σ|λ|(1 + iβ 2 π<br />

(signλ) ln λ) <strong>for</strong> α = 1<br />

where the stable index 0 < α ≤ 2, the skewness −1 ≤ β ≤ 1, the center<br />

a ∈ R, and the scale σ ≥ 0 [49]. The Gaussian distribution is the special


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 15<br />

case α = 2, and in that case the skewness is superfluous. Stable distributions<br />

represent the only possible limits of normalised sums of independent random<br />

variables, so they are in some sense universal models <strong>for</strong> <strong>diffusion</strong> processes.<br />

Stable densities generalise the more familiar Gaussian densities to include<br />

the possibility of skewness and heavy power law tails, since <strong>for</strong> α < 2 the<br />

stable density k(y) ≈ Cqα|y| −α−1 as y → −∞ and k(y) ≈ Cpαy −α−1 as<br />

y → ∞ where β = p − q. The weights p, q represent the relative chance of<br />

large negative versus positive movements, respectively. In the symmetric case<br />

one has β = 0 since p = q. Figure 1 shows the standard σ = 1, µ = 0 stable<br />

densities in the case α = 1.6 to illustrate the skewness. The symmetric stable<br />

density is similar to a Gaussian density but with power-law tails.<br />

β = 0<br />

0.30<br />

0.25<br />

β = −1<br />

β =1<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

-6 -4 -2 0 2 4 6<br />

0.00<br />

Fig. 1 Standard stable dispersal kernels with α = 1.6 illustrating the bell shape<br />

and skewness.<br />

For stable semigroups the generator <strong>for</strong>mula is<br />

[ 1 − β ∂ α u<br />

Au = D<br />

2 ∂(−x) α + 1 + β ∂ α ]<br />

u<br />

2 ∂x α<br />

where ∂ α u/∂(−x) α has Fourier trans<strong>for</strong>m (−iλ) α û [3]. This <strong>for</strong>mula is obtained<br />

from (26) by computing the integral in the last term, using the α-stable<br />

jump intensity φ(y) = Cq|y| −α−1 <strong>for</strong> y < 0 and φ(y) = Cpy −α−1 <strong>for</strong> y > 0.<br />

In the symmetric case β = 0, one often writes Au = D∂ α u/∂|x| α the Riesz<br />

fractional derivative. This <strong>for</strong>m also coincides with the classical fractional<br />

power of the Laplacian or second derivative operator [1,24].<br />

The fractional derivative can be computed in real space from the generator<br />

<strong>for</strong>mula (26), see [3]. For example, in the simplest case p = 1 − q = 1 and


16 Baeumer, Kovács, Meerschaert<br />

0 < α < 1 the last term in the integral (26) converges, and hence we can<br />

choose a so that<br />

Au(x) =<br />

∫ ∞<br />

0<br />

u(x − y) − u(x)<br />

y<br />

Cy −α dy. (39)<br />

Then we see that the fractional derivative is a weighted average of difference<br />

quotients, with power law weights deriving from the underlying jump<br />

intensity. Using the fractional derivative generator in the abstract <strong>reaction</strong><strong>diffusion</strong><br />

<strong>equation</strong> (5) corresponds to an α-stable dispersal kernel in the<br />

corresponding discrete-time integro-difference <strong>equation</strong> (6). Along with the<br />

Gaussian kernel (the special case α = 2), these dispersal kernels are distinguished<br />

by the stability property: k t (y) = t −1/α k(yt −1/α ) <strong>for</strong> all y ∈ R and<br />

all t > 0. Hence they are strongly indicated when the shape of the dispersal<br />

kernel is time-independent. Furthermore, since they derive from central<br />

limit theorems, they are in some sense universal. Finally, although these α-<br />

stable probability density functions cannot usually be written in closed <strong>for</strong>m,<br />

probabilists have developed fast numerical methods <strong>for</strong> computing them [42].<br />

Combining this with Theorem 1 allows the efficient solution of the fractional<br />

<strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong>, as well as monotone error bounds. We will illustrate<br />

the application of these methods in the next section.<br />

Remark 4 A wide variety of alternative models can be obtained by considering<br />

different infinitely divisible densities k as the dispersal kernel. The infinitely<br />

divisible densities are convenient because one can define the integrodifference<br />

<strong>equation</strong> (6) at any time scale τ based on the convolution power k τ .<br />

The results of this paper also yield the corresponding <strong>reaction</strong>-<strong>diffusion</strong> model<br />

with a <strong>diffusion</strong> operator A exhibiting as the generator of the corresponding<br />

infinitely divisible semigroup. Lockwood et al. [30] employ a Laplace (double<br />

exponential), or more generally, a double Gamma family of dispersion<br />

kernels, which are infinitely divisible with Lévy representation [ta, 0, tφ] with<br />

jump intensity φ(y) = |y| −1 e −c|y| . In this case the last term in the integral<br />

(26) converges, and hence we can choose a so that<br />

Au(x) =<br />

∫ ∞<br />

−∞<br />

u(x − y) − u(x)<br />

|y|<br />

e −c|y| dy (40)<br />

an exponentially weighted derivative, which is similar to the fractional derivative<br />

<strong>for</strong>mula (39) but with different weights [29]. For an exponential or<br />

Gamma dispersal kernel, the generator <strong>for</strong>mula is the same as (40) except<br />

that the integral is taken over y > 0. Clarke et al. [16,17] employ the Student-t<br />

dispersal kernel, which is infinitely divisible with Lévy representation as specified<br />

in [23, Remark 2.3] in terms of Bessel functions. Substituting into (26)<br />

yields the corresponding generator, connecting the integro-difference <strong>equation</strong><br />

(6) with a Student-t dispersal kernel to the analogous <strong>reaction</strong>-<strong>diffusion</strong><br />

<strong>equation</strong> (5).


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 17<br />

4 Numerical Experiments<br />

Consider a fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> using Fisher’s growth <strong>equation</strong><br />

and a symmetric (Riesz) fractional <strong>diffusion</strong> term of order 1 < α ≤ 2:<br />

( )<br />

∂u<br />

∂t (x, t) = D ∂α u<br />

∂|x| α (x, t) + ru(x, t) u(x, t)<br />

1 − , u(x, 0) = u 0 (x). (41)<br />

K<br />

Solutions to this <strong>equation</strong> can be computed using the results of Theorem 1,<br />

using the integro-difference model (6) as an approximation where the dispersal<br />

kernel k τ (y) is the symmetric α-stable probability density function whose<br />

Fourier trans<strong>for</strong>m is e −Dτ|λ|α <strong>for</strong> some D > 0.<br />

u(x,32)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

t=32<br />

α=1.8<br />

α=2<br />

0<br />

−100 −50 0 50 100<br />

x<br />

Fig. 2 Solution of the fractional Fisher’s <strong>equation</strong> (41) with α = 1.8 versus α = 2<br />

at t = 32 with K = 1, r = 0.25 and D = 0.1 showing heavier tails and faster<br />

spreading in the fractional case α < 2.<br />

Numerical simulations were per<strong>for</strong>med with K = 1, c = 0.25 and D = 0.1<br />

and a smooth step-like initial function u 0 which takes the constant value<br />

u = 0.8 around the origin and rapidly decays to 0 away from the origin. S(τ)<br />

was computed analytically at each time step using the explicit analytical solution<br />

(14) and T (τ)u n was computed by numerically convolving u n against<br />

the symmetric α-stable dispersal kernel k τ , which was computed numerically<br />

using the method of Nolan [42]. Note that the dispersal kernel can be computed<br />

<strong>for</strong> any time scale τ, and that it only has to be computed once <strong>for</strong><br />

each time scale. Figure 2 shows that the use of the conventional <strong>diffusion</strong><br />

term α = 2 in (41) produces a rapidly decaying solution away from the origin.<br />

However, if one replaces the <strong>diffusion</strong> term by a fractional one, even with<br />

an order α which is close to 2, then the solution picks up heavy tails and is<br />

spreading faster, similar to the results reported in del-Castillo-Negrete et al.<br />

[19] <strong>for</strong> the one-sided fractional <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> (2).<br />

Figure 3 visualises the conclusions of Theorem 1; that is, the sequential<br />

splitting approximations (15) and (16) converge in a pointwise monotone<br />

increasing and decreasing fashion, respectively, to the solution of (41).


18 Baeumer, Kovács, Meerschaert<br />

1<br />

t=32<br />

0.8<br />

u(x,32)<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−100 −50 0 50 100<br />

x<br />

Fig. 3 Sequential splitting approximations of the solution of the fractional Fisher’s<br />

<strong>equation</strong> (41) with α = 1.8, K = 1, r = 0.25 and D = 0.1. Dashed lines are<br />

computed from (16) with time step τ = 32, 8, 2 and dotted lines are computed<br />

from (15) at the same time steps to show monotone convergence to the exact (solid<br />

line) solution.<br />

Figure 4 shows evidence of an accelerating front. The selected population<br />

density level u = c has spread a distance x c (t) by time t > 0, and since the<br />

graph of x c (t) versus t <strong>for</strong> various c closely resembles a straight line on the<br />

semi-log plot, we conclude that the front expands exponentially with time,<br />

in agreement with results reported by del-Castillo-Negrete et al. [19] <strong>for</strong> the<br />

one-sided model (2).<br />

Acknowledgements We would like to thank Shona Lamoureaux of AgResearch<br />

New Zealand, Jon Pitch<strong>for</strong>d at the University of Leeds, and Alex James at the<br />

University of Canterbury, New Zealand <strong>for</strong> stimulating discussions.<br />

References<br />

1. Arendt, W., Batty, C., Hieber, M., Neubrander, F.: Vector-valued Laplace<br />

trans<strong>for</strong>ms and Cauchy problems. Monographs in Mathematics. Birkhaeuser-<br />

Verlag, Berlin (2001)<br />

2. Arendt, W., Grabosch, A., Greiner, G., Groh, U., Lotz, H.P., Moustakas, U.,<br />

Nagel, R., Neubrander, F., Schlotterbeck, U.: One-parameter semigroups of<br />

positive operators, Lecture Notes in Mathematics, vol. 1184. Springer-Verlag,<br />

Berlin (1986)<br />

3. Baeumer, B., Meerschaert, M.M.: Stochastic Solutions <strong>for</strong> <strong>Fractional</strong> Cauchy<br />

Problems. <strong>Fractional</strong> Calculus and Applied Analysis 4(4), 481–500 (2001)<br />

4. Baeumer, B., Meerschaert, M.M., Benson, D.A., Wheatcraft, S.W.: Subordinated<br />

advection-dispersion <strong>equation</strong> <strong>for</strong> contaminant transport. Water Resources<br />

Research 37, 1543–1550 (2001)


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 19<br />

10 2 c=0.1<br />

c=0.2<br />

c=0.3<br />

t<br />

Location of x c<br />

10 1<br />

10 0<br />

0 10 20 30 40 50<br />

Fig. 4 Exponential spreading front <strong>for</strong> the fractional Fisher’s <strong>equation</strong> (41) with<br />

α = 1.8, K = 1, r = 0.25 and D = 0.1. The farthest distance to which a population<br />

density of c has spread by time t is plotted on a semi-log scale to illustrate the<br />

(nearly) exponential front growth over time.<br />

5. Barkai, E., Metzler, R., Klafter, J.: From continuous time random walks to the<br />

fractional fokker-planck <strong>equation</strong>. Phys. Rev. E 61, 132–138 (2000)<br />

6. Benson, D., Wheatcraft, S., Meerschaert, M.: Application of a fractional<br />

advection-dispersion <strong>equation</strong>. Water Resources Research 36(6), 1403–1412<br />

(2000)<br />

7. Benson, D.A., Schumer, R., Meerschaert, M.M., Wheatcraft, S.W.: <strong>Fractional</strong><br />

dispersion, Lévy motions, and the MADE tracer tests. Transport in Porous<br />

Media 42, 211–240 (2001)<br />

8. Benson, D.A., Wheatcraft, S.W., Meerschaert, M.M.: The fractional-order governing<br />

<strong>equation</strong> of Lévy motion. Water Resources Research 36, 1413–1424<br />

(2000)<br />

9. Bertoin, J.: Lévy processes, Cambridge Tracts in Mathematics, vol. 121. Cambridge<br />

University Press, Cambridge (1996)<br />

10. Blumen, A., Zumofen, G., Klafter, J.: Transport aspects in anomalous <strong>diffusion</strong>:<br />

Lévy walks. Phys. Rev. A 40, 3964–3973 (1989)<br />

11. Bouchaud, J., Georges, A.: Anomalous <strong>diffusion</strong> in disordered media - statistical<br />

mechanisms, models and physical applications. Phys. Rep. 195, 127–293 (1990)<br />

12. Brézis, H., Pazy, A.: Convergence and approximation of semigroups of nonlinear<br />

operators in Banach spaces. J. Functional Analysis 9, 63–74 (1972)<br />

13. Bullock, J.M., Clarke, R.T.: Long distance seed dispersal by wind: measuring<br />

and modelling the tail of the curve. Oecologia 124(4), 506–521 (2000)<br />

14. Chambers, J.M., Mallows, C.L., Stuck, B.W.: A method <strong>for</strong> simulating stable<br />

random variables. J. Amer. Statist. Assoc. 71(354), 340–344 (1976)<br />

15. Chaves, A.: A fractional <strong>diffusion</strong> <strong>equation</strong> to describe Lévy flights. Phys. Lett.<br />

A 239, 13–16 (1998)<br />

16. Clark, J.S., Silman, M., Kern, R., Macklin, E., HilleRisLambers, J.: Seed dispersal<br />

near and far: Patterns across temperate and tropical <strong>for</strong>ests. Ecology<br />

80(5), 1475–1494 (1999)


20 Baeumer, Kovács, Meerschaert<br />

17. Clark,J. S., Lewis, M., Horvath, L.: Invasion by Extremes: Population Spread<br />

with Variation in Dispersal and Reproduction The American Naturalist 157(5),<br />

537–554 (2001)<br />

18. Cliff, M., Goldstein, J.A., Wacker, M.: Positivity, Trotter products, and blowup.<br />

Positivity 8(2), 187–208 (2004)<br />

19. del Castillo-Negrete, D., Carreras, B.A., Lynch, V.E.: Front dynamics in<br />

<strong>reaction</strong>-<strong>diffusion</strong> systems with levy flights: A fractional <strong>diffusion</strong> approach.<br />

Physical Review Letters 91(1), 018302 (2003)<br />

20. Engel, K.J., Nagel, R.: One-parameter semigroups <strong>for</strong> linear evolution <strong>equation</strong>s,<br />

Graduate Texts in Mathematics, vol. 194. Springer-Verlag, New York<br />

(2000). With contributions by S. Brendle, M. Campiti, T. Hahn, G. Metafune,<br />

G. Nickel, D. Pallara, C. Perazzoli, A. Rhandi, S. Romanelli and R. Schnaubelt<br />

21. Feller, W.: An Introduction to Probability Theory and Applications. Volumes<br />

I and II. John Wiley and Sons (1966)<br />

22. Gorenflo, R., Mainardi, F., Scalas, E., Raberto, M.: Problems with using existing<br />

transport models to describe microbial transport in porous media. In:<br />

Trends Math., pp. 171–180. American Society <strong>for</strong> Microbiology, Mathematical<br />

finance (Konstanz, 2000) (2001)<br />

23. Heyde, C.C., Leonenko, N.N.: Student Processes. Advances in Applied Probability<br />

37, 342–365 (2005)<br />

24. Hille, E., Phillips, R.S.: Functional analysis and semi-groups. American Mathematical<br />

Society, Providence, R. I. (1974). Third printing of the revised edition<br />

of 1957, American Mathematical Society Colloquium Publications, Vol. XXXI<br />

25. Jacob, N.: Pseudo-differential operators and Markov processes, Mathematical<br />

Research, vol. 94. Akademie Verlag, Berlin (1996)<br />

26. Katul, G.G., Porporato, A., Nathan, R., Siqueira, M., Soons, M.B., Poggi, D.,<br />

Horn, H.S., Levin, S.A.: Mechanistic analytical models <strong>for</strong> long-distance seed<br />

dispersal by wind. The American Naturalist 166, 368–381 (2005)<br />

27. Klafter, J., Blumen, A., Shlesinger, M.F.: Stochastic pathways to anomalous<br />

<strong>diffusion</strong>. Phys. Rev. A 35, 3081–3085 (1987)<br />

28. Klein, E.K., Lavigne, C., Picault, H., Renard, M., Gouyon, P.H.: Pollen dispersal<br />

of oilseed rape: estimation of the dispersal function and effects of field<br />

dimension. Journal of Applied Ecology 43(10), 141–151 (2006)<br />

29. Kozubowski, T., Meerschaert, M., Podgórski, K: <strong>Fractional</strong> Laplace Motion.<br />

Advances in Applied Probability 38(2), to appear (2006) http://www.maths.<br />

otago.ac.nz/%7Emcubed/fLaplace.pdf<br />

30. Lockwood, D.R., Hastings, A.: The effects of dispersal patterns on marine<br />

reserves: Does the tail wag the dog Theoretical Population Biology 61, 297–<br />

309 (2002)<br />

31. Meerschaert, M., Mortensen, J., Wheatcraft, S.: <strong>Fractional</strong> vector calculus <strong>for</strong><br />

fractional advection-dispersion. Physica A p. to appear (2006)<br />

32. Meerschaert, M.M., Benson, D.A., Baeumer, B.: Multidimensional advection<br />

and fractional dispersion. Phys. Rev. E 59, 5026–5028 (1999)<br />

33. Meerschaert, M.M., Benson, D.A., Baeumer, B.: Operator Lévy motion and<br />

multiscaling anomalous <strong>diffusion</strong>. Phys. Rev. E 63, 1112–1117 (2001)<br />

34. Meerschaert, M.M., Benson, D.A., Scheffler, H., Baeumer, B.: Stochastic solution<br />

of space-time fractional <strong>diffusion</strong> <strong>equation</strong>s. Phys. Rev. E 65, 1103–1106<br />

(2002)<br />

35. Meerschaert, M.M., Benson, D.A., Scheffler, H.P., Becker-Kern, P.: Governing<br />

<strong>equation</strong>s and solutions of anomalous random walk limits. Phys. Rev. E 66,<br />

102R–105R (2002)<br />

36. Meerschaert, M.M., Scheffler, H.P.: Limit Distributions <strong>for</strong> Sums of Independent<br />

Random Vectors: Heavy Tails in Theory and Practice. Wiley Interscience,<br />

New York (2001)<br />

37. Miller, K., Ross, B.: An Introduction to the <strong>Fractional</strong> Calculus and <strong>Fractional</strong><br />

Differential Equations. Wiley and Sons, New York (1993)<br />

38. Miyadera, I., Ôharu, S.: Approximation of semi-groups of nonlinear operators.<br />

Tôhoku Math. J. (2) 22, 24–47 (1970)<br />

39. Morales, J., Carlo, T.: The effects of plant distribution and frugivore density<br />

on the scale and shape of dispersal kernels. Ecology, to appear (2005)


<strong>Fractional</strong> <strong>reaction</strong>-<strong>diffusion</strong> <strong>equation</strong> 21<br />

40. Murray, J.D.: Mathematical biology. I,II, Interdisciplinary Applied Mathematics,<br />

vol. 17,18, third edn. Springer-Verlag, New York (2002)<br />

41. Neubert, M., Caswell, H.: Demography and dispersal: Calculation and sensitivity<br />

analysis of invasion speed <strong>for</strong> structured populations. Ecology 81(6),<br />

1613–1628 (2000)<br />

42. Nolan, J.P.: Numerical calculation of stable densities and distribution functions.<br />

Heavy tails and highly volatile phenomena. Comm. Statist. Stochastic<br />

Models 13(4), 759–774 (1997).<br />

43. Paradis, E., Baillie, S.R., Sutherland, W.J.: Modeling large-scale dispersal distances.<br />

Ecological Modelling 151(2-3), 279–292 (2002)<br />

44. Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential<br />

<strong>equation</strong>s, Applied Mathematical Sciences, vol. 44. Springer-Verlag, New<br />

York (1983)<br />

45. Raberto, M., Scalas, E., Mainardi, F.: Waiting-times and returns in highfrequency<br />

financial data: an empirical study. Physica A 314, 749–755 (2002)<br />

46. Sabatelli, L., Keating, S., Dudley, J., Richmond, P.: Waiting time distributions<br />

in financial markets. Eur. Phys. J. B 27, 273–275 (2002)<br />

47. Saichev, A.I., Zaslavsky, G.M.: <strong>Fractional</strong> kinetic <strong>equation</strong>s: solutions and applications.<br />

Chaos 7(4), 753–764 (1997)<br />

48. Samko, S., Kilbas, A., Marichev, O.: <strong>Fractional</strong> Integrals and derivatives: Theory<br />

and Applications. Gordon and Breach, London (1993)<br />

49. Samorodnitsky, G., Taqqu, M.S.: Stable Non-Gaussian Random Processes.<br />

Chapman & Hall/CRC (1994)<br />

50. Sato, K.: Lévy Processes and Infinitely Divisible Distributions. Cambridge<br />

studies in advanced mathematics. Cambridge University Press (1999)<br />

51. Scalas, E., Gorenflo, R., Mainardi, F.: <strong>Fractional</strong> calculus and continuous-time<br />

finance. Phys. A 284, 376–384 (2000)<br />

52. Scalas, E.: The application of continuous-time random walks in finance and<br />

economics. Physica A 362 225–239 (2006)<br />

53. Scalas, E., Gorenflo, R., Luckock, H., Mainardi, F., Mantelli, M., Raberto, M.:<br />

On the Intertrade Waiting-time Distribution. Finance Letters 3, 38–43 (2005)<br />

54. Schilling, R.L.: Growth and Hölder conditions <strong>for</strong> sample paths of Feller proceses.<br />

Probability Theory and Related Fields 112, 565–611 (1998)<br />

55. Schumer, R., Benson, D.A., Meerschaert, M.M., Baeumer, B.: Multiscaling<br />

fractional advection-dispersion <strong>equation</strong>s and their solutions. Water Resources<br />

Research 39, 1022–1032 (2003)<br />

56. Schumer, R., Benson, D.A., Meerschaert, M.M., Wheatcraft, S.W.: Eulerian<br />

derivation of the fractional advection-dispersion <strong>equation</strong>. Journal of Contaminant<br />

Hydrology 48, 69–88 (2001)<br />

57. Sokolov, I.M., Klafter, J.: From <strong>diffusion</strong> to anomalous <strong>diffusion</strong>: a century<br />

after Einstein’s Brownian motion. Chaos 15(2), 026,103, 7 (2005)<br />

58. Soons, M.B., Heil, G.W., Nathan, R., Katul, G.G.: Determinants of longdistance<br />

seed dispersal by wind in grasslands. Ecology 85(11), 3056–3068<br />

(2004)<br />

59. Tackenberg, O.: Modeling long-distance dispersal of plant diaspores by wind.<br />

Ecological Monographs 73(2), 173–189 (2003)<br />

60. Tackenberg, O., Poschlod, P., Kahmen, S.: Dandelion seed dispersal: The horizontal<br />

wind speed does not matter <strong>for</strong> long-distance dispersal - it is updraft!<br />

Plant Biology 5(5), 451–454 (2003)<br />

61. Zaslavsky, G.: <strong>Fractional</strong> kinetic <strong>equation</strong> <strong>for</strong> hamiltonian chaos. chaotic advection,<br />

tracer dynamics and turbulent dispersion. Phys. D 76, 110–122 (1994)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!