You are on page 1of 21

SIAM J. APPL. MATH.

c 1999 Society for Industrial and Applied Mathematics



Vol. 60, No. 2, pp. 371391

ON TRAVELING WAVE SOLUTIONS OF FISHERS EQUATION IN


TWO SPATIAL DIMENSIONS
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

PAVEL K. BRAZHNIK AND JOHN J. TYSON

Abstract. It is shown for the quadratic Fisher equation in two spatial dimensions that, along
with a plane wave, there exist several other traveling waves with nontrivial front geometry. Some
of the solutions are found in explicit form; others are constructed approximately. The dispersion
relationship and velocity-curvature dependence generated by these solutions are studied.

Key words. Fisher equation, bistable medium, traveling waves, wave patterns, reaction-
diffusion equation

AMS subject classification. 35K57

PII. S0036139997325497

1. Introduction. The scalar nonlinear reaction-diffusion equation for u(t, x)


with quadratic kinetics

u 2u
(1.1) = D 2 + su (1 u)
t x
has a long-standing history in mathematical modeling of propagation phenomena in
distributed dissipative systems. Sixty years ago Fisher proposed it as a model for the
propagation of a mutant gene [1]. The same equation occurs in logistic population
growth models [2], flame propagation [3], neurophysiology [4], autocatalytic chemical
reactions [5], branching Brownian motion processes [6], and nuclear reactor theory
[7]. It is incorporated as an important constituent of nonscalar models describing ex-
citable media (EM), e.g., the BelousovZhabotinsky (BZ) reaction [8]. In chemical
media the function u (t, x) is the concentration of the reactant, D represents its diffu-
sion coefficient, and the positive constant s specifies the rate of the chemical reaction.
In media of other natures, u might be temperature or electric potential, while D might
be the thermal conductivity or specific electrical conductivity. The medium described
by (1.1) is often referred to as a bistable medium because it has two homogeneous sta-
tionary states, u = 0 and u = 1. A kink-like traveling wave solution of (1.1) describes
a constant-velocity front of transition from one homogeneous state to another.
In one-dimensional space, (1.1) has received considerable attention (see, e.g.,
[9, 10]). On the other hand, many realistic systems are essentially two- or three-
dimensional and surprisingly little work has been done in this direction so far [11, 12].
In particular, possible traveling wave solutions have not yet been characterized.
Finding solutions of nonlinear models is a difficult and challenging task. Several
analytical methods have been developed for obtaining wave solutions for pure disper-
sive nonlinear systems in one spatial dimension: the inverse scattering transfer [13],
the Hirota method [14], Lambs ansatz [15], etc. [16]. Some of these methods may be
extended for (2+1)-dimensional systems (two spatial dimensions and one temporal

Receivedby the editors August 4, 1997; accepted for publication (in revised form) October 27,
1998; published electronically December 28, 1999. This work was supported by NSF grant CHE
95-00763.
http://www.siam.org/journals/siap/60-2/32549.html
Department of Biology, Virginia Polytechnic Institute and State University, Blacksburg, VA

24061 (brazhnik@vt.edu, tyson@vt.edu).


371
372 PAVEL K. BRAZHNIK AND JOHN J. TYSON
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

Fig. 1.1. Shapes of stationary propagating excitation wave fronts in two-dimensional EM pre-
dicted by the kinematic model [23]. Patterns can be characterized by their velocity of propagation:
two patterns propagate with the velocity of a plane wave: the plane wave and the separatrix front;
the V-pattern and the Y-wave move with speeds which depend on the asymptotic angle between their
wings 2 and move faster than a plane front; the space-oscillating fronts propagate with velocities
(smaller than plane waves) that depend on the spatial period of the wave.

variable). The problem of obtaining solutions for systems including dissipative losses
turned out to be more complex. Even for the (1+1) case, most of the above-mentioned
methods do not work. The usual way of treating the problem is by perturbation theory
or numerical investigation.
In two and three spatial dimensions, propagating wave fronts may change shape
and evolve to some stationary configuration (wave pattern) which is not necessarily
unique. Such stationary structures usually possess some kind of symmetry. For a two-
dimensional bistable medium, leaving one space dimension for the wave propagation
direction, the only symmetry we may expect is reflection symmetry of the wave front
with respect to its propagation direction.1 Patterns with this property are known for
nonlinear wave models of a different nature, e.g., for two-dimensional Kortevegde
Vries (KdV) waves [17],2 or for the sine-Gordon equation [18, 19]. The patterns are
building blocks for constructing possible stationary wave configurations in a layered
medium, e.g., stationary refracting waves [20]. For a bistable medium, some insight
into what patterns to expect can be gained from a lower level, geometrical model
which treats the wave front as a curved line that moves according to its curvature.
Such a kinematic model [21] developed in the context of EM3 provides a set of

1 Spiral-like stationary circulating solutions are not possible in a bistable medium because its

elements do not return to their initial state after the passage of the wave.
2 An extension of the KdV equation for two dimensions is known as the KadomtsevPetviashvili

equation [16].
3 An excitable medium is similar to a bistable medium, except the excited state is metastable:

after some time the system leaves the excited state and returns to the single stable resting state.
TRAVELING WAVES OF FISHERS EQUATION 373

noncirculating solutions [22, 23] whose front lines are depicted in Figure 1.1. Each
pattern passes through the medium only once, so it does not matter whether or not
the medium regains excitability.
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

In this paper we derive analytical representations of several traveling wave solu-


tions for the Fisher equation in two dimensions. Our work relies significantly on results
for the one-dimensional Fisher equation, which we summarize briefly in section 2. In
section 3 we construct solutions for two space dimensions, and in section 4 we analyze
the velocity-curvature dependence for these solutions. Section 5 contains conclusions
and discussion.
2. Solutions for the quadratic Fisher equation in one dimension. After
1/2
rescaling time t0 = st and space x0 = (s/D) x, and dropping the prime, (1.1)
becomes

(2.1) ut = uxx + u(1 u).

A traveling wave solution u(x, t) = u( = x ct), propagating with a speed c, then


obeys the ordinary differential equation (ODE)

(2.2) cu = u + u(1 u).

Phase plane analysis can be used to characterize solutions of (2.2). For most ap-
plications, u () is restricted to be positive and bounded. Therefore the boundary
conditions for the traveling wave solution are usually

(2.3) u ( ) 1 while u ( ) 0.

The speed of the waves then has to be found as the solution for the eigenvalue problem
(2.2), (2.3).
Fisher found that (2.1) has an infinite number of traveling wave solutions for
which 0 u (x, 0) 1 and wave speeds are c cmin = 2. Kolmogorov, Petrovsky,
and Piskunov [24] proved that for bounded u(x, 0), if u (x, 0) = 1 for x < a, and
u (x, 0) = 0 for x > b, there is a unique solution of (2.1) and that this solution evolves
into a traveling monotonic wave solution with a speed c = cmin . Higher velocity waves
arise when an appropriate initial gradient is present in the system: the less steep the
wave profile is, the faster it moves [10]. McKean [25] showed that any wave speed
c > 2 is stable (with respect to small perturbations) if the initial datum has the right
behavior at the tails. A traveling wave solution for (2.1) in explicit form was found
by Ablowitz and Zeppetella [26] for one special case:

1
(2.4) u(x, t) =  2 .
1 + exp / 6

The kink (2.4) propagates from left to right with a speed c = 5/ 6 2.041. The
problem of selection of appropriate speed has been discussed in [10, 27, 28]. Fronts
initiated on a compact support evolve to minimum velocity solutions [27]. Traveling
wave solutions for (2.2) with c < 2 also exist but they are considered to be physically
unrealistic since u becomes negative for some : u 0 at the leading edge with
decreasing oscillations about u = 0.
One approach for finding traveling wave solutions for bistable medium, introduced
by Rinzel and Keller (RK) [29], is to replace (2.1) by a piecewise-linear approximation
374 PAVEL K. BRAZHNIK AND JOHN J. TYSON

that retains the essential features of the reaction term: two roots (at u = 0 and u = 1)
connected by a continuous function with its unique maxima located between the roots.
For instance, we might seek a solution of two independent linear equations
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

1
(2.5) ut = uxx + u for u ,
2

1
(2.6) ut = uxx + 1 u for u
2
satisfying appropriate boundary conditions and continuously joined at the point where
u = 1/2. Such a solution is
    
1 1   1   1
(2.7) u () = exp + + + exp for u ,
2 2 2 2
1   1
(2.8) u () = 1 exp + for u ,
2 2

2
where = c c2 4 /2, and the negative constant, = 2c+ c +4 , ap-
2 c2 4
proaches 1/2 when c (the contribution of the second term in (2.7) becomes
negligibly small for high-velocity waves). The expression (2.7) is not defined for c = 2
because in this case . Thus the velocity of the traveling waves is notuniquely
defined but only restricted to the interval c (cmin = 2, ). For c = 5/ 6, (2.7),
(2.8) can be recasted as follows:
       
2 1 3 1
(2.9) u () = exp exp for u ,
6 2 6 2
  
1 1 1
(2.10) u () = 1 exp for u ,
2 6 2
which coincides with corresponding asymptotes of the exact solution (2.4) except for
the numerical coefficients at the second terms which represent a next-order correction.
The velocity of the Fisher traveling waves may depend on the wave amplitude.
This can be seen by considering the equation

(2.11) ut = uxx + au u2 .

For traveling waves(2.11) converts in terms of new amplitudeu = u/a to (2.2) with
the space variable a and velocity multiplied by the factor a. By making use of
the specific solution (2.4) it gives a kink
a
(2.12) u(x, t) =  p a 2
1 + exp 6
p
that propagates with speed V = 5 a6 . In this case solutions with negative amplitudes
may become meaningful in the context of their applications to EM (nonscalar) models.
For instance, the simplest models for waves in BZ reaction or neuromuscular tissue
involve only two components (excitation u and recovery v), and the equation for u is
of the form [8, 30]

(2.13) ut = uxx + u u2 v.
TRAVELING WAVES OF FISHERS EQUATION 375
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(a) (b)
Fig. 2.1. Phase plane for one-dimensional quadratic Fisher equation in a stationary regime
(original (a) and piecewise-continuous version (b); dashed line is the line of concatenation).

Here v can be considered in its first approximation


 as being a small constant. Then
(2.13), after shifting u, u = u + 1 1 4v /2, converts into equation (2.11) for

u with a = 1 4v.
Equation (2.1) also admits a one-parameter family of nontrivial standing-wave
solutions, i.e., structures which do not evolve in time (ut = 0). The phase plane cor-
responding to a stationary version of (2.1) is depicted in Figure 2.1(a) and exhibits
the existence of bounded oscillating and separatrix solutions and two families of un-
bounded solutions. The explicit solutions can be obtained in terms of the Jacobian
elliptic functions [31] which, for some specific cases, reduce to elementary functions:
Oscillating solutions can be presented as
r !
2 1 2
(2.14) u (x) = c + (b c) sn x, k ,
g 3
h p i
where b and c = 0.5 (b 1.5) (b 1.5)2 4b(b 1.5) are maximum and min-
imum amplitudes of u, respectively, sn(y, k)is the Jacobian elliptic sine, k 2 = (b c) /
(1.5 b 2c) is its modulus, and g = 2/ 1.5 b 2c; since the real period of the
function sn2 is 2K, K being the complete elliptic integral of the first kind, the wave-
length of u is given by

(2.15) = 6gK.

The loop of the separatrix is formed by the solution following from (2.14) as
a specific case (b = 1, c = 1/2, k = 1) :
3/2
(2.16) u(x) = 1 2,
[cosh (x/2)]
3/2
while the wings of the separatrix correspond to u(x) = 1 + [sinh(x/2)]2
.
376 PAVEL K. BRAZHNIK AND JOHN J. TYSON

The counterpart of the trivial solution (u (x) = 0) is an unbounded solution


of the form
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

3/2
(2.17) u(x) =  2 .
cos x2

It is useful to compare here the original phase plane in Figure 2.1(a) with the one
produced by a corresponding RK piecewise-linear approximation, Figure 2.1(b). For
u 1/2, circle-like orbits are formed by the solutions

(2.18) u (x) = c cos (x)

coinciding with the small b approximation (c b) for (2.14). The branches of the
orbit for u 1/2 correspond to the following solutions:
(1) Curves lying inside the separatrix, for 1/2 u 1, are given by

(2.19) u (x) = 1 A cosh (x) ,

while curves lying to the right of the separatrix, for u 1, are given by

(2.20) u (x) = 1 + A cosh (x) ;

(2) the pieces of the separatrix, for 1/2 < u 1, are given by

(2.21) u (x) = 1 exp (x) ,

while its wings (u 1) are given by

(2.22) u (x) = 1 + exp (x) ;

(3) unbounded curves lying outside the separatrix, which continue the circle-like
orbits on the u < 1/2 half-plane, are given by

(2.23) u (x) = 1 B sinh (x) ,

where A and B are positive constants parametrizing different solu-


tions within each family. Comparing (2.14)(2.23) one may notice
that inside the separatrix loops the nonlinear Fisher equation and
piecewise-linear versions give qualitatively similar results. Outside,
where Fishers equation predicts the unbounded oscillating solution
(2.17), the piecewise-linear version gives the unbounded but mono-
tonically increasing solution (2.20).
3. Two-dimensional stationary structures. In a two-dimensional isotropic
medium Fishers equation in rescaled variables is

(3.1) ut = uxx + uyy + u(1 u).

We assume further that a stationary wave moves along the X-axis. The velocity of
the wave, Vp , may be, generally speaking, different from that for the plane wave, c.
For traveling wave solutions, then, u(x, y, t) = u( = x Vp t, y), and u(, y) satisfies
the equation

(3.2) Vp u = u + uyy + u(1 u).


TRAVELING WAVES OF FISHERS EQUATION 377

Here we do not restrict u to be positive but just require the absence of motion as
. Hence, far in front and far behind the transition region, u (, y) must
approach asymptotic states u ( , y) u (y) smoothly; that is,
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

   2 
u u
(3.3) = = 0.
2

Note that the asymptotic states may depend nontrivially on y because they obey the
2
one-dimensional stationary Fisher equation: ddyu2 + u (1 u ) = 0. This equation,
as we mentioned, admits inhomogeneous solutions (2.14)(2.17).
The trivial generalization of one-dimensional solutions for the two-dimensional
case is a plane wave propagating with velocity Vp = c along the X-axis: u does
not depend on y, and the front line associated with, say, the line of a constant level
(u = const) is a straight line perpendicular to the X-axis. A similar generalization
admits traveling wave solutions with damped oscillations in front of the leading edge.
The simplest solutions involving both x and y space variables can be constructed from
a plane wave by rotating the frame of reference.4 For example,

(3.4)
  2
x sin y cos 5 t   2
6 sin y cos
u(x, t) = 1 + exp = 1 + exp
6 6

corresponds to the specific solution (2.4) propagating at an angle from the X-


axis. Equivalently, the wave can be thought of as a tilted plane wave propagating
along the X-axis with velocity Vp = c/ sin ( = x Vp t). Are any other (nontrivial)
two-dimensional traveling wave front configurations possible?
3.1. Linear approximation. In order to get an idea of what to expect, let us
first consider the linearized version of (3.2):

(3.5) Vp u = u + uyy + u.

This diffusion-like equation approximates (3.2) at small u. It is known that (3.5),


with uyy 0, admits a traveling wave solution but its amplitude is not bounded.
Hence, the solutions we may find for (3.5) also will be unbounded, but at this point
we are interested in possible spatial configurations of fronts rather than in their am-
plitudes. Moreover, in the RK approach, (3.5) serves only as the equation for the
bottom of the wave (u 1/2); the top of the wave (u 1/2) is described by a
different equation which is responsible for the wave amplitude to be bounded.
Solutions for the linear equation (3.5) are separable. In (3.5) substituting u of the
form

(3.6) u (, y) = K()Y (y)

yields
00 0
(3.7) K + Vp K + ( + 1) K() = 0,

4 Note that this simple idea does not necessarily work for any nonlinear wave model. It works for

the sine-Gordon equation but fails when applied to certain two-dimensional extensions of KdV or
Burgers equations.
378 PAVEL K. BRAZHNIK AND JOHN J. TYSON

(3.8) Y 00 Y (y) = 0,
where is an unknown separation constant. General solutions for (3.7), (3.8) are
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

   
(3.9) K() = K1 exp + () + K2 exp () ,
h i h i
(3.10) Y (y) = Y1 exp y + Y2 exp y if 6= 0,
(3.11) Y (y) = Y3 y + Y4 if = 0.
h q i
Here () = 0.5 Vp Vp2 4 ( + 1) are roots of the characteristic polyno-
mial, and K1,2 and Y1,...,4 are integration constants. Notice that () = Vp
+ (). Also, if we choose Y1 6= 0 and Y2 = 0 (or vice versa), we find the leading edge
of a plane wave inclined to the X-axis (e.g., (3.4) for ). If both Y1 and Y2 are
nonzero, then symmetry requires that either Y1 = Y2 (i.e., Y (y) even) or Y1 = Y2
(i.e., Y (y) odd) and the absolute value of Y (y) must be taken. Clearly, the solutions
we can construct for u (, y) now crucially depend on the choice of the constants K1,2
and Y1,...,4 and .
For = 0 and Y3 = 0, u does not depend on y and is presented by an expression
of the form (2.9); hence, it corresponds to a plane wave propagating along the X-axis
with velocity Vp = c. (Note that (0) of this section is identical to of section
2.) This solution approximates the bottom part of the traveling wave connecting
homogeneous asymptotic states u ( , y) = 1 and u ( , y) = 0. For the
specific case c = 5/ 6 the coefficients K1,2 can be identified by comparing (3.9) with
the expansion of (2.4) at .
If both Y3 and Y4 are not equal to zero but still = 0, we get the solution of the
form
    
(3.12) u (, y) = K11 exp + (0) + K22 exp (0) (|y| + Y4 /Y3 )

(here K11,22 are constants). The lines of constant level for the above solution (u =
ul const) are depicted in Figure 3.1(a). For positive ul they straighten out when
y and become orthogonal to the X-axis, and, therefore, they propagate with
the velocity of a plane wave (Vp = c). Since the wave is stationary (each of its points
propagates along the X-axis with the same speed), we conclude that solution (3.12)
describes a pattern of nontrivial geometry propagating with the same velocity as a
plane wave. The configuration of the pattern is shown in Figure 3.1(b). The solution
near u = 0 (3.12) presents two pieces of wings of the traveling wave connecting the
asymptotic (inhomogeneous) state at given by (2.16) and the homogeneous
p state at . Therefore, the value for the constant Y4 /Y3 , namely,
zero asymptotic
2 arccosh ( 3/2), can be determined from the comparison of (3.12) with the ex-
pansion of (2.16) for small u. The solution is not valid along the line y = 0, where
the dependence of the appropriate solution on y must be of at least the second order
(the line is the line of the local extremum) and therefore the visual discontinuity of
the wave-front-surface is artificial. In the kinematic theory this pattern corresponds
to the separatrix solution (see Figure 1.1).
2
Positive lead to the restriction (Vp /2) > + 1, which means that possible
waves can propagate with velocity faster than a plane wave. Y (y) becomes this time
either a hyperbolic cosine-function or the absolute value of a hyperbolic sine-function.
If Y (y) is expressed through the hyperbolic cosine, the solution has the form
      
(3.13) u (, y) = M11 exp + () + M22 exp () cosh y
TRAVELING WAVES OF FISHERS EQUATION 379
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

(a) (b)
Fig. 3.1. The separatrix solution for the linearized version of the two-dimensional Fisher
equation (here and in all figures below, waves propagate from left to right): (a) lines of constant
levels: ul = 0.05, 0.01, and 0.1; (b) wave front.

y
u

x
Fig. 3.2. The V-wave solution for the linear approximation.

(here M11,22 are some phase constants) and is depicted in Figure 3.2. It is clearly
reminiscent of the V-pattern known in EM [32]: the front consists of two extended,
almost flat wings colliding at a certain asymptotic angle; in the point of collision
the wings are
connected with each other by a smooth, highly curved, relatively short
area ( 1/ ) where the front line turns rapidly. Like the plane wave, a V-pattern
is supposed to connect homogeneous asymptotic states. For this solution we can find
the connection of the separation constant to the asymptotic angle between each
wing and the X-axis. The angle can be evaluated from (3.13) by using an expression
u
for a level line (u (, y) = ul const): dy
d = ( uy ),y = tan ( ). Supplementing
380 PAVEL K. BRAZHNIK AND JOHN J. TYSON
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

(a) (b)
Fig. 3.3. The Y-wave front (from the linear approximation): (a) lines of constant levels:
ul = 0.01, 0.01, 0.1, and 0.3; (b) wave front.

it with the relationship Vp = c/sin( ) , we get


 
2
(3.14) = + cos ( ) , = cos ( ) for c = 5/ 6 .
6
2
Thus, for positive , must be less than [ + ] , the quantity equal to unity for
c = cmin and decreasing to zero as c . Note that an expression identical to (3.14)
also follows just from a formal comparison of the asymptotic expansions of the exact
solution (3.4) and the expression (3.13) (this also provides values for the constants
M11,22 ). Equation (3.14) allows us to establish a relationship between Vp and ,
c
(3.15) Vp = q ,
2
1 / [ + ]

and to further express () either as functions of or as Vp ,


 
+ + + c
(3.16) () = sin ( ) = .
Vp

Substituting (3.16) into (3.9), we can rewrite K (; ) in terms of Vp (for any pattern)
or in terms of (for V-patterns). Note that the expression (3.15) is exact and
pattern-independent. It is obviously true for plane and separatrix waves ( = 0) but
also must hold for the Y-wave and space-oscillating solutions discussed below.
The expression constructed from the solution with Y (y) proportional to the ab-
solute value of the hyperbolic sine leads to a Y-wave (see Figure 3.3), a traveling
wave connecting asymptotic states just as for the separatrix solution. In fact, a Y-
wave differs from the separatrix solution only by the angle between asymptotes 2 :
for the former it can be any positive value between zero and , while for the latter it
is exactly . The connection between and this time is identical to (3.14).
TRAVELING WAVES OF FISHERS EQUATION 381
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

(a) (b)
Fig. 3.4. The space-oscillating traveling waves (from the linear approximation): (a) lines of
constant levels: ul = 0.01, 0.1, and 0.2; (b) wave front.

For negative the function Y (y) turns into the trigonometric cosine, leading
therefore to the front oscillating in space:
      
(3.17) u (, y) = K11 exp + () + K22 exp () cos y

(see Figure 3.4), where K11,22 are undetermined constants. The surface corresponds
to the leading edge (near u = 0) of the space-oscillating traveling wave connecting the
oscillating asymptotic state (2.14) at and the homogeneous zero asymptotic
state at .
The wavelength of the front oscillations is defined by the value
of = 2/ and the only connection imposed on Vp and at this level of
approximation is Vp2 > 4 ( + 1) , which just means that oscillating fronts may travel
with a velocity smaller than cmin . But if we adopt the result received from the analysis
of V-wave solution, (3.15), then the dispersion relationship becomes (see Figure 3.5)
c
(3.18) Vp = q ,
2
1 + (2/ + )

showing that the smaller the wavelength, the slower the speed. When the front
turns into a plane wave and Vp c. For large c, + 0, and therefore the dispersion
curve turns into a linear function. Equation (3.18) does not show any restrictions on
, since Vp 0 as 0. On the other hand, in (3.17), the second term has to
be a next-order correction to the first term, which requires () < + (). This is
equivalent to > 2 (for c > 2), which gives a minimal period for the space-oscillating
front.
Thus at the level of the linear approximation we have reconstructed all patterns
predicted by the kinematic model. The temptation arises then to proceed further
with the RK algorithm and to construct tops (bottoms, where necessary, as in
Figure 3.1(b)) of patterns. However, in two dimensions, this approach does not work.
Solutions of the (2+1) linear equation analogous to (2.6) do not satisfy conditions nec-
essary for smooth concatenation with the solutions we have just found. For instance,
382 PAVEL K. BRAZHNIK AND JOHN J. TYSON
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

Fig. 3.5. The dispersion curves for c = cmin : upper curve corresponds to the linear approxi-
mation solution, (3.18); its part for < 2 is depicted by the dashed line; lower curve follows from
the approximate space-oscillating solution, (3.25).

the solution for the top of the V-wave has the form
 
(3.19) u (, y) = 1 u00 exp + () cosh ( y)

(u00 and are constants) and fails to satisfy the asymptotic boundary conditions
u ( , y) 1, u (, y ) 1 simultaneously.
3.2. Beyond the linear approximation. Now we try to explore what kind of
traveling wave solutions can be constructed for (3.2) beyond the linear approximation.
Let us substitute (3.6) into the original nonlinear equation (3.2):

(3.20) Vp K Y = K Y + KYyy + KY K 2 Y 2 .

Comparing phase portraits in Figures 2.1(a) and 2.1(b) one can see that the oscillating
asymptotic state for small amplitudes (u < 0.5) is well reproduced by a linearized
version of the equation. Therefore, heuristically, we replace Y 2 in the last term of
(3.20) by Y, because this gives us, as we will see below, an oscillating front in y with
proper asymptote behavior as . Such a replacement allows us to separate
variables, and we get two independent equations
00 0
(3.21) K + Vp K + (1 + ) K() K 2 = 0,

(3.22) Y 00 Y (y) = 0,

where is again a separation constant.


The oscillating solution for (3.22),
 
(3.23) Y (y) cos y ,

exists only for negative . Equation (3.21) is simply the one-dimensional Fisher equa-
tion (2.11.11), which provides a kink-like solution propagating with velocity

(3.24) Vp = c 1 + .

Thus, (3.23) and (3.24) lead to the restriction for variations: 1 < 0. For
waves of minimal speed (c = cmin ) and small , (3.24) coincides with the corresponding
TRAVELING WAVES OF FISHERS EQUATION 383
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

Fig. 3.6. The approximate (nonperturbative) space-oscillating solution for the two-dimensional
Fisher equation.


equationp derived in the context of V-waves, (3.15). In the specific case when c = 5/ 6
(Vp = 5 (1 + )/6), the kink takes the form (2.4), with a = 1 + . The space-
oscillating traveling wave solution for the Fisher equation then is

u0 (1 + ) cos y
(3.25) u (, y) =  q 2
1+
1 + exp 6

(see also Figure 3.6), where the constant u0 is restricted to u0 < 0.5 in order to justify
our approximation. Thus both the velocity and the amplitude of the oscillating front
are smaller than corresponding quantities for a plane front. The asymptotic state of
(3.25) at coincides with the linear approximation solution (3.17), but in con-
trast to the latter, this time the velocity of the pattern Vp and its wavelength
in the
transversal direction are connected to each other via (3.24) (since = 2/).5
Hence patterns with shorter wavelength move more slowly and develop smaller am-
plitude. Equation (3.24) also indicates that the wavelength of the oscillating traveling
front has a minimum critical value 2 (when = 1), the result we obtained from
the linear approximation solutions, but this time we see that for minimal both the
velocity of the wave and its amplitude degenerate to zero. At first sight this restriction
may look like an artifact of our approximation since it does not follow directly from
the exact dispersion relationship (3.18). But the following argument speaks in favor of
its existence. The asymptotic state of (3.25) at is (3.23). This is an approxi-
mate expression for the exact asymptotic state given by (2.14). But the latter exhibits
a minimal wavelength = 2 when the amplitude drops to zero (b 0). Accounting
for this restriction in an exact dispersion relationship, we find that, when the ampli-
tude of the wave decreases to zero, the velocity
of the wave decreases, approaching a
finite, nonzero value Vp = c + 2 (= 2 for c = 2).
1+[ ]

5 It is interesting to note that expression (3.24) coincides with an analogous expression for oscil-

lating patterns in EM derived from kinematic theory [22], except for the numerical prefactor at the
second term under the radical: in kinematic theory the term is multiplied by the ratio sD/c2 .
384 PAVEL K. BRAZHNIK AND JOHN J. TYSON
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

Fig. 3.7. The approximate (nonperturbative) separatrix solution for the two-dimensional Fisher
equation.

The separatrix solution propagates with the same velocity as a plane front, Vp = c.
Suppose that for this case K () in (3.20) remains as for a plane wave, that is, one-
dimensional, solution of (2.4). This allows us to eliminate from (3.20) derivatives with
respect to , obtaining

(3.26) KYyy + K 2 Y Y 2 = 0.

For the one-dimensional problem we know that K () and K 2 are restricted to the
interval (0,1), monotonically decreasing functions of , and smoothly approaching
the same limits as . For the specific solution (2.4), they deviate from each
other by at most 25% within the narrow transition region. Therefore replacing K 2
by K in (3.26) seems to be a decent approximation. This leaves us with an ODE for
Y (y) which coincides with the equation for one-dimensional stationary patterns.
For
our purposes, solution (2.16) is appropriate. Thus for the specific case c = 5/ 6 the
separatrix traveling front in two dimensions takes the form
3/2
1 cosh(y/2)2
(3.27) u (, y) =  2 ,
1 + exp / 6

which is depicted in Figure 3.7. The simplification of (3.27) for small u leads to the
expression derived earlier in the linear approximation.
To study V-pattern solutions it is convenient to use a standard perturbation
technique. For this we rewrite (3.2) in terms of a new space variable = /Vp :
 
1
(3.28) u = u + uyy + u(1 u).
Vp2

For a plane wave Vp = c 2 and 1/Vp2 becomes a small parameter. It is known


from the comparison with numerical solutions that, for a plane wave of speed 2,
(3.28), omitting the first term in the right-hand side, gives a traveling wave solution
which deviates only a few percentage points from the numerically computed solution.
For higher speeds the precision of such an approximation is even better. V-patterns
TRAVELING WAVES OF FISHERS EQUATION 385
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

Fig. 3.8. The approximate V-wave solution for the two-dimensional Fisher equation.

propagate with velocity higher than c because of the scissors effect, and therefore
omitting the term 1/Vp2 in (3.28) is even more justified. Without this term, (3.28)
becomes a one-dimensional Fisher equation, where plays the role of time. Taking
into account the symmetry of the V-pattern with respect to the Y -axis, we come to
the conclusion that in a singular perturbation limit the problem of finding V-pattern
solutions is equivalent to the problem of one-dimensional colliding transition waves.
Exact solutions for the latter are unfortunately not known. Therefore we proceed with
further simplification: for the moment assume that we can omit in (3.28) the term uyy .
The first-order differential equation we are left with admits a traveling wave solution
of the form
1
(3.29) u (, y) = ,
1 + A (y) exp (/Vp )

where the function A(y) is unspecified. It can be identified by comparing (3.29) at


with the corresponding solution for small amplitudes (3.13). This gives, for
the V-wave solution, the expression

1
(3.30) u (, y) =   ,
1 + A exp (/Vp ) / cosh y

where A is a phase constant, Vp = c/ sin , and is specified by (3.14). For second


derivatives, (3.30) gives u 1/Vp2 and uyy ,which means that our approximation
(based on uyy negligible) is good for slightly curved V-waves ( /2) and improves
for larger c. The front described by (3.30) is depicted in Figure 3.8.6
In concluding this section, we mention that we were also able to construct several
exact traveling wave solutions for Fishers equation in two dimensions. A triplet of
solutions can be found by substituting into (3.2) the ansatz

1
(3.31) u (x, y, t) = 2
{ (y) + exp [ (x Vp t)]}

6 It is interesting to compare here the expansion of (3.30) for , which gives terms de-

scribing the top of the V-wave, with the expression for the top of the V-wave provided by the RK
algorithm, (3.19).
386 PAVEL K. BRAZHNIK AND JOHN J. TYSON
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

y
u

y y
u u

x x

Fig. 3.9. Exact singular solutions for the two-dimensional quadratic Fisher equation.

and solving the arising ODE for (y) and algebraic equations for and Vp . The
resulting expressions are

3/2
(3.32) u (, y) = 2
[cos (y/2) + exp (/2)]

and
3/2
(3.33) u (, y) = 2.
[|cos (y/2)| exp (/2)]

The traveling wave structures corresponding to (3.32), (3.33) are depicted in Figure
3.9. The waves propagate with the minimum velocity for a plane wave (Vp = 2) and
their amplitudes are not bounded. Wave patterns (3.32), (3.33) connect (though not
continuously) the inhomogeneous asymptotic state (2.17) (at ) with a zero
homogeneous asymptotic state at and show explicitly that such a connection
is not necessarily unique.
4. Velocity-curvature dependence. For systems with dissipative loss, a low-
level model can often be formulated which in the two-dimensional case approximates
the propagating wave front with an evolving front line [18, 33, 34]. The crucial func-
tional parameter referring such geometric models to their microscopical partial dif-
ferential equation (PDE) counterparts is the dependence of the local velocity of the
TRAVELING WAVES OF FISHERS EQUATION 387

wave on the curvature of the front line, k. This dependence is known to be a linear
function (in the first approximation) for crystal growth problems [35, 34] and for EM
(eikonal approximation) [30]. A known extension of the eikonal approximation for
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

EM indicates that V (k) remains almost a linear function for negative k but deviates
from the eikonal approximation significantly for positive k, exhibiting a critical value
beyond which stationary propagation of the curved front is impossible. Experiments
with waves in BZ-based EM confirm such behavior qualitatively. Theoretical deriva-
tions of V -dependence on k are usually based on semiphenomenological arguments or
on a perturbation analysis of corresponding PDE models.
The solutions we have found in section 3 allow us for the first time to construct
nonlinear analytical expressions for the V (k) function from the nontrivial solutions of
a two-dimensional PDE model. We associate a front line with a line of constant level
u (, y) = ul = const. This defines a curve in the plane (, y). The curvature of the
d2 y/d 2
front line is then given by k = 3/2 . Furthermore, the velocity of stationary
[1+(dy/d)2 ]
patterns propagating along the X-axis is related to the local normal velocity of the
front line as V = Vp sin , where the angle between the X-axis and the tangent to
the front line, , can be evaluated from the relationship dy d = tan . Appropriately
combining these three expressions, we may find V (k; ul ).
We start from the linear approximation solutions. Since in (3.9) the second term
is a next-order correction, for simplicity we omit it in our calculations. Then the
separatrix, V-wave (equation (3.13)), and oscillating solutions (equation (3.17)) lead
remarkably to the same expression7
"  2 #  
+ V V
(4.1) k = 1 ,
c c

2
where = (c/Vp ) . For the separatrix solution, Vp = c and = 1. For the V-shaped
2 2
wave, can be expressed via the asymptotic angle: (c/Vp ) = (sin ) . For the
oscillating front the ratio can be related to the period of the space oscillations, , as
2 2
(c/Vp ) = 1 + 2 + . The cubic-like curve (4.1) is depicted in Figure 4.1. The similar
antisymmetric branch for negative V isnot shown in the figure. The dependence (4.1)
exhibits a critical curvature at V = c/ 3. For small curvatures (4.1) becomes
 c 
(4.2) V =c+ k.
2 +
For the wave of minimal speed, (4.2) recovers the eikonal approximation for EM
(cmin = 2 and + (cmin ) = 1). In deriving (4.2) we also took into account the fact
that small k can be achieved for the V-wave solution only at /2 and for
the oscillating solution when . For large negative curvature, (4.1) gives the
deviation of V from the linear approximation (4.2) towards smaller V .
Equation (4.1) is constructed for small u solutions and hence is valid for low-level
cuts, when ul is small, but it turns out to be independent on the value of ul . Accounting
for both terms in (3.9) will introduce a dependence on ul into k (V ) expressions, which
therefore become more complicated, but the separatrix solution for c = cmin recovers
(4.1) exactly.

7 Here, to eliminate we used the relationship between V and derived from the consideration
p
of the V-wave solution in (3.15).
388 PAVEL K. BRAZHNIK AND JOHN J. TYSON
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

Fig. 4.1. V -of-k dependence from the solutions for the linearized version of the two-dimensional
Fisher equation (solid line) compared to the linear (eikonal) approximation (dashed line).

Solutions we have found beyond the linear approximation can also be analyzed
for V -of-k dependence. For the oscillating (equation (3.25)) and separatrix (equation
(3.27)) solutions the relationship between V and k cannot be expressed explicitly but
can be found in a parametric form (V = V (y) , k = k (y)). For the negligibly small
ul , they reduce to (4.1). The V-wave solution (3.30), on the other hand, provides an
expression different from (4.1) only by the prefactor ( + in (4.1) is replaced by 1/c),
although, according to the construction algorithm, it should be valid only for small k,
while the derivation of (4.1) did not imply any restrictions on k. Surprisingly, exact
solutions (3.32), (3.33) give expressions for k (V ) identical to (4.1).
5. Conclusions. Our investigation shows that Fisher traveling waves in a two-
dimensional bistable medium may have different geometry which, together with a
reaction rate and diffusion coefficient, affects the propagation velocity of waves. Five
bounded solutions have been identified and characterized: plane, V-, and Y-waves,
separatrix, and space-oscillating fronts. At fixed diffusion coefficient D and reaction
rate constant s, the velocity of the wave Vp may be considered as the bifurcation
parameter (see Figure 1.1). The slowest wave is an oscillating front, but its velocity
increases with the increase of the wavelength of the space oscillations. When the
velocity reaches the value corresponding to the plane wave, c, the wave front bifurcates
into two distinct configurations: a plane wave and a separatrix wave. The latter
two give birth, with an increase of Vp , to V- and Y-waves, respectively. The Cauchy
problem for the two-dimensional Fisher equation, answering the question of which
initial conditions lead to which stationary traveling waves, remains to be explored.
Fishers equation is a core for many nonlinear models of different natures. There-
fore, its solutions are reminiscent of patterns known from different fields. V-waves
were characterized in crystal-growth-related models [34] and in EM [32] but only in
the framework of geometrical models. Space-oscillating fronts exhibit similarity with
solidification fingers [36] and may be relevant to cellular flame structures and patterns
arising from diffusion-induced instability of the planar front in chemical reaction-
diffusion systems [12]. The solutions we have constructed may become a good start-
ing point for developing more elaborate PDE-based theories of these patterns. This
TRAVELING WAVES OF FISHERS EQUATION 389

approach for waves in EM is currently underway.


One-dimensional EM may support a train of pulses which propagate with velocity
depending on the spacing between pulses ; this dependence is called the dispersion
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

relation of the EM. Bistable media are not able to support a train of pulses, hence
the concept of dispersion is meaningless. We have shown that traveling waves with
space-oscillating fronts may exist in two-dimensional, bistate media characterized by
Fishers equation. (The front is periodic in the direction orthogonal to its direction of
propagation.) Such waves exhibit a nontrivial dispersion relationship, which relates
the velocity of the space-oscillating stationary propagating front to its wavelength
. This dispersion relation is different from that mentioned in the context of EM
because, in the latter case, dispersion is due to the influence of a second component,
an inhibitor, which is absent in the bistable medium. It was shown recently [23]
with a geometrical model that EM is also capable of supporting space-oscillating
fronts and, hence, may exhibit, in addition to the traditional dispersion relationship,
a second dispersion relationship similar to that found in a bistable medium. Are these
two dispersion relationships independent? What is confusing here is the fact that
the dispersion relationship we have found for the bistable medium is qualitatively
similar to the one found in one-dimensional EM, that is, to what we call here the
traditional dispersion relationship. Moreover, the V -of-k dependence we found here
shows a dependence of the coefficient of k on , similar to that elucidated recently
from experiments with EM [37] for . Are these observations just coincidences or
indications of some common underlying mechanism? This question remains to be
answered.
The V -of-k dependence we found exhibits a critical curvature similar to one for
the EM [30]. But in EM the critical curvature is thought to be attributable to the
presence of an inhibitor in the system. Our study shows, however, that the critical
curvature is present even in the scalar model. In EM a front whose curvature exceeds
the critical curvature breaks at this point and may consequently evolve into spirals.
The scenario for the evolution of supercritical fronts in a bistable medium is not yet
clear.
We showed in section 4 that the problem of finding V-wave solutions for a two-
dimensional equation in a singular perturbation limit is equivalent to the problem of
one-dimensional colliding transition waves. Further, we have constructed an approxi-
mate solution for the V-pattern, which therefore means that we also have constructed
a perturbation solution for two head-on colliding transition waves in one dimension.

REFERENCES

[1] R.A. Fisher, The wave of advance of advantageous genes, Ann. of Eugenics, 7 (1937), pp.
355369.
[2] J.D. Murray, Lectures on Nonlinear-Differential-Equations Models in Biology, Oxford Uni-
versity Press, London, 1977; N.F. Britton, Reaction-Diffusion Equations and Their Ap-
plications to Biology, Academic Press, London, 1986.
[3] D.A. Frank-Kamenetskii, Diffusion and Heat Exchange in Chemical Kinetics, Princeton Uni-
versity Press, Princeton, NJ, 1955; F.A. Williams, Combustion Theory, Addison-Wesley,
Reading, MA, 1965.
[4] H.C. Tuckwell, Introduction to Theoretical Neurobiology, Cambridge Stud. Math. Biol. 8,
Cambridge University Press, Cambridge, UK, 1988.
[5] H. Cohen, Nonlinear diffusion problems, in Studies in Applied Mathematics, MAA Studies
in Math. 7, A.H. Taut, ed., Math. Assoc. Amer. (distributed by Prentice-Hall, Englewood
Cliffs, NJ), 1971, pp. 2764; P.C. Fife and J.B. McLeod, The approach of solutions of
nonlinear diffusion equations to travelling front solutions, Arch. Rational Mech. Anal., 65
390 PAVEL K. BRAZHNIK AND JOHN J. TYSON

(1977), pp. 335361; D.G. Aronson and H.F. Weinberger, Nonlinear diffusion in popu-
lation genetics, combustion and nerve pulse propagation, in Partial Differential Equations
and Related Topics, Lecture Notes in Math. 446, J.A. Goldstein, ed., Springer-Verlag, New
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

York, 1975, pp. 549.


[6] M.D. Bramson, Maximal displacement of branching Brownian motion, Comm. Pure Appl.
Math., 31 (1978), pp. 531581.
[7] J. Canosa, Diffusion in nonlinear multiplicative media, J. Math. Phys., 10 (1969), pp. 1862
1868.
[8] J.J. Tyson and P.C. Fife, Target patterns in a realistic model of the Belousov-Zhabotinskii
reaction, J. Chem. Phys., 73 (1980), pp. 22242257.
[9] P. Fife, Mathematical Aspects of Reacting and Diffusing Systems, Lecture Notes in Biomath.
28, Springer-Verlag, Berlin, 1979.
[10] J.D. Murray, Mathematical Biology, Springer-Verlag, Berlin, 1989.
[11] D.G. Aronson and H.F. Weinberger, Multidimensional nonlinear diffusion arising in pop-
ulation genetics, Adv. in Math., 30 (1978), pp. 3376; C.K.R.T. Jones, Spherically sym-
metric solutions of a reaction-diffusion equation, J. Differential Equations, 49 (1983), pp.
142169; R. Gardner, Multidimensional Travelling Fronts, Lectures in Appl. Math. 23,
AMS, Providence, RI, 1986.
[12] S.K. Scott and K. Showalter, Simple and complex propagating reaction-diffusion front, J.
Chem. Phys., 96 (1992), pp. 87028711; K. Showalter, Quadratic and cubic reaction-
diffusion fronts, Nonlinear Sci. Today, 4 (1995), pp. 110.
[13] S. Novikov, S.V. Manakov, L.P. Pitaevskii, and V.E. Zakharov, Theory of Solitons. The
Inverse Scattering Method, Consultants Bureau, New York, 1984.
[14] R. Hirota, Exact solution of the modified Korteweg-de Vries equation for multiple collisions
of solitons, J. Phys. Soc. Japan, 33 (1972), pp. 14561458.
[15] G.L. Lamb Jr., Analytical description of ultrashort optical pulse propagation in a resonant
medium, Rev. Mod. Phys., 43 (1971), pp. 99124; G. Leibbrandt, Exact solutions of the
elliptic sine equation in two space dimensions with application to the Josephson effect,
Phys. Rev. B, 15 (1976), pp. 33533361.
[16] E. Infeld and G. Rowlands, Nonlinear Waves, Solitons and Chaos, Cambridge University
Press, Cambridge, UK, 1990.
[17] H. Segur, Some open problems, Phys. D, 18 (1986), pp. 111.
[18] B.A. Malomed, Dynamics of quasi-one-dimensional kinks in the two-dimensional sine-Gordon
model, Phys. D, 52 (1991), pp. 157170.
[19] N.K. Vitanov, On travelling waves and double-periodic structures in two-dimensional sine-
Gordon systems, J. Phys. A: Math. Gen., 29 (1996), pp. 51955207.
[20] P.K. Brazhnik and J.J. Tyson, Propagation of waves through a line of discontinuity in two-
dimensional excitable media: Refraction and reflection of autowaves, Phys. Rev. E, 54
(1996), pp. 19581968.
[21] P.K. Brazhnik, V.A. Davydov, and A.S. Mikhailov, Kinematic approach to the description
of autowave processes in active media, Theoret. and Math. Phys., 74 (1988), pp. 300306.
[22] P.K. Brazhnik, Exact solutions for the kinematic model of autowaves in two-dimensional
media, Phys. D, 94 (1996), pp. 205220.
[23] P.K. Brazhnik and J.J. Tyson, Nonspiral excitation waves beyond the eikonal approximation,
Phys. Rev. E, 54 (1996), pp. 43384346.
[24] A.N. Kolmogorov, I.G. Petrovsky, and N.S. Piskunov, Study of the diffusion equation with
growth of the quantity of matter and its application to a biology problem, in Dynamics of
Curved Fronts, P. Pelce, ed., Academic Press, New York, 1988, pp. 105130.
[25] H.P. McKean, Nagumos equation, Adv. Math., 4 (1970), pp. 209223.
[26] M.J. Ablowitz and A. Zeppetella, Explicit solution of Fishers equation for a special wave
speed, Bull. Math. Biol., 41 (1979), pp. 835840.
[27] J.H. Merkin and D.J. Needham, Propagating reaction-diffusion waves in a simple isothermal
quadratic autocatalytic chemical system, J. Engrg. Math., 23 (1989), pp. 343356.
[28] P. Gray, J.H. Merkin, D.J. Needham, and S.K. Scott, The development of travelling waves
in simple isothermal chemical system, Proc. Roy. Soc. London Ser. A, 430 (1990), pp. 509
524.
[29] J. Rinzel and J.B. Keller, Travelling wave solutions of a nerve conduction equation, Biophys.
J., 13 (1973), pp. 13131337.
[30] V.S. Zykov, Simulation of Wave Processes in Excitable Media, A.T. Winfree, ed., Manchester
University Press, Manchester, New York, 1987.
[31] P.F. Byrd and M.D. Friedman, Handbook of Elliptical Integrals for Engineers and Scientists,
Springer-Verlag, Berlin, 1971.
TRAVELING WAVES OF FISHERS EQUATION 391

[32] P.K. Brazhnik and V.A. Davydov, Non-spiral autowave structures in unrestricted excitable
media, Phys. Lett. A, 199 (1995), pp. 4044.
[33] V.A. Davydov, V.S. Zykov, and A.S. Mikhailov, Complex dynamics of spiral waves and
Downloaded 11/15/14 to 130.207.50.37. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

motion of curves, Phys. D, 70 (1994), pp. 139.


[34] D.W. Schwendeman, A front dynamics approach to curvature-dependent flow, SIAM J. Appl.
Math., 56 (1996), pp. 15231538.
[35] W.K. Burton, N. Cabrera, and F.C. Frank, The growth of crystals and the equilibrium
structures of their surfaces, Philos. Trans. Roy. Soc. London Ser. A, 243 (1951), pp. 299
358.
[36] P. Pelce, Dynamics of Curved Fronts, Academic Press, New York, 1988.
[37] A.M. Pertsov, M. Wellner, and J. Jalife, Eikonal relation in highly dispersive excitable
media, Phys. Rev. Lett., 78 (1997), pp. 26562659.

You might also like