Вы находитесь на странице: 1из 25

Physica D 205 (2005) 100124

Long dynamo waves


Joanne Mason
a,b,
, Edgar Knobloch
a,c
a
Department of Applied Mathematics, University of Leeds, Leeds LS2 9JT, UK
b
High Altitude Observatory, National Center for Atmospheric Research,
Boulder, CO 80307, USA
c
Department of Physics, University of California, Berkeley, CA 94720, USA
Available online 5 February 2005
Abstract
A simple mean-eld model of magnetic eld generation in the Sun is considered. The model is characterized by spatially
disjoint locations of the and effects, believed to take place in the solar convection zone and in the solar tachocline, respectively,
and includes -quenching. The model admits a long wave dynamo instability, whose evolution is described by a perturbed mKdV
equation. Solutions of this equation, the so-called snoidal waves, describe nonlinear waves of magnetic activity migrating towards
the equator, as observed in the Sun.
2005 Elsevier B.V. All rights reserved.
Keywords: Mean eld dynamo theory; mKdV equation
1. Introduction
Surface observations of the Suns magnetic activity reveal an array of features varying on many spatial and
temporal scales. One of the most widely studied surface manifestations of the Suns large-scale magnetic eld are
sunspots. Sunspots are regions of intense magnetic eld, and exhibit a 22 year cycle that takes the form of a wave
of magnetic activity. They are born in pairs of opposite polarity, and at the beginning of a cycle appear at about
30

latitude. Their lifetimes are short, however, of the order of several days or weeks, and as they decay new spots
emerge at a slightly lower latitude. The resulting wave of activity reaches the equator in about 11 years, at which
point the process starts afresh, with spot pairs beginning to reappear near 30

, but this time with reversed polarities.


Thus the whole cycle takes approximately 22 years, a period that uctuates somewhat but is in fact remarkably
stable over long times [3].

Corresponding author. Tel.: +1 303 497 1503; fax: +1 303 497 1589.
E-mail address: jmason@hao.ucar.edu (J. Mason).
0167-2789/$ see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.physd.2005.01.006
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 101
Mean eld dynamo theory [20,25] provides a metaphor for understanding the origin of the solar magnetic cycle.
The theory describes the evolution of the large-scale azimuthally-averaged magnetic eld, and relies on two basic
mechanisms: the shearing of a pre-existing poloidal eld by the Suns differential rotation to produce a toroidal
eld (the effect), and a process known as the effect that is responsible for regenerating the poloidal eld from
the toroidal eld. Helioseismology (see, for example, [24,27]) suggests that the most important shearing motion
is radial and occurs at (or just below) the base of the solar convection zone in a region called the tachocline.
In contrast, the classical effect of Parker [21] describes the net effect of convection in a stratied rotating
shell, and hence is distributed throughout the solar convection zone, although the sign of this effect is believed
to be correct for equatorward migration of active regions only towards the lower reaches of this region (see, for
example, [5,32]). Parker [22] has explored a simple model, known as the interface dynamo, that incorporates
the above ideas, and in particular the suggestion that the and effects are physically separated. Nonlinear
versions of Parkers model that include the nonlinear suppression of the effect by the mean magnetic eld have
been explored by (among others) Charbonneau and MacGregor [9] and Tobias [28], showing that the interface
scenario is effective and may generate equipartition strength magnetic elds, and that solutions of the model
partial differential equations can resemble qualitatively a number of the observed properties of the solar magnetic
cycle [6].
Our purpose in this article is not so much to model the solar cycle, but to draw attention to some intriguing
properties of the interface dynamo. In particular, following Mason et al. [19], we note that the equations admit both
long wave modes with small wavenumber, and short wave modes with wavenumber comparable to the depth of the
region responsible for the -effect. Traditionally it is believed that the latter modes are the relevant ones for the solar
dynamo; however, with the development of the interface dynamo it makes sense to consider the long wave modes
as well [29,30]. Somewhat unexpectedly (given the dissipative nature of the model) we nd that the long wave
modes evolve according to the modied Kortewegde Vries equation, and trace this observation to the fact that the
mean-eld dynamo equations are written in terms of the vector potential for the poloidal eld. At leading order this
potential is independent of the depth and is therefore phase-like. Consequently the theory of the long wave mode
bears considerable similarity to phase dynamics, and in particular resembles the development of the celebrated
KuramotoSivashinsky equation as the phase equation for the evolution of the Eckhaus instability [18]. However,
the fact that the sign of the magnetic eld is arbitrary and that the waves have a preferred direction of propagation
(equatorward) changes the form of the phase equation that results, and leads to a modied Kortewegde Vries
equation at leading order, with weak damping and forcing at supercritical dynamo numbers entering only at higher
order.
The paper is organized as follows. In Section 2 we describe the basic problem we study, followed in Section 3 by
a summary of the basic properties of the linear dispersion relation. The bulk of the paper is contained in Section 4
where the leading order amplitude equation, the mKdV equation, is derived. Section 5 contains a derivation of
the perturbed mKdV equation and discusses its solutions; the relation of these solutions to the solar magnetic
activity cycle is summarized in Section 6. Certain aspects of the (somewhat lengthy) derivation are relegated to
Appendix A.
2. The model and governing equations
We consider an idealized nonlinear mean-eld dynamo in which the and effects are spatially separated. For
simplicity we take both of these to be spatially localized, with the former located at z = 1 (representing the effect
of the convection zone) and the latter located at z = 0 (representing the solar tachocline). We write the magnetic
eld B(x, z, t) in the form
B = Ae
y
+Be
y
, (2.1)
102 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
Fig. 1. The geometry of the model. Local Cartesian coordinates are dened on the interface of the convection zone and the tachocline at a point
in the northern hemisphere, with x increasing poleward and z with radius.
where A(x, z, t)e
y
is the vector potential of the poloidal magnetic eld and B(x, z, t) represents the toroidal eld.
The dimensionless dynamo equations [15,20] then read
A
t
= (z)B +
2
A, (2.2)
B
t
= DG(z)
A
x
+
2
B, (2.3)
where (z) = (z 1)/(1 +B
2
), G(z) = (z) and D
0
G
0
z
3
0
/
2
0
is the dynamo number. The formof represents
quenching of the effect as the eld amplies [13], and provides the sole nonlinearity in the problem.
The above equations are to be solved in the semi-innite domain < x < , L z L for waves that
travel inthe negative x direction, i.e., towards the equator (Fig. 1). Here z = L > 1represents the topof the convection
zone, while z = Llies in the radiative interior belowthe tachocline. In the following we adopt boundary conditions
obtained by matching the magnetic eld inside the layer to an external potential eld. In a thin layer geometry such
a procedure leads to the boundary conditions [23]
B(x, z = L, t) = 0,
A
z
(x, z = L, t) = 0. (2.4)
Thus the toroidal magnetic eld is conned in L < z < L while the poloidal magnetic eld is normal to the layer
at z = L at leading order in its aspect ratio. In the following we increase the dynamo number D to trigger the
onset of the dynamo instability.
It should be noted that, except at the locations of the and effects responsible for magnetic eld generation,
the equations for A and B are diffusion equations. Solutions of these equations in the three regions L < z < 0,
0 < z < 1, 1 < z < L satisfying the boundary conditions at z = L are therefore simple to write down. These
then have to be matched across z = 0 and z = 1 subject to the requirement that A and B are continuous and their
derivatives satisfy the jump conditions
_
A
z
_
z=0
= 0,
_
B
z
_
z=0
+ D
A
x

z=0
= 0, (2.5)
_
A
z
_
z=1
+
B
1 +B
2

z=1
= 0,
_
B
z
_
z=1
= 0, (2.6)
obtained by integrating the model equations across z = 0 and 1, respectively. We employ here the usual notation in
which square brackets denote the jump in a quantity across the specied surface.
Before continuing we draw the readers attention to an unusual feature of the above problem. Since the scalar
eld A is a potential and subject to Neumann boundary conditions, it is only dened up to a constant. This fact
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 103
implies that A behaves like a phase variable in phase dynamics, a fact that is responsible for a number of unexpected
properties of the above problem.
3. Linear theory
We begin by considering the linear problem, i.e., by replacing (z) by (z) = (z 1). In each of the three
regions we seek solutions of the form
A(x, z, t) = a(z) exp (pt +ikx), B(x, z, t) = b(z) exp (pt +ikx),
where p = +i is the complex growth rate. Applying the continuity conditions on A and B, and the matching
and boundary conditions, leads to the dispersion relation [19]
4q
2
sinh 2qL ikD sinh[2q(L 1)] = 0, (3.1)
where q
2
p +k
2
.
3.1. Threshold for instability
As shown by Mason et al. [19] this dispersion relation describes in general two types of modes, a long wave
mode with wavenumber k 1 and a short wave mode with k = O(1). We focus here on the former and take
k = 1.
To compute the marginal stability curve we set = 0 and compute D = D
c
(k) in the form of a series in . Since
within the model the direction of the waves is arbitrary we anticipate that D
c
will be even in k while will be odd.
We therefore write
D
c
= D
0
+
2
D
2
+ , =
10
+
3

30
+ ,
and expand the dispersion relation (3.1) in powers of q
2
i +
2
1:
4q
2
(2L +
4
3
L
3
q
2
+
4
15
L
5
q
4
+
8
315
L
7
q
6
+ ) ikD(2(L 1) +
4
3
(L 1)
3
q
2
+
4
15
(L 1)
5
q
4
+
8
315
(L 1)
7
q
6
+ ) = 0. (3.2)
At O() the resulting problem is purely imaginary and we obtain
D
0
=
4L
10
L 1
. (3.3)
At O(
2
) the problem is purely real and yields

10
=
_
3
2(2L 1)
.
Since L > 1 this quantity is real. In the following we suppose that D
0
> 0 and hence choose the positive sign in
this expression. With this choice of sign all disturbances travel in the negative x direction, i.e., towards the equator.
104 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
At O(
3
) we obtain
D
2
(L 1) 4L
30
=
4
15

10
L[4L
2
+26L 13], (3.4)
while at O(
4
) we obtain
D
2
(L 1)
3
4L
30
(L
2
+2L 1) =
4
105

10
L[12L
4
+46L
3
203L
2
+180L 45]. (3.5)
From these equations we readily deduce the values of D
2
and
30
. In particular D
2
> 0 for all L > 1.
3.2. Growth rate of supercritical dynamo waves
We next suppose that the dynamo number is supercritical, so that > 0. We again write k = , and suppose that
D = D
0
+
2
D
2
+
4
D
4
+ +
2
,
where = O(), but is otherwise an independent small parameter. We anticipate that the growth rate is an even
function of while the frequency is odd in :
=
2

02
+
2

22
+
4

04
+O(
4

2
;
2

4
;
6
),
=
10
+
3

30
+
2

12
+O(
5
;
3

2
;
4
).
The expansion for incorporates the fact that = 0 when = 0, i.e., on the neutral stability curve. Substituting
the above expansions into the dispersion relation and focusing on the coefcients of
2
,
2
,
2

2
and
4
we obtain

02
= 0,
12
=
L 1
4L
,
22
=

10
(2L
2
3L +1)
3L
,
04
= 0, (3.6)
respectively. In summary,
=
10
+
3

30
+
2

12
+O(
5
;
3

2
;
4
), =
2

22
+O(
4

2
;
2

4
;
6
). (3.7)
It follows that when = d, d = O(1), the amplitude of the waves must depend on three distinct timescales: t =
O(
1
), t = O(
3
) and t = O(
4
). This fact complicates considerably the derivation of the amplitude equation.
4. The amplitude equation
We now turn to the nonlinear problem and write
(z) =
(z 1)
1 +B
2
(z 1)(1 B
2
).
The dynamo equations (2.2)(2.3) can be written in the matrix form
L = N(),
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 105
where L is the matrix
L =
_

t

xx

zz
(z 1)
D(z)
x

t

xx

zz
_
,
N is the vector of nonlinear terms
N =
_
(z 1)B
3
0
_
,
and
=
_
A
B
_
.
The linear theory of the preceding section tells us that although waves set in as soon as Dexceeds D
c
these waves
have a O() wavenumber and a O() frequency. This fact suggests that the simplest way to obtain an amplitude
equation for the waves is to use multiple space and time scales. Indeed, as already mentioned, Eq. (3.7) suggests that
we introduce a large spatial scale X = x, together with the time scales T
10
= t, T
30
=
3
t, T
12
=
2
t, T
22
=
2

2
t.
Here x is a short lengthscale and t a fast timescale, although neither scale will appear in the solutions that follow.
As in the linear problem is dened by k = , while
2
is the departure from criticality, i.e. D = D
0
+
2
D
2
+
+
2
. Since as already mentioned the potential A is phase-like we seek a solution in the form
=
_
A
0
0
_
+
_
A
1
B
1
_
+
2
_
A
2
B
2
_
+ ,
where A
0
is O(1) but depends on x and t in the form (x +ct), as appropriate for a traveling wave. Despite
this unusual Ansatz it should be clear that whenever DD
c
= O(
2
) the physical magnetic eld will in fact
be O() since it depends on the derivative of A
0
with respect to x. To proceed we replace
x
by
X
and
t
by

T
10
+
3

T
30
+
2

T
12
+
2

T
22
. The above matrix problem, with the boundary, jump and continuity conditions
specied in Section 2, now becomes a series of problems to be solved at each order in (where = d). Since these
computations are somewhat involved they are relegated to Appendix A. We summarize here the results.
At O(1) we conclude immediately that A
0
= A
0
(x, t) only. At O() the linear problem for (A
1
, B
1
) is inhomo-
geneous and we must impose a solvability condition. This condition yields the relation
A
0
T
10
c
0
A
0
X
= 0, (4.1)
where
c
0
=
D
0
(L 1)
4L
. (4.2)
Note that, as expected, c
0
=
10
/k.
This result suggests that the remainder of the calculation is best performed in a frame moving with a speed c,
where
c = c
0
+
2
c
20
+
2
c
02
+
106 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
and c
0
is given by Eq. (4.2). When = d we can combine the two O(
3
) timescales in a timescale T
3
, relabel the
O(
4
) timescale T
4
, and write instead c = c
0
+
2
c
2
+ etc., where c
2
c
20
+d
2
c
02
=
30
+d
2

12
, etc. This
procedure permits us to look for solutions of the form A = A(, T
3
, T
4
), B = B(, T
3
, T
4
), where = X+cT
10
.
This is accomplished by replacing
t
by c

+
3

T
3
+ and
x
by

; the remaining derivative with respect to


T
3
thus refers to any T
3
dependence in addition to that arising through .
In the moving reference frame the solvability condition (4.1) is replaced by (4.2). The solution of the O()
problem is then (see Appendix A):
A
1
= c
0
A
0

z
2
2
c
0
A
0

Lz +b
4
in z > 1, (4.3)
A
1
= c
0
A
0

z
2
2
+c
0
A
0

Lz +b
5
in z < 1, (4.4)
B
1
=
D
0
2
A
0

z +
LD
0
2
A
0

in z > 0, (4.5)
B
1
=
D
0
2
A
0

z +
LD
0
2
A
0

in z < 0. (4.6)
Here
b
4
= 2c
0
L
A
0

+b
5
but the function b
5
remains undetermined at this order. Note that b
5
is again phase-like, i.e., it is independent of z.
At O(
2
) the solvability condition for A
2
, B
2
is automatically satised and A
2
, B
2
are readily found. At O(
3
)
we once again require a nontrivial solvability condition:
A
0
T
3
a
A
0

3
A
0

3
+
6L
2
c
0
2L 1
_
A
0

_
3
= 0. (4.7)
Here
a = c
2
+
L 1
4L
(D
2
+d
2
) =
c
0
15
(4L
2
+26L 13) > 0,
a relation that follows from Eqs. (3.4) and (3.6). In the following it will be useful to write this equation in the form

A
0
T
3
a

A
0

3

A
0

3
+b

A
2
0

A
0

= 0, (4.8)
where

A
0
A
0

and b = 18L
2
c
0
/(2L 1) > 0. The variable

A
0
is analogous to the local wavenumber in phase
dynamics.
Note that the linearization of Eq. (4.7) with A
0
exp(i
3
p
3
t +i), = (x +ct), c = c
0
+
2
c
2
+ yields
p
3
0, conrming that A
0
exp[i(
10
+
2

30
+
2

12
+ )t +ix] as assumed in Section 3.
4.1. The mKdV equation
Eq. (4.8) is the modied KdVequation and is completely integrable [1]. The equation describes a variety of other
systems as well, including ion acoustic solitons [26,31] and interfacial waves in two-layer liquids with gradually
varying depth [12]. However, in the present case a > 0, b > 0 and no solutions in the form of solitary or cnoidal
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 107
waves are possible. Instead we nd a class of uniformly traveling nonlinear solutions called snoidal waves. Since
these waves are nonlinear we expect that they travel at a speed that differs from the linear speed. To permit drift
with respect to the frame = const., we go into a reference frame traveling with speed v northwards, i.e., we let
= vT
3
, and look for steady solutions of a given period in . These satisfy the equation
a

3

A
0

3
+(a +v)

A
0

A
2
0

A
0

= 0, (4.9)
where

A
0
is now a function of alone. Integrating this equation twice we obtain
1
2

2
0
+
1
2
_
1 +
v
a
_

A
2
0

b
12a

A
4
0
= E, (4.10)
where the prime denotes derivative with respect to . This equation has the solution

A
0
= N
1/2
sn(, s), (4.11)
where
N =
6(a +v)
b
_
s
2
1 +s
2
_
,
2
=
_
1 +
v
a
_
1
1 +s
2
, E =
1
2
N
2
. (4.12)
Here sn is the elliptic function of the rst kind and 0 s 1 is its modulus. Its period is 4K(s)/, where K(s) is
the complete elliptic integral of the rst kind. It follows that solutions of period 2 travel with speed v given by
v = a
_
4(1 +s
2
)K(s)
2

2
1
_
= a
_
3
2
s
2
+
27
32
s
4
+
_
. (4.13)
When s = 0 the solutions are innitesimal in amplitude and sinusoidal; such solutions are stationary in the frame
traveling with the linear velocity c, and indeed v(0) = 0. The amplitude N and speed v of the solutions increase
monotonically with s. Thus nonlinear waves travel towards the equator more slowly than innitesimal waves. In
contrast, on the real line we have a two parameter family of solutions, specied by s and v. It should be emphasized
that, to this order, both s and v are determined by initial conditions; these specify the initial energy E, as well as the
initial momentum. The resulting description is appropriate for the nonlinear evolution of dynamo waves for O(
3
)
times when DD
c
= O(
2
). However, on longer times both forcing and dissipation enter in the description, and
produce a slow drift in the modulus s. The xed point of this drift determines s in terms of the distance d
2
above
the instability threshold, and hence the amplitude and speed of the waves, cf. [17]. This calculation is the focus of
the next section.
5. The perturbed mKdV equation
In order to determine the equilibrated state of the waves for supercritical values of the dynamo number we need
to include the behavior of the system on the time scale T
4
= O(1). To perform this calculation we rst solve for
A
3
and B
3
(see Appendix A), and then proceed to O(
4
). The solvability condition at this order yields an equation
involving both the undetermined function b
5
and the amplitude A
0
satisfying Eq. (4.7). It therefore determines the
unknown function b
5
:
b
5
T
3
a

3
b
5

3
a
b
5

+b
_
A
0

_
2
b
5

2
A
0
T
3
+
_
A
0

_
2

2
A
0

2
+

4
A
0

4
+

2
A
0

2
+
A
0
T
4
= 0,
(5.1)
108 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
where
=
c
0
6
(2L
2
14L +7), =
9L
2
(2L
2
+2L 1)
2(2L 1)
2
, (5.2)
=
1
420(2L 1)
[120L
4
+740L
3
742L
2
+372L 93],
= c
2
+
(2L
3
3L +1)c
0
24L
(D
2
+d
2
).
Note that b
5
= 0.
This result can also be compared with the linear theory discussed in Section 3. We suppose that A
0
, b
5

exp(i
3
p
3
t +i
4
p
4
t +i), where = (x +ct), c = c
0
+
2
c
2
+O(
4
), and linearize Eq. (5.1). Using the results
(3.4)(3.5) relating
30
and D
2
together with the result (3.6) for
12
now shows that p
4
= id
2

22
, and hence that
A
0
, b
5
exp[i(
10
+
2

30
+
2

12
+ )t +ix +
2

22
t] as assumed in Section 3.
To obtain an evolution equation that includes the behavior of the system on the T
4
timescale we follow Aspe and
Depassier [2] and reconstitute the amplitude equation. In order to do this we rst differentiate Eq. (5.1) with respect
to , and rewrite it in terms of

A
0
A
0

,

b
5
b
5

. We then return to Eqs. (4.3)(4.4) for A


1
in z > 1 and z < 1
and observe that these equations contain a common phase-like quantity b
45
(b
4
+b
5
)/2 = c
0
L(A
0
/) +b
5
.
Thus we set C =

A
0
+

b
45
and construct an evolution equation in the moving reference frame for the redened
phase-like variable C(, ):
C

a
C

3
C

3
+bC
2
C

+f = O(
2
). (5.3)
Here / = /T
3
+/T
4
+ , and
f = (a +)

2
C

2
+(a +)

4
C

4
+( b)

2

2
_
1
3
C
3
_
. (5.4)
We refer to Eq. (5.3) as the perturbed mKdV equation.
5.1. The bifurcation diagram
In this Section we use the perturbed mKdVequation to construct the bifurcation diagramfor long dynamo waves.
As already noted we expect that the energy E or equivalently the modulus s of the waves will evolve as a consequence
of the perturbation f, cf. [14]. To determine the effect of this perturbation we write Eq. (5.3) in the moving frame
a

3
C

3
+(a +v)
C

bC
2
C

= f +O(
2
), (5.5)
multiply by it Cand integrate over a period of C. This procedure yields the following exact condition for the presence
of a periodic solution of the perturbed mKdV equation:
_
Cf d = O(). (5.6)
This condition can be approximated using C =

A
0
+O() and 4K(s)/ for the period, yielding the condition
_
4K(s)/
0

A
0
f d = O(), (5.7)
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 109
Fig. 2. The amplitude N as a function of d
2
computed from Eq. (5.8) (solid line) compared with the perturbation result (5.17) (dashed line)
(L = 10).
where

A
0
and are given by Eqs. (4.11)(4.12), with v given by (4.13). The quantity (5.7) is therefore a function
of s, h(s) say, and we therefore seek the root s = s of h(s) = 0 as a function of the supercriticality parameter d
2
,
i.e., of [7]
h(s)
1

_
4E
_
+
b
a
_

_
1 +
v
a
_ _
(a +)
v
a
_
_
NI
2
+
1

_
b
3a
_
(a +)
v
a
_
3
_
1 +
v
a
_
_
+
b
a
__
N
2
I
4
+
2b
3a
_
+
b
a
_
N
3
I
6
= 0. (5.8)
Here I
n
=
_
4K(s)
0
sn
n
d. From s we reconstruct the bifurcation diagram N(d
2
) (Fig. 2), and compute the speed
v (Fig. 3) as a function of d
2
. Fig. 4 shows the corresponding snoidal wave for d
2
= 100 ( s 0.79). Note that
dh/ds at s = s determines the stability of the solution: the solution is stable in time if dh/ds > 0 and unstable if
dh/ds < 0. Fig. 5 illustrates h(s) for d
2
= 25, with the positive gradient at s 0.5 indicating a stable solution.
5.2. Weakly nonlinear theory
We can check the results of the preceding section using weakly nonlinear theory. We start with the rst integral
of Eq. (5.5):
Fig. 3. The speed v as a function of d
2
computed from Eq. (5.8) (solid line) compared with the perturbation results (5.13) and (5.15) (dashed
line) (L = 10).
110 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
Fig. 4. The snoidal wave

A
0
for d
2
= 100 ( s 0.79, L = 10).
a

2
C

2
+(a +v)C
1
3
bC
3
=
_
(a +)
C

+(a +)

3
C

3
+( b)C
2
C

_
+O(
2
), (5.9)
and suppose that
C =
1/2
C
0
+
3/2
C
1
+
5/2
C
2
+ , v = v
1
+
2
v
2
+ , =
c
+
1
. (5.10)
Substituting these expansions into Eq. (5.9) we obtain at O(
1/2
) the result
C

0
+C
0
= 0 (5.11)
with the solution C
0
= Rsin up to an arbitrary phase. At O(
3/2
) we obtain
C

1
+C
1
=
b
3a
C
3
0
+
1
a
_
v
1
C
0
+(a +
c
)C

0
+(a +)C

0
_
. (5.12)
The solvability conditions for this problem are
v
1
=
1
4
bR
2
,
c
= . (5.13)
When these conditions hold we can solve for C
1
and obtain C
1
= (b/96a)R
3
sin 3. Finally, at O(
5/2
) we obtain
C

2
+C
2
=
1
a
[bC
2
0
C
1
v
2
C
0
v
1
C
1
+(a +
c
)C

1
+(a +)C

1
+
1
C

0
+( b)C
2
0
C

0
]. (5.14)
Fig. 5. The function h(s) dened by Eq. (5.8) for d
2
= 25, L = 10.
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 111
The solvability conditions are now
v
2
=
b
2
384a
R
4
,
1
+
1
4
( b)R
2
= 0. (5.15)
The latter equation determines the direction of branching of long dynamo waves. We obtain
R
2
=
2L 1
9L
2

1
. (5.16)
Since
1
= (L 1)(2L 1)(c
0
/3L)(d
2
/) > 0 the primary bifurcation is supercritical for all values of L > 1. In
Fig. 3 the dashed line represents the speed v as a function of d
2
as computed from the above expansion, while the
corresponding line in Fig. 2 represents the resulting amplitude N:
N =
(2L 1)
2
(L 1)
27L
3
c
0
d
2
. (5.17)
It is of interest to compare these results with the exact result computed in Section 4.1. From the expansion
(4.13) for v and the corresponding expansion for N we conclude that
s
2
=
_
bR
2
6a
_

5
8

2
_
bR
2
6a
_
2
+ , (5.18)
and hence that
v =
1
4
bR
2

2
b
2
R
4
384a
+ ,
as obtained via perturbation theory. To obtain expression (5.18) we used the fact that sn(p, s) = [1 +(s
2
/16) +
] sin q +(s
2
/16 + ) sin 3q + , where q = (/2K)p.
Similarly, expanding the result (5.8) in powers of s
2
we obtain
h(s) =
N

_
+
3
8
( )s
2
+
3a
2b
(b )s
2
+
_
.
The results (5.13b) and (5.15b) now follow on writing =
c
+
1
, and using (5.18).
5.3. Wavelength selection
In Section 4 we imposed the wavelength of the solution; without loss of generality we chose this wavelength
to be 2. Although this is normal procedure in problems of this kind we expect that in the physical situation the
wavelength, or equivalently the wave speed v, will be selected by physical processes. To see how this comes about
we multiply Eq. (5.9) by C and integrate the result over a period. We obtain
v =
_
[a(C

2
C
2
) +(b/3)C
4
] d
_
C
2
d
+O(
2
). (5.19)
This is an exact expression. Note, however, that the perturbation f has dropped out. For this reason this equation is
identically satised to O(), a fact that can be readily checked using an expansion of the form C =

A
0
+C
1
+
in both (5.19) and (5.9). Thus any selection of the speed comes about at O(
2
), a calculation that is beyond the scope
of this paper. However, in contrast to mechanisms responsible for wavelength selection in other pattern forming
systems, exemplied by the so-called Busse balloon, we expect the selection based on Eq. (5.19) to be sharp.
112 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
Fig. 6. Prole of B
1
(, z) at = /2 corresponding to a maximum value of

A
0
when d
2
= 100. At = 3/2, corresponding to a minimum
value of

A
0
, the sign of B
1
(, z) is reversed. Note that = /2 also corresponds to a zero of A
0
(with the mean subtracted off), and at = 0, ,
corresponding to minimum and maximum values of A
0
, respectively, B
1
(, z) = 0 ( s 0.79, L = 10).
5.4. Physical manifestation of the solution
The corresponding solutions for the elds A
0
, A
1
, B
1
can be reconstructed from the solution (4.11) to give
A
0
=
N
1/2
s
ln[dn(, s) s cn(, s)],
to within an arbitrary constant of integration, and
A
1
= N
1/2
(
1
2
c
0
z
2
c
0
Lz +2c
0
L) sn(, s) +b
5
in z > 1,
A
1
= N
1/2
(
1
2
c
0
z
2
+c
0
Lz) sn(, s) +b
5
in z < 1,
B
1
=
1
2
N
1/2
D
0
(L z) sn(, s) in z > 0,
B
1
=
1
2
N
1/2
D
0
(L +z) sn(, s) in z < 0.
Fig. 7. The toroidal (B/, solid line) and poloidal elds (|B
P
|/ for and |B
P
|/ for , |B
P
|
2

2
[(A
1
/z)
2
+

A
2
0
], dashed line)
at z = 0 when d
2
= 100 ( s 0.79, L = 10).
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 113
Fig. 6 shows the resulting vertical prole of B
1
(, z) at = /2, chosen to correspond to a maximum of

A
0
.
The leading order contributions to both the toroidal and poloidal elds are shown together in Fig. 7; both are
of order , despite the fact that A
0
= O(1), with the toroidal eld approximately twice as strong as the poloidal
eld.
6. Discussion
In this paper we have derived an amplitude equation for slightly supercritical long wavelength dynamo waves.
The derivation was based on a simplied model of the mean-eld dynamo equations, and led to a leading order
description of the waves in terms of a modied Kortewegde Vries equation. This equation is exactly solvable,
and in our case describes strongly nonlinear waves called snoidal waves. For positive dynamo numbers the waves
propagate in the negative x direction, corresponding to propagation towards the equator, as observed in the Sun. In
order to describe the growth and equilibration of the magnetic eld we considered dynamo numbers O(
2
) above
critical, where
1
measures the wavelength of the waves (in units of the depth of the layer where dynamo action
takes place). At such values of the dynamo number the amplitude of the waves grows on a yet slower timescale and
nonlinear quenching of the effect saturates the magnetic eld at O() amplitude. The amplitude and speed of the
resulting dynamo wave are related to the distance d
2
from threshold for the instability, and are approximated well
by second order perturbation theory. Our results indicate that the bifurcation producing these waves is supercritical,
suggesting that the waves are stable.
Because of their long wavelength the waves are described at leading order by an integrable amplitude equation,
with forcing and dissipation entering only at higher order. Consequently we have had to employ multiple scale
methods involving three distinct timescales (in addition to a slow spatial scale), followed by a procedure that has
been termed reconstitution to describe the effects of the higher order terms on the leading order dynamics. This
procedure is frequently used in applied mathematics, e.g., [2], although from the point of view of asymptotics the
results cannot be justied. However, in many instances an alternative iterative procedure leads to identical results,
e.g., [17], and in some of these the results can be justied rigorously via normal form theory [10,11]. We have not
attempted here these extensions of the theory.
Although our results are based on a specic and highly idealized model of the dynamo process, we believe that
many aspects of our results have general applicability. We have already noted that the structure of the expansion
owes much to the fact that the poloidal magnetic eld is described in terms of a (vector) potential. As a consequence
the potential A is phase-like, and the resulting expansion procedure resembles that familiar fromstudies of xed-ux
convection [8,16] and of phase dynamics [18]. As is well known, in phase dynamics the leading order description
generally leads to Burgers equation (e.g., [4]), or if one deals with low-frequency long waves, the Korteweg
de Vries equation. The reason we obtain here the modied Kortewegde Vries equation can be traced to the
symmetry of the mean eld equations under B B. These considerations suggest that long wave, low frequency
dynamo waves will in general be described by the modied Kortewegde Vries equation, albeit it with different
coefcients, and raise the possibility that solitary dynamo waves may be a general property of interface dynamo
models.
Acknowledgements
The second author spent 5 months in 1988 in the group of Professor Kuramoto as a JSPS Fellow [16], and
is grateful to Professor Kuramoto and all members of his group at that time for their warm and memorable
hospitality in Kyoto, and friendships that continue to this day. This work was supported in part by EPSRC
under grant GR/R52879/01, an EPSRC studentship, and the High Altitude Observatory. We are grateful to Prof.
D.W. Hughes and Dr S.M. Tobias for suggesting this investigation, and to Dr M.C. Depassier for a helpful
discussion.
114 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
Appendix A. Derivation of the mKdV equation
In this appendix we summarize the results of the various steps required to derive the amplitude equation (4.8).
We begin with the matrix problem
L = N(),
where L is the matrix
L =
_

t

xx

zz
(z 1)
D(z)
x

t

xx

zz
_
,
N is the vector of nonlinear terms
N =
_
(z 1)B
3
0
_
,
and
=
_
A
B
_
,
together with the boundary conditions (2.4) and jump relations (2.5), (2.6). As explained in Section 4, we write
D = D
0
+
2
D
2
+ +
2
(where = d), replace
x
by
X
and
t
by
T
10
+
3

T
30
+
2

T
12
+
2

T
22
, and
seek a solution in the form
=
0
+
1
+
2

2
+ =
_
A
0
0
_
+
_
A
1
B
1
_
+
2
_
A
2
B
2
_
+
The matrix problem then becomes a sequence of problems to be solved at each order in , i.e. we solve
L
i

i
= q
i
for i = 0, 1, . . .
We shall nd that the linear operator L
0
has a nontrivial nullspace (spanned by vectors of the form (A
0
, 0), where
A
0
is independent of z), so that we need to impose a solvability condition at each order. For this purpose we require
the solution of the adjoint problem, as described next.
A.1. The adjoint problem
To nd the adjoint problem to
L
0

i
= 0
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 115
(including appropriate boundary conditions, jump relations and continuity conditions) we use integration by parts
to write the scalar product , L
0

i
in the form
, L
0

i
=
i
, L

0
+surface terms.
The adjoint problem is then dened to be
L

0
= 0,
subject to boundary conditions that eliminate the surface terms for any
i
. Here
, L
0

i
=
_
z
(
1
,
2
)
_

zz
A
i
(z 1)B
i

zz
B
i
_
=
_
z
(A
i
, B
i
)
_

zz

zz

2
(z 1)
1
_

1
A
i
z
A
i

1
z
+
2
B
i
z
B
i

2
z
_
L
L
. (A.1)
Since B
i
(x, z = L, t) = 0 and (A
i
/z)(x, z = L, t) = 0, the surface terms are eliminated by choosing
2
(z =
L) = 0,
1
/z(z = L) = 0, leaving the adjoint problem
_

zz
0
(z 1)
zz
__

2
_
= 0,

1
z
(x, z = L, t) = 0,
2
(x, z = L, t) = 0.
In addition
1
and
2
must be continuous throughout the domain, and satisfy appropriate jump conditions derived
by integrating the equations across z = 1. We obtain

1
= independent of z

2
=

1
(1 +L)z
2L
+

1
(1 +L)
2
in z > 1, (A.2)

2
=

1
(1 L)z
2L


1
(1 L)
2
in z < 1.
A.2. The amplitude equation
The leading order problem is considerably simpler than those that followdue to B
0
= 0. We are required to solve

2
A
0
z
2
= 0,
A
0
z
(x, z = L, t) = 0.
Since A
0
is continuous throughout the domain it must be independent of z, although it will depend on X, T
10
, etc.
At O() we obtain the problem
L
0

1
= q
1
, (A.3)
116 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
where
L
0
=
_

zz
(z 1)
0
zz
_
,
q
1
=
_

T
10
0
D
0
(z)
X

T
10
__
A
0
0
_
,
with the boundary conditions
B
1
(x, z = L, t) = 0,
A
1
z
(x, z = L, t) = 0.
In addition A
1
and B
1
must be continuous throughout the domain and satisfy the jump conditions
_
B
1
z
_
z=0
+D
0
A
0
X
(x, z = 0, t) = 0, (A.4)
_
A
1
z
_
z=1
+B
1
(x, z = 1, t) = 0. (A.5)
The problem (A.3) has a solution if and only if
, q
1
= 0, (A.6)
where is given by Eq. (A.2). The solvability condition is thus
A
0
T
10

D
0
(L 1)
4L
A
0
X
= 0.
Equivalently, if we look for solutions in the form A
0
(X, T
10
, T
30
, . . .) A
0
(, T
30
, . . .), where = X+cT
10
, and
write
c = c
0
+
2
c
20
+
2
c
02
+O(
4
),
we obtain at O() the problem
L
0
_
A
1
B
1
_
= L
1
_
A
0
0
_
,
where L
0
is as above and
L
1
=
_
c
0

0
D
0
(z)

c
0

_
,
subject to the boundary conditions
B
1
(x, z = L, t) = 0,
A
1
z
(x, z = L, t) = 0. (A.7)
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 117
Moreover, A
1
and B
1
are continuous throughout the domain, and the jump conditions read
_
B
1
z
_
z=0
+D
0
A
0

(x, z = 0, t) = 0, (A.8)
_
A
1
z
_
z=1
+B
1
(x, z = 1, t) = 0. (A.9)
The solvability condition , L
1

0
= 0 for this problem yields
A
0

_
1 +
D
0
(1 L)
4Lc
0
_
= 0, (A.10)
or, equivalently,
c
0
=
D
0
(L 1)
4L
,
as obtained above.
With this choice of c
0
we can solve the O() problem. The equation for B
1
reads

2
B
1
z
2
= D
0
(z)
A
0

.
Hence,
B
1
= a
2
z +b
2
in z > 0,
B
1
= a
3
z +b
3
in z < 0,
where a
2
, a
3
, b
2
, b
3
are independent of z but may be functions of the large space and slowtime scales. The boundary
conditions (A.7) give
a
2
L +b
2
= 0, b
3
a
3
L = 0,
while continuity of B
1
across z = 0 gives
b
2
= b
3
,
and the jump relation (A.8) yields
a
2
a
3
+D
0
A
0

= 0.
These equations can be solved for the unknowns a
2
, a
3
, b
2
, b
3
and we obtain
B
1
=
D
0
2
A
0

z +
LD
0
2
A
0

in z > 0,
118 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
B
1
=
D
0
2
A
0

z +
LD
0
2
A
0

in z < 0.
For A
1
we solve

2
A
1
z
2
+(z 1)B
1
= c
0
A
0

,
obtaining
A
1
= c
0
A
0

z
2
2
+a
4
z +b
4
in z > 1,
A
1
= c
0
A
0

z
2
2
+a
5
z +b
5
in z < 1,
where a
4
, a
5
, b
4
, b
5
are independent of z. The boundary conditions (A.7) give
a
4
= c
0
L
A
0

, a
5
= a
4
.
Applying continuity in A
1
across z = 1 gives
b
4
= 2c
0
L
A
0

+b
5
, (A.11)
and applying the jump relation (A.9) yields
A
0

_
1
2
D
0
(L 1) 2c
0
L
_
= 0,
a condition that is automatically satised by our choice of c
0
. Thus
A
1
= c
0
A
0

z
2
2
c
0
A
0

Lz +b
4
in z > 1,
A
1
= c
0
A
0

z
2
2
+c
0
A
0

Lz +b
5
in z < 1,
where b
4
is related to b
5
through Eq. (A.11). Note that b
5
remains undetermined at this order.
At O(
2
) we obtain the problem
L
0
_
A
2
B
2
_
= L
1
_
A
1
B
1
_
+L
2
_
A
0
0
_
,
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 119
where L
0
and L
1
are as above, and
L
2
=
_

0
0

_
,
subject to
B
2
(x, z = L, t) = 0,
A
2
z
(x, z = L, t) = 0. (A.12)
As before A
2
and B
2
are continuous throughout the domain, and satisfy the jump conditions
_
B
2
z
_
z=0
+D
0
A
1

(x, z = 0, t) = 0, (A.13)
_
A
2
z
_
z=1
+B
2
(x, z = 1, t) = 0. (A.14)
The solvability condition for this problem reads
, L
1

1
+L
2

0
= 0,
and yields
_
z
(
1
,
2
)
_

A
0
c
0

A
1
D
0
(z)

A
1
c
0

B
1
_
= 0,
and is automatically satised. The O(
2
) problem is therefore solvable and we obtain
B
2
=
c
0
D
0
4

2
A
0

2
_
2L
3
3
Lz
2
+
z
3
4
_
+
D
0
2
b
5

(L z) in z > 0, (A.15)
B
2
=
c
0
D
0
4

2
A
0

2
_
2L
3
3
Lz
2

z
3
4
_
+
D
0
2
b
5

(L +z) in z < 0. (A.16)


The equation for A
2
,

2
A
2
z
2
+(z 1)B
2
= c
0
A
1


2
A
0

2
,
yields
A
2
=
c
2
0
24

2
A
0

2
z
4

c
2
0
L
6

2
A
0

2
z
3
+
_
c
0
b
4


2
A
0

2
_
z
2
2
+a
8
z +b
8
in z > 1,
A
2
=
c
2
0
24

2
A
0

2
z
4
+
c
2
0
L
6

2
A
0

2
z
3
+
_
c
0
b
5


2
A
0

2
_
z
2
2
+a
9
z +b
9
in z < 1.
120 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
The functions a
8
, a
9
, b
8
and b
9
are related by the boundary conditions (A.12),
a
8
= L

2
A
0

2
c
0
L
b
4

+
c
2
0
L
3
3

2
A
0

2
,
a
9
= c
0
L
b
5

2
A
0

2

c
2
0
L
3
3

2
A
0

2
,
and continuity of A
2
across z = 1:
b
8
=
c
2
0
L
3

2
A
0

2
+
c
0
2
b
5

c
0
2
b
4

+a
9
+b
9
a
8
. (A.17)
The jump condition (A.14) is then automatically satised. The solution at this order therefore depends on two
unknown phase-like functions, b
5
and b
9
.
At O(
3
) we have
L
0
_
A
3
B
3
_
= L
1
_
A
2
B
2
_
+L
2
_
A
1
B
1
_
+L
3
_
A
0
B
0
_
+N
3
,
where L
0
, L
1
, L
2
are as above and
L
3
=
_
c
2

T
3
0
(D
2
+d
2
)(z)

0
_
, N
3
=
_
(z 1)B
3
1
0
_
.
Here c
2
= c
20
+d
2
c
02
and
T
3
=
T
30
+d
2

T
12
. The boundary conditions are
B
3
(x, z = L, t) = 0,
A
3
z
(x, z = L, t) = 0, (A.18)
and A
3
and B
3
are continuous throughout the domain and subject to the jump conditions
_
B
3
z
_
z=0
+D
0
A
2

(x, z = 0, t) +(D
2
+d
2
)
A
0

(x, z = 0, t) = 0, (A.19)
_
A
3
z
_
z=1
+B
3
(x, z = 1, t) B
3
1
(x, z = 1, t) = 0. (A.20)
The solvability condition for this problem
, L
1

2
+L
2

1
+L
3

0
+N
3
= 0,
leads to the result (4.7). Differentiating with respect to and writing

A
0
= A
0
we obtain the mKdV Eq. (4.8),
namely

A
0
T
3
a

A
0

3

A
0

3
+b

A
2
0

A
0

= 0, (A.21)
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 121
where a = c
0
(4L
2
+26L 13)/15 > 0, and b = 18L
2
c
0
/(2L 1) > 0. With the solvability condition (4.7) we
can now solve the O(
3
) problem:
B
3
=
c
2
0
D
0
240

3
A
0

3
z
5
+
c
2
0
LD
0
48

3
A
0

3
z
4
+
_
D
0
2

3
A
0

3
+c
0
a
6

_
z
3
6
+
_
c
0
b
6

LD
0
2

3
A
0

3
_
z
2
2
+a
10
z +b
10
in z > 0, (A.22)
B
3
=
c
2
0
D
0
240

3
A
0

3
z
5
+
c
2
0
LD
0
48

3
A
0

3
z
4
+
_
c
0
a
7

D
0
2

3
A
0

3
_
z
3
6
+
_
c
0
b
7

LD
0
2

3
A
0

3
_
z
2
2
+a
11
z +b
11
in z < 0, (A.23)
where a
10
, a
11
, b
10
, b
11
are to be determined by continuity and the relations (A.18) and (A.19). The boundary
conditions (A.18) give
c
2
0
D
0
L
5
60

3
A
0

3

D
0
L
3
6

3
A
0

3
+
c
0
L
3
6
a
6

+
c
0
L
2
2
b
6

+a
10
L +b
10
= 0, (A.24)
and
c
2
0
D
0
L
5
60

3
A
0

3

D
0
L
3
6

3
A
0

3

c
0
L
3
6
a
7

+
c
0
L
2
2
b
7

a
11
L +b
11
= 0. (A.25)
The remaining conditions lead to
a
10
= a
11
=
1
2
_
D
0
b
9

+(D
2
+d
2
)
A
0

_
,
and
b
10
= b
11
=
c
2
0
D
0
L
5
60

3
A
0

3
+
D
0
L
3
6

3
A
0

3

c
0
L
3
6
a
6

c
0
L
2
2
b
6

+
LD
0
2
b
9

+
L(D
2
+d
2
)
2
A
0

.
(A.26)
Similarly,
A
3
=
c
3
0
720

3
A
0

3
z
6

c
3
0
L
120

3
A
0

3
z
5
+
_
c
0

2
b
4

2
2

3
A
0

3
_
c
0
z
4
24
+
c
0
z
3
6
_
a
8

+L

3
A
0

3
_
+
_
c
0
b
8


2
b
4

2
+ c
2
A
0

+
A
0
T
3
_
z
2
2
+a
12
z +b
12
in z > 1, (A.27)
A
3
=
c
3
0
720

3
A
0

3
z
6
+
c
3
0
L
120

3
A
0

3
z
5
+
_
c
0

2
b
5

2
2

3
A
0

3
_
c
0
z
4
24
+
c
0
z
3
6
_
a
9

3
A
0

3
_
+
_
c
0
b
9


2
b
5

2
+ c
2
A
0

+
A
0
T
3
_
z
2
2
+a
13
z +b
13
in z < 1, (A.28)
122 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
where a
12
, a
13
, b
12
, b
13
are independent of z and are related through conditions (A.18) and (A.19). The boundary
conditions (A.18) imply
a
12
=
c
3
0
L
5
30

3
A
0

3

c
0
L
3
6
_

3
A
0

3
+c
0

2
b
4

2
_

c
0
L
2
2
a
8

+L
_

2
b
4

2
c
0
b
8

c
2
A
0

A
0
T
3
_
,
(A.29)
and
a
13
=
c
3
0
L
5
30

3
A
0

3
+
c
0
L
3
6
_
c
0

2
b
5

2
+

3
A
0

3
_

c
0
L
2
2
a
9

+L
_
c
0
b
9


2
b
5

2
+ c
2
A
0

+
A
0
T
3
_
,
(A.30)
while continuity of A
3
across z = 1 gives
b
12
=
c
3
0
L
60

3
A
0

3
+
c
2
0
24

2
b
5

2
+
c
0
6
_
a
9

2L

3
A
0

3
_
+
c
0
2
b
9

1
2

2
b
5

2
+a
13
+b
13

c
2
0
24

2
b
4

c
0
6
a
8

c
0
2
b
8

+
1
2

2
b
4

2
a
12
. (A.31)
The solvability condition (4.7) is recovered from the jump condition (A.20). Note that we now have three arbitrary
phase-like functions, namely b
5
, b
9
and b
13
.
Finally, at O(
4
), we obtain the problem
L
0
_
A
4
B
4
_
= L
1
_
A
3
B
3
_
+L
2
_
A
2
B
2
_
+

L
3
_
A
1
B
1
_
+L
4
_
A
0
0
_
+N
4
,
where L
0
, L
1
, L
2
are as above

L
3
=
_
c
2

T
3
0
(D
2
+d
2
)(z)

c
2

T
3
_
, L
4
=
_

T
4
0
0 0
_
,
with
T
4
= d
2

T
22
, and
N
4
=
_
3(z 1)B
2
1
B
2
0
_
.
The boundary conditions are
B
4
(x, z = L, t) = 0,
A
4
z
(x, z = L, t) = 0, (A.32)
with A
4
and B
4
continuous throughout the domain, and satisfying the jump conditions
_
B
4
z
_
z=0
+D
0
A
3

(x, z = 0, t) +(D
2
+d
2
)
A
1

(x, z = 0, t) = 0, (A.33)
_
A
4
z
_
z=1
+B
4
(x, z = 1, t) 3B
2
1
(x, z = 1, t)B
2
(x, z = 1, t) = 0. (A.34)
J. Mason, E. Knobloch / Physica D 205 (2005) 100124 123
The solvability condition for this problem reads
, L
1

3
+L
2

2
+

L
3

1
+L
4

0
+N
4
= 0,
and yields Eq. (5.1) relating b
5
and A
0
.
As explained in Section 5, we then seek the reconstituted amplitude equation for the phase-like variable
C =

A
0
+(

b
4
+

b
5
)/2. Differentiating (5.1) with respect to , multiplying by , and adding the resulting equation
to (4.8) yields
C

a
C

3
C

3
+bC
2
C

f = O(
2
), (A.35)
where / = /T
3
+/T
4
+ , and

f =

2
C

2
C

2
+

2
_
C
3
3
_
+

4
C

4
. (A.36)
The rst term in

f involving the time derivative can be eliminated using Eq. (A.35), resulting in Eqs. (5.3) and (5.4).
References
[1] M.J. Ablowitz, D.J. Kaup, A.C. Newell, H. Segur, The inverse-scattering transform Fourier analysis for nonlinear problems, Stud. Appl.
Math. 53 (1974) 249315.
[2] H. Aspe, M.C. Depassier, Evolution equation of surface waves in a convecting uid, Phys. Rev. A 41 (1990) 31253128.
[3] J. Beer, S. Tobias, N. Weiss, An active Sun throughout the Maunder minimum, Solar Phys. 181 (1998) 237249.
[4] A.J. Bernoff, Slowly varying fully nonlinear wavetrains in the GinzburgLandau equation, Physica D 30 (1988) 363381.
[5] N.H. Brummell, N.E. Hurlburt, J. Toomre, Turbulent compressible convection with rotation. II. Mean ows and differential rotation,
Astrophys. J. 493 (1998) 955969.
[6] P. Bushby, J. Mason, Understanding the solar dynamo, Astron. Geophys. 45 (2004) 713.
[7] P.F. Byrd, M.D. Friedman, Handbook of Elliptic Integrals for Engineers and Scientists, Springer-Verlag, New York, 1971.
[8] C.J. Chapman, M.R.E. Proctor, Nonlinear RayleighB enard convection between poorly conducting boundaries, J. Fluid Mech. 101 (1980)
759782.
[9] P. Charbonneau, K.B. MacGregor, On the generation of equipartition-strength magnetic elds by turbulent hydromagnetic dynamos,
Astrophys. J. 473 (1996) L59L62.
[10] J. Guckenheimer, P. Holmes, Nonlinear Oscillations, Dynamical Systems and Bifurcations of Vector Fields, Springer-Verlag, New York,
1997.
[11] J. Guckenheimer, E. Knobloch, Nonlinear convection in a rotating layer: amplitude expansions and normal forms, Geophys. Astrophys.
Fluid Dyn. 23 (1983) 247272.
[12] K.R. Helfrich, W.K. Melville, J.W. Miles, On interfacial solitary waves over slowly varying topography, J. Fluid Mech. 149 (1984) 305
317.
[13] S.A. Jepps, Numerical models of hydromagnetic dynamos, J. Fluid Mech. 67 (1975) 625646.
[14] V.I. Karpman, E.M. Maslov, Perturbation theory for solitons, Sov. Phys. JETP 46 (1977) 281291.
[15] N.I. Kleeorin, A.A. Ruzmaikin, Properties of a nonlinear solar dynamo model, Geophys. Astrophys. Fluid. Dyn. 17 (1981) 281296.
[16] E. Knobloch, Pattern selection in binary uid convection at positive separation ratios, Phys. Rev. A 40 (1989) 15491559.
[17] E. Knobloch, M.R.E. Proctor, Nonlinear periodic convection in double-diffusive systems, J. Fluid Mech. 108 (1981) 291316.
[18] Y. Kuramoto, Chemical Oscillations, Waves and Turbulence, Springer-Verlag, New York, 1984.
[19] J. Mason, D.W. Hughes, S.M. Tobias, The competition between surface and deep-seated -effects, Astrophys. J. 580 (2002) L89L92.
[20] H.K. Moffatt, Magnetic Field Generation in Electrically Conducting Fluids, Cambridge University Press, 1978.
[21] E.N. Parker, Hydromagnetic dynamo models, Astrophys. J. 122 (1955) 293314.
[22] E.N. Parker, A solar dynamo surface wave at the interface between convection and nonuniform rotation, Astrophys. J. 408 (1993) 707
719.
[23] A.A. Ruzmaikin, A.M. Shukurov, D.D. Sokoloff, Magnetic Fields of Galaxies, Kluwer Academic Publishers, 1988.
124 J. Mason, E. Knobloch / Physica D 205 (2005) 100124
[24] J. Schou, et al., Helioseismic studies of differential rotation in the solar envelope by the solar oscillations investigation using the Michelson
Doppler imager, Astrophys. J. 505 (1998) 390417.
[25] M. Steenbeck, F. Krause, K.H. R adler, A calculation of the mean electromotive force in an electrically conducting uid in turbulent motion
under the inuence of Coriolis forces, Z. Naturforsch. 21a (1966) 369376.
[26] M. Tajiri, K. Nishihara, Solitons and shock waves in two-electron-temperature plasmas, J. Phys. Soc. Jpn. 54 (1985) 572578.
[27] M.J. Thompson, J. Christensen-Dalsgaard, M.S. Miesch, J. Toomre, The internal rotation of the Sun, Annu. Rev. Astron. Astrophys. 41
(2003) 599643.
[28] S.M. Tobias, Diffusivity quenching as a mechanism for Parkers surface dynamo, Astrophys. J. 467 (1996) 870880.
[29] S.M. Tobias, M.R.E. Proctor, E. Knobloch, The role of absolute instability in the solar dynamo, Astron. Astrophys. 318 (1997) L55L58.
[30] S.M. Tobias, M.R.E. Proctor, E. Knobloch, Convective and absolute instabilities of uid ows in nite geometry, Physica D 113 (1998)
4372.
[31] S. Watanabe, Ion acoustic soliton in plasma with negative ion, J. Phys. Soc. Jpn. 53 (1984) 950956.
[32] H. Yoshimura, A model of the solar cycle driven by the dynamo action of the global convection in the solar convection zone, Astrophys.
J. Suppl. 294 (1975) 467494.

Вам также может понравиться