1: \documentclass[a4paper,11pt]{article}
2: \usepackage{amssymb} %,amsfonts,amsmath}
3: %\usepackage{psfrag}
4: \usepackage{epsfig}
5: \usepackage{natbib}
6: \bibpunct{(}{)}{;}{a}{,}{;}
7: %\let\citet=\cite
8: %\let\citep=\cite
9:
10: \usepackage{latexsym}
11: \newcommand{\eg}{{e.g.\ }}
12: \newcommand{\etc}{{etc.\ }}
13: \newcommand{\etal}{\mbox{\it et al.\ }}
14: \newcommand{\Rey}{\mbox{\it Re}}
15: %\newcommand{\Rm}{\Rey_m}
16: \newcommand{\Rm}{R_m}
17: \newcommand{\Pm}{P_m}
18: \newcommand{\upartial}{\partial}
19: \newcommand{\upi}{\pi}
20: \newcommand{\bnabla}{\mbox{\boldmath $\nabla$}}
21: %\renewcommand{\vec}[1]{\mbox{\boldmath $#1$}}
22: \renewcommand{\vec}[1]{{\bf #1}}
23: \newcommand{\vechat}[1]{{\skew3\hat{\vec{#1}}}}
24:
25: \font\sls = cmssi10 % s-s bold maths, italic
26: \newcommand{\mat}[1]{\mbox{\sls #1}}
27:
28: \newcommand{\rf}[1]{(\ref{#1})}
29:
30: \newcommand{\laplace}{\nabla^2} % -> \Delta
31: \newcommand{\cross}{\wedge}
32: \newcommand{\grad}{\bnabla}
33: \newcommand{\dvgnce}{\bnabla \cdot}
34: \newcommand{\curl}{\bnabla \wedge}
35: \newcommand{\avge}[1]{\langle #1 \rangle}
36: \newcommand{\inner}[2]{\langle #1 , #2 \rangle}
37: \newcommand{\norm}[1]{\| #1 \|}
38: \newcommand{\mod}[1]{\, |#1| \,}
39: \newcommand{\sgn}{\,\mbox{sgn}\,}
40: \newcommand{\der}[3]{\frac{{\mathrm d}^{#3}#1}{{\mathrm d} #2^{#3}}}
41: \newcommand{\pder}[3]{\frac{\upartial^{#3}#1}{\upartial #2^{#3}}}
42: \newcommand{\pd}[1]{\upartial_{#1}}
43: \newcommand{\dt}{{\Delta t}}
44: \newcommand{\bess}{{\mathcal{B}}}
45: \newcommand{\ord}{{\mathcal{O}}}
46: \newcommand{\ex}{{\mathrm e}}
47: \newcommand{\im}{{\mathrm i}}
48:
49:
50: %---------------------------------------------------------------------------
51:
52: \title{
53: Kinematic dynamo action in a sphere:
54: Effects of periodic time-dependent flows on solutions with axial dipole
55: symmetry.
56: }
57:
58: \author{
59: Ashley P. Willis\ \ and\ \ David Gubbins
60: %}
61: %
62: \\ \it
63: %\address{
64: School of Earth Sciences, University of Leeds, LS2 9JT, UK
65: }
66:
67: \date{\today}
68:
69: %---------------------------------------------------------------------------
70: %\linespread{1.6}
71: \begin{document}
72:
73: \maketitle
74:
75: {\abstract\noindent
76: %
77: Choosing a simple class of flows, with characteristics that may be
78: present in the Earth's core, we study the ability to generate
79: a magnetic field when the flow is permitted to oscillate periodically
80: in time. The flow characteristics are parameterised by
81: $D$, representing a differential rotation,
82: $M$, a meridional circulation, and
83: $C$, a roll component characterising convective rolls.
84: %These parameters are subject to the
85: %condition that the total kinetic energy is constant.
86: The dynamo action of all solutions with fixed parameters
87: (steady flows)
88: is known from previous studies.
89: Dynamo action is sensitive to these flow parameters
90: and fails spectacularly for much of the
91: parameter space where magnetic flux is concentrated into small
92: regions, leading to high diffusion.
93: In addition, steady flows generate only steady
94: or regularly reversing oscillatory fields
95: and cannot therefore reproduce
96: irregular geomagnetic-type reversal behaviour.
97: %
98: Oscillations of the flow are introduced by varying the flow
99: parameters in time, defining a closed orbit in the space $(D,M)$.
100: When the frequency of the oscillation is small,
101: the net growth rate of the magnetic field over one period
102: approaches the average of the growth rates for steady flows
103: along the orbit.
104: %
105: At increased frequency time-dependence appears to smooth out flux
106: concentrations, often enhancing dynamo action.
107: Dynamo action can be impaired, however, when flux
108: concentrations of opposite signs occur close together as
109: smoothing destroys the flux by cancellation.
110: %
111: It is possible to produce geomagnetic-type reversals by making the
112: orbit stray into a region where the steady flows generate
113: oscillatory fields. In this case, however,
114: dynamo action was not found to be
115: enhanced by the time-dependence.
116: %
117: A novel approach is taken to solving the time-dependent eigenvalue
118: problem, where by combining Floquet theory with
119: a matrix-free Krylov-subspace method we
120: avoid large memory requirements for storing the matrix required by
121: the standard approach.
122: \\ \\
123: {\it Keywords:} Kinematic dynamos; Time dependent stability;
124: Geomagnetism; Floquet theory; Eigenvalue problems.
125: }
126:
127:
128: %---------------------------------------------------------------------------
129: %---------------------------------------------------------------------------
130: \section{Introduction}
131: %
132: Unlike the Sun or Jupiter, the Earth's dynamo
133: runs on a tight heat budget and may therefore rely on a
134: significant large-scale component to the flow.
135: In addition to the dominant dipole component of the geomagnetic
136: field, there is some observational evidence to suggest the
137: field contains a persistent non-axisymmetric component
138: \citep{gubbins93}.
139: Both may be indicative of
140: a steady component to the underlying core flow.
141: This prompted a kinematic study of dynamo action from a class
142: of large-scale steady candidate core-flows
143: (\citet{gubbins00a}, \citet{gubbins00b}, \citet{gubbins02}
144: hereafter referred to as papers I--III).
145: Steady flows, particularly if they contain stagnation points,
146: tend to concentrate magnetic flux into small regions with large
147: energy loss due to diffusion.
148: Often, increasing the flow speed to overcome diffusive loss simply
149: results in more concentrated flux and faster decay.
150: On the other hand, chaotic flows appear to make better dynamos,
151: perhaps because the mixing properties of the flow prevents
152: permanent flux concentration and exponential separation of
153: neighbouring particles in the flow lead to stretching of the
154: magnetic field \citep{brummell98}.
155: The dynamo mechanisms of complicated chaotic flows
156: are difficult to understand. We are therefore motivated to first
157: study the effects of simple flows that fluctuate about a steady mean.
158: Although the Earth's dipole has persisted for a long time,
159: secular variation including excursions of the magnetic pole may
160: indicate that fluctuations of the large-scale flow are
161: present in the Earth's core.
162:
163: Kinematic theory ignores the nonlinearity of back-reaction by the
164: magnetic field on the flow,
165: and considers only the time evolution of the magnetic field
166: $\vec{b}$ as governed by the induction equation
167: \begin{equation}
168: \label{eq:indeqn}
169: \partial_t \vec{b} = \Rm \curl ( \vec{u} \cross \vec{b} )
170: + \laplace \vec{b} .
171: \end{equation}
172: The induction equation has been non-dimensionalised with the
173: timescale for magnetic diffusion, the length scale
174: $d$, the radius of the sphere,
175: and in this work the
176: velocity $\vec{u}$ is normalised such that the
177: magnetic Reynolds number $R_m$ is unity for a flow
178: of unit kinetic energy. For a given steady flow the induction
179: equation is linear in $\vec{b}$ and has eigenfunction
180: solutions of the form
181: $ %\begin{equation}
182: % \label{eq:eigfn}
183: \vec{b}(r,\theta,\phi;t) = \ex^{\sigma t}\,\vec{B}(r,\theta,\phi) .
184: $ %\end{equation}
185: Dynamo action is established if $\Re({\sigma})>0$.
186: This simple test is the major
187: advantage of the kinematic approach. The
188: alternative is to integrate in time the nonlinear problem
189: for both the velocity and magnetic field
190: until one is convinced the magnetic field will not ultimately decay;
191: this is expensive and the results can be uncertain.
192: %
193: The advantage remains, however, when the flow varies in time
194: but is periodic;
195: Floquet theory gives an eigenvalue problem for the growth rate.
196:
197: \citet{backus58} was first to show kinematic dynamo action by
198: a time-dependent flow.
199: His dynamo employed periods of stasis while high harmonics in the field
200: decayed, enabling him to establish convergence of the solution.
201: Time dependence
202: may even lead to dynamo action when no single snapshot of the flow can
203: generate magnetic field on its own.
204: Magnetic fields can grow during an initial transient period
205: under the influence of a
206: subcritical steady flow, the most familiar example being
207: the production of toroidal field from the action of differential rotation
208: on poloidal field.
209: If the induction equation were self-adjoint and its eigenfunctions
210: orthogonal it would be a simple matter to prove that all such transients
211: decay; it is the non-normal property that allows transients to grow.
212: The initial fields which optimise transient growth for flows in a sphere,
213: including one of the flows here, have been studied by \citet{livermore04}.
214: Unfortunately, if the flow is steady the field eventually dies away
215: and only the slowest decaying mode remains.
216: If the flow is permitted to be time dependent, however,
217: once a transient field associated with the initial flow has grown,
218: a change in the flow can encourage further growth.
219: In plane-layer flow it has been shown that by repeatedly switching
220: the orientation of the flow it is possible to take
221: advantage of these transients \citep{gog99},
222: and to find dynamo action where each flow in isolation does not
223: dynamo kinematically.
224:
225: Another reason to extend the studies to time-dependent flows
226: is that steady flows cannot
227: account for the irregularity of geomagnetic reversals.
228: The induction equation is linear with eigenfunctions $\vec{b}$ that
229: change with $t$ only in magnitude, when $\Im(\sigma)=0$,
230: or oscillatory solutions
231: that reverse polarity with fixed period $2\pi/\Im(\sigma)$.
232: Geomagnetic-type reversals require changes in the flow.
233: \citet{sarson99} described
234: irregular reversals that occurred in simulations with their
235: $2\frac1{2}$-dimensional model.
236: The mechanism could be interpreted kinematically, and reversals were
237: observed to occur when fluctuations in the flow lead to a
238: reduced meridional circulation.
239: More recently \citet{wicht04} studied reversals in a fully self-consistent
240: but quasi-periodic system.
241: The reversal mechanism they proposed also appeals largely to kinematic
242: principles and appears to reverse with approximately fixed
243: period even when nonlinearity through the Lorentz force is
244: omitted.
245:
246: %Periodic flows can be used to study geomagnetic-type reversals. Although
247: %they generate magnetic fields that are also periodic in time, the magnetic
248: %field can reverse in a time that is short compared with the period of
249: %the cycle, so the reversal process can be isolated from the longer-term
250: %cyclical behaviour.
251:
252: The class of steady flows explored in I--III was originally
253: prescribed by \citet{kumar75} and, with parameters chosen to mimic
254: flows near the limit of \citet{braginsky64}, was shown to be capable
255: of dynamo action. Dependence of the dynamo on a much wider
256: range of parameter values was later found in I.
257: The Kumar--Roberts flow is confined to the sphere of unit radius,
258: the exterior of which is assumed to be perfectly insulating.
259: Three components of the flow represent
260: a differential rotation, a meridional circulation and a
261: convective overturn,
262: \begin{equation}
263: \vec{u} =
264: \epsilon_0 \vec{t}_1^0 +
265: \epsilon_1 \vec{s}_2^0 +
266: ( \epsilon_2 \vec{s}_2^{2c} +
267: \epsilon_3 \vec{s}_2^{2s} ) .
268: \end{equation}
269: Following the nomenclature detailed in I,
270: the $\epsilon_i$ are constrained such that $\epsilon_2=\epsilon_3$ and
271: the kinetic energy of the flow is unity.
272: %\begin{equation}
273: % \int \vec{v}^2 \, {\mathrm d}V = \sum_i \alpha_i \, \epsilon_i = 1 .
274: %\end{equation}
275: The flow is parameterised by $(D,M)$-space, where
276: $D=D(\epsilon_0)$, $M=M(\epsilon_1)$ and $|D|+|M|\le 1$.
277: The parameters $D$ and $M$ are measures of the
278: differential rotation and meridional circulation respectively.
279:
280: For a steady forcing flow, writing
281: $\vec{b}(t)=\ex^{\sigma t}\vec{B}$, where $\vec{B}$ is independent of $t$,
282: \rf{eq:indeqn}
283: can be expressed as the eigenvalue problem
284: \begin{equation}
285: \label{eq:eigprobfnsteady}
286: \sigma \, \vec{B} = {\mathcal F} \, \vec{B} .
287: \end{equation}
288: In paper I dynamo action was established for
289: approximately half the $(D,M)$-space
290: (Fig.~\ref{fig:diamond}).
291: \citet{sarson96} and III
292: found a number of oscillatory solutions for steady flows
293: in a region which corresponds to the dynamo wave solutions of the
294: $\alpha\omega$ equations in the Braginsky limit ---
295: $|D|\to 1$ in a manner such that $1-|D|^2=c\,|M|$ where $c$ is a constant.
296: The oscillatory region was found in I
297: to extend only for a very narrow range in $M$, shown schematically
298: in Fig.~\ref{fig:diamond}. The majority of solutions are steady.
299: Given the narrow range for $M$, it is apparent that
300: only a small degree of meridional circulation is required to stabilise
301: the field to steady solutions. On the other hand, the existence of
302: oscillatory solutions for low $M$ appears to be a fairly robust feature
303: as the range in $D$ for which they exist is large, and extends well beyond
304: the limit of Braginsky.
305:
306: In this work, the exploration above is extended to the
307: dynamo action of flows with $D=D(t)$ and $M=M(t)$ periodic in time,
308: with a given period $T$.
309: The induction equation \rf{eq:indeqn} can be written as
310: \begin{equation}
311: \label{eq:indop}
312: \pd{t} \vec{b} = {\mathcal F}(t) \, \vec{b},
313: \end{equation}
314: with periodic forcing
315: ${\mathcal F}(T+t) = {\mathcal F}(t)$.
316: It follows from Floquet's theorem (see \S\ref{sect:nummeth})
317: that solutions may be written in the form
318: $\vec{b}(T+t)=\ex^{\sigma_1 T}\,\vec{b}(t)$
319: where the real part of $\sigma_1$ is the net growth rate over one
320: period. Setting $\vec{b}(t)=\ex^{\sigma_1 t}\,\vec{B}(t)$, so that
321: $\vec{B}(T+t)=\vec{B}(t)$,
322: substitution into \rf{eq:indop} defines the
323: eigenvalue problem for $\vec{B}$,
324: \begin{equation}
325: \label{eq:eignonst}
326: \sigma_1 \, \vec{B} = \left( {\mathcal F} - \pd{t} \right) \vec{B} .
327: \end{equation}
328: The critical magnetic Reynolds number for which the field is marginally
329: stable, $\Re(\sigma)=0$, is denoted $\Rm^c$.
330:
331: Both the steady and non-steady
332: eigenvalue problems \rf{eq:eigprobfnsteady} and \rf{eq:eignonst}
333: permit solutions for four linearly independent spatial
334: symmetries, axial dipole, axial quadrupole, equatorial dipole
335: and equatorial quadrupole. Symmetry selection in the steady case
336: was studied in II. Here only the geophysically interesting axial dipole
337: symmetry will be considered.
338: %---------------------------------------------------------------------------
339: %---------------------------------------------------------------------------
340: \section{Numerical method}
341: \label{sect:nummeth}
342: %
343:
344: Steady flows have been studied using extensions of the method first
345: developed by \citet{bullard54}. Toroidal and poloidal potentials for
346: the magnetic field are expanded in
347: spherical harmonics, with
348: truncation at degree $L$. A finite difference scheme is applied
349: on $N_r$ points in the radial dimension
350: leading to
351: the discretised eigenvalue problem
352: \begin{equation}
353: \label{eq:eigprobmat}
354: \sigma \, \vec{B} = \mat{E} \, \vec{B} .
355: \end{equation}
356: The matrix $\mat{E}$ has dimensions $N_rN_h\times N_rN_h$, where after
357: symmetry considerations the number of harmonics $N_h \sim \frac1{2} L^2$.
358: As the finite difference scheme only connects neighbouring points,
359: $\mat{E}$ is block banded where each block has size $N_h\times N_h$.
360: Eigenvectors are then
361: calculated by either by inverse iteration or by the Implicitly
362: Restarted Arnoldi Method (IRAM) on the inverse.
363: Due to the performance of both methods with respect to the
364: distribution of the eigenvalues, both operate on the
365: inverse and require the (banded) LU factorisation of $\mat{E}$.
366: Memory requirements scale like several times $N_h^2 N_r$, depending on
367: the stencil size of the finite difference approximation.
368: Solutions have generally been calculated with second order differences,
369: and $L$ not much larger than twenty. The storage requirement for the large
370: matrix is the limiting factor for the calculation.
371:
372: For the time-dependent eigenvalue problem \rf{eq:eignonst} with
373: the same spatial representation, applying
374: a Fourier expansion in time introduces at least another factor
375: $N_t$ to the storage requirements. This can be minimised by
376: permitting only sinusoidal forcings, but due to the structure of the
377: matrix memory requirements are prohibitive
378: with respect to calculation of the
379: LU factorisation (a few times $N_h^2 N_r^2 N_t$).
380: Storage is a significant difficulty in multiplying by the inverse
381: or in calculating the inverse of a suitable preconditioner for the
382: time-dependent problem.
383:
384: %Another possible method makes use of the eigenfunctions already calculated
385: %for steady flows. The solution is expanded in these eigenfunctions using
386: %time-dependent coefficients to be found from \rf{eq:eignonst}.
387: %The scheme is similar to Backus' (1958) use of piecewise-steady
388: %flows. Its efficiency depends on the number of modes needed for an accurate solution,
389: %since a full set of modes would require as much work and storage as
390: %Fourier expansion. It is likely to work well only for time dependent flows
391: %whose snapshots have similar eigenfunctions, restricting it to small changes
392: %in magnetic behaviour. Induction mode expansion was not pursued further.
393:
394: Instead we have adopted a method that does not require storage of the
395: matrix, which we call the matrix-free Krylov subspace method.
396: It is an adaptation of a
397: method used to find steady solutions of the Navier-Stokes equations
398: by \citet{edwards94}. Periodicity of the flow is incorporated in
399: the following manner \citep{verhulst96}.
400: Writing the discrete form of (\ref{eq:indop}) as
401: \begin{equation}
402: \label{eq:indopdesc}
403: \pd{t}\vec{b} = \mat{F}(t) \vec{b},
404: \end{equation}
405: the matrix $\mat{G}(t)$ satisfying
406: $ %\begin{equation}
407: \pd{t}\mat{G}(t) = \mat{F}(t)\,\mat{G}(t),
408: $ with $%\quad
409: \mat{G}(0) = \mat{I}, \,
410: $ %\end{equation}
411: is the fundamental matrix of the system (\ref{eq:indopdesc}).
412: Evolution of a starting solution is then given by
413: \begin{equation}
414: \label{eq:fundevol}
415: \vec{b}(t) = \mat{G}(t) \, \vec{b}(0) .
416: \end{equation}
417: For any $T$-periodic $\mat{F}(t)$, there exist matrices $\mat{P}(t)$ and
418: $\mat{E}$ such that the fundamental matrix can be written
419: \begin{equation}
420: \label{eq:floqthm}
421: \mat{G}(t) = \mat{P}(t) \, \ex^{\mat{E}\,t},
422: \end{equation}
423: where $\mat{E}$ is independent of $t$ and $\mat{P}(t)$
424: is %non-singular and
425: $T$-periodic
426: (Floquet's theorem).
427: %From (\ref{eq:fundevol}) and (\ref{eq:floqthm})
428: %Setting $t_0=0$ it
429: It follows immediately that
430: the change in the solution over one period is given by
431: \begin{equation}
432: \vec{b}(T) = \mat{G}(T) \, \vec{b}(0),
433: \qquad
434: \mat{G}(T) = \ex^{\mat{E}\,T} = \mat{A} .
435: \end{equation}
436: The stability of solutions to
437: (\ref{eq:indopdesc}) is determined by the eigenvalues $\lambda$
438: of the constant matrix $\mat{A}$.
439: If $\vec{b}(0)$ is an eigenvector of $\mat{A}$
440: with eigenvalue $\lambda=\ex^{\sigma_1 T}$,
441: %and $\vec{b}(t)=\mat{G}(t)\,\vec{b}(0)$,
442: we find that
443: $\vec{b}(T+t)=\ex^{\sigma_1 T}\,\vec{b}(t)$ for any $t$.
444: The real part of the Floquet exponents $\sigma_1$
445: correspond to growth rates of the solutions.
446: Although $\mat{A}$ is unknown,
447: from (\ref{eq:fundevol}) we see that
448: the effect of multiplying by $\mat{A}$
449: is equivalent to the result of timestepping through one period.
450: Therefore we do not have to calculate and store $\mat{A}$ explicitly.
451: Note that for a steady forcing $\mat{F}$, the period
452: $T$ can be chosen arbitrarily.
453:
454:
455: The eigenvalue problem for $\mat{A}$
456: is well suited to the Arnoldi process \citep{arnoldi51},
457: which tends to pick out eigenvalues isolated in the complex plane.
458: The many %non-physical
459: decaying modes have $\lambda$ clustered about the origin, marginal
460: modes correspond to $|\lambda|$ close to unity.
461: %The matrix $\mat{A}$ does not need to be known explicitly and is
462: %accessed only once per iteration.
463: %
464: At each iteration we add to the Krylov-subspace given by
465: $\,\mathrm{span}\{\vec{b}, \mat{A}\vec{b},...,\mat{A}^{k-1}\vec{b}\}$
466: which we hope contains our solutions.
467: In exact arithmetic the $k^{\mathrm{th}}$ Krylov subspace is
468: equivalent to $\mathrm{span}\{\vec{b}_1,...,\vec{b}_k\}$
469: where the basis vectors $\vec{b}_k$ are obtained from the Arnoldi method.
470: Numerically the latter set is better suited to span the space.
471: %
472: The Arnoldi process is summarised as follows: (1)
473: Take a suitable normalised initial basis vector
474: $\vec{b}_1=\vec{b}/\|\vec{b}\|_2$.
475: (2)
476: At the $k^{\mathrm{th}}$ iteration
477: evaluate (timestep)
478: %$\vec{b}_k$ is timestepped,
479: $\tilde{\vec{b}}_{k+1} = \mat{A}\,\vec{b}_k$.
480: (3) The result $\tilde{\vec{b}}_{k+1}$ is then
481: orthogonalised against previous vectors in the basis by the
482: modified Gram--Schmidt method:
483: \[%begin{*equation}
484: h_{jk} = \langle \tilde{\vec{b}}_{k+1} , \vec{b}_j \rangle,
485: \quad
486: \tilde{\vec{b}}_{k+1} := \tilde{\vec{b}}_{k+1} - h_{jk}\vec{b}_j;
487: \quad
488: j \le k .
489: \]%end{*equation}
490: (4) Setting $h_{k+1,k} = \| \tilde{\vec{b}}_{k+1} \|_2$, the process
491: continues from (2) with the next basis vector
492: $\vec{b}_{k+1} = \tilde{\vec{b}}_{k+1} / h_{k+1,k}$.
493: %
494: %In exact arithmetic the $k^{\mathrm{th}}$ Krylov subspace is given
495: %by
496: %$\,\mathrm{span}\{\vec{b}, \mat{A}\vec{b},...,\mat{A}^{k-1}\vec{b}\}
497: %=\mathrm{span}\{\vec{b}_1,...,\vec{b}_k\}$.
498: %Numerically the latter is better suited to span the space,
499: % which we hope contains our solutions.
500: Construct $\mat{Q}_k = \left[ \vec{b}_1,...,\vec{b}_k \right]$
501: and $\mat{H}_k = \left[ h_{jm} \right ]_{j,m\le k}$.
502: From steps (3) and (4) we expect $h_{k+1,k}\to 0$.
503: In this case,
504: looking carefully at the steps above,
505: the results of the Arnoldi process %after $k$ steps
506: can be written
507: $\mat{A}\mat{Q}_k = \mat{Q}_k \mat{H}_k$.
508: Multiplying on the right by
509: eigenvectors $\vec{x}$ of $\mat{H}_k$ we find that they are related
510: to those of $\mat{A}$ by
511: $\vec{b} = \mat{Q}_k \vec{x}$.
512: For non-zero $h_{k+1,k}$ eigenvectors have
513: residual
514: $\|\mat{A}\,\vec{b} - \lambda\vec{b}\|_2 = |h_{k+1,k}| |x_k|$,
515: where $x_k$ is the last element of the $k$-vector $\vec{x}$.
516: Thus, at each iteration eigenvalues $\lambda$
517: of $\mat{H}_k$ are approximate eigenvalues of $\mat{A}$.
518:
519: %Now, letting $\mat{Q}_k = \left[ \vec{b}_1,...,\vec{b}_k \right]$, and
520: %$\mat{H}_k = \left[ h_{jm} \right ]_{j,m\le k}$ it turns out that
521: %eigenvalues of $\mat{H}_k$ are approximate eigenvalues of $\mat{A}$.
522: %The corresponding eigenvectors $\vec{x}$ of $\mat{H}_k$ are related
523: %to those of $\mat{A}$ by
524: %$\vec{b} = \mat{Q}_k \vec{x}$.
525: %The residual
526: %$\|\mat{A}\,\vec{b} - \lambda\vec{b}\|_2 = |h_{k+1,k}| |x_k|$,
527: %where $x_k$ is the last element of the $k$-vector $\vec{x}$.
528:
529: In practice the residual $|h_{k+1,k}| |x_k|$
530: tends to overestimate the error, and in our
531: calculations the Arnoldi process is stopped when the largest
532: eigenvalues of $\mat{H}_k$
533: are sufficiently converged. The number of iterations required is
534: typically of order 100 or less, and so the eigenvalues of the
535: small matrix $\mat{H}_k$ can be
536: cheaply calculated by the QR algorithm. The memory required to store
537: the basis vectors scales like $N_h N_r k$.
538: It is possible to restart the Arnoldi process with a more suitable
539: starting vector obtained from the process so far, but without
540: completely restarting the new process from scratch.
541: This implicit restarting
542: allows further reduction of memory requirements by reducing the
543: number of basis vectors at each restart \citep{lehoucq98}.
544: With $k$ small, restarting
545: was not found to be necessary, however.
546: The stencil of the finite
547: difference scheme does not explicitly affect the storage requirements.
548: Basis vectors were therefore timestepped with a fourth order finite
549: difference scheme.
550: Timestepping was performed with the benchmarked code of
551: Gibbons, Jones and Zhang \citep{christensen01}.
552: %For little adaptation of the timestepping code
553: %the Arnoldi process offers accelerated convergence over
554: %direct timestepping alone, and allows calculation of more than one
555: %eigenvalue.
556:
557: Another advantage of the matrix-free method
558: is that, given a timestepping code, only a few extra lines of
559: code are required to incorporate the Arnoldi process, therefore leaving
560: significantly less room for error. The code was verified by comparison
561: with the matrix method used in I--III
562: for the steady problem, adjustments for the periodic case in our
563: matrix-free method are minimal.
564: %
565: Table \ref{tbl:grtest}({\it a})
566: shows the leading two eigenvalues for the
567: of the steady Kumar--Roberts flow,
568: $(D,M)=(0.98354915,0.0001632689)$,
569: at $\Rm=1000$ calculated using the method in I--III.
570: Table \ref{tbl:grtest}({\it b}) shows the same eigenvalues calculated
571: using our method.
572: %
573: The higher order radial differences used in the
574: timestepping code leads to more rapid convergence with $N_r$.
575: % and affects memory requirements very little relative to the matrix method.
576: %
577: Table \ref{tbl:grtest}({\it c}) shows that incorporating the Arnoldi
578: method accelerates convergence relative to timestepping alone
579: (and calculation of more than one eigenvalue is possible).
580: The same starting vector was used for both calculations.
581: The period $T$ can be chosen arbitrarily for the
582: steady flow case, but if chosen too small more iterations are required
583: and therefore more basis vectors must be stored. For these calculations
584: we set $T=0.001$.
585:
586: As the structure of the eigenfunctions varies with $D$ and $M$,
587: so does the convergence with $N_r$ and $L$.
588: For most of the following analysis a radial resolution of $N_r=50$
589: and a spectral truncation of $L=18$ was adopted.
590: Checks at higher resolutions were calculated where growth rates were found
591: to vary rapidly with the parameters.
592:
593:
594: %---------------------------------------------------------------------------
595: %---------------------------------------------------------------------------
596: \section{Results}
597: %
598: Periodic flows are defined by a closed orbit in $(D,M)$-space.
599: We restrict ourselves to simple sinusoidal variations in time
600: with a single frequency $\omega$:
601: %
602: %Changes in the growth rate are a clear indicator of whether changes
603: %in the flow are beneficial or not to its dynamo properties.
604: %For steady flows, small parameter changes can lead to significant
605: %differences in the generated field. Here the flow is changed
606: %continuously with smooth variations described by
607: \begin{eqnarray}
608: D(t) & = & D_0 + A_D \cos(\omega t), \nonumber \\
609: M(t) & = & M_0 + A_M \sin(\omega t).
610: \end{eqnarray}
611: It is the aim of this section to assess how the amplitude of variations
612: $A_{D,M}$ and frequency $\omega$ affect the growth rates and therefore
613: the dynamo action.
614: %
615: \subsection{Magnetic growth rates for time-varying flows}
616: %
617: Figure \ref{fig:grAamp} shows growth rates for different amplitudes of
618: variations about the point $(D_0,M_0)=(0.25,-0.14)$,
619: marked A in Fig.~\ref{fig:diamond}, which lies on a
620: line of minimum $\Rm$ extending from the Braginsky limit point $(1,0)$
621: (see I, Table~5). The majority of neighbouring points
622: have lower growth rates for the given value of $\Rm=87$.
623: Figure \ref{fig:grAamp} shows that the effect of oscillations of the
624: flow on the growth rates is more pronounced with increased
625: oscillation amplitude.
626: %
627: %growth rate decreases with
628: %increasing amplitude of oscillation, $A$: this is because the
629: %the steady flows far from the central point $(D_0,M_0)$ have lower
630: %growth rates for the same magnetic Reynolds number.
631:
632: For a steady flow, given any small real number $\varepsilon>0$
633: there exists a finite
634: time $t$ such that transients are reduced so that
635: $|\sigma-\sigma(t)|<\varepsilon$, where $\sigma$ is the growth
636: rate corresponding to the steady flow at each point on the orbit,
637: and $\sigma(t)$ is the growth rate of an
638: arbitrary initial field as it adjusts to the flow.
639: Provided that growth rates are
640: piecewise continuous (for example $D(t)$ could be discontinuous but
641: periodic, see Backus 1958),
642: a frequency $\omega$ can always be selected low
643: enough such that net growth rate over the cycle is close to
644: the mean $\bar{\sigma}$ of those on the orbit.
645: The limit $\sigma_1(\omega)\to\bar{\sigma}$ as $\omega\to 0$
646: is observed in the numerical
647: calculations. What is more interesting, however, is that with
648: finite $\omega$ the dynamo can do much better than this mean,
649: $\sigma_1(\omega)>\bar{\sigma}$,
650: as seen in most figures for the growth rate.
651: %as for $A=0.1$ and
652: %$\ln\omega\approx 5$ in Figure \ref{fig:grAamp}.
653: Figure \ref{fig:grARm} shows that the effect increases
654: with $\Rm$ and that the peak occurs at a frequency $\omega$
655: that increases in proportion to $\Rm$.
656:
657: Meridional sections of the magnetic field for this flow are plotted in
658: Fig.\ \ref{fig:AmerBph90}. The two times correspond closely to the
659: points on the orbit which have the maximum (upper row) and
660: minimum growth rates for steady flows ($\omega\to 0$).
661: The structure of these eigenfunctions is similar;
662: regions of $B_\phi$ are generally well separated in space.
663: The dissipation for these fields
664: is larger for the lower panel where fields of opposite
665: sign are squeezed towards the equator.
666: For non-zero $\omega$ the location of the flux changes over
667: the cycle, and at $\log\omega=2.6$ the field represents a smoothed version
668: of the two eigenfunctions. Fewer small-scale features are present and
669: the flow performs well as a dynamo (see Fig. \ref{fig:grARm}).
670: %
671: At $\Rm=150$ the peak frequency for $\sigma_1$
672: is $\log\omega\approx 2.6$;
673: taking $T\sim (\delta/d)^2$
674: as an approximate timescale for diffusion,
675: if $\delta\sim d/8$ is an approximate length scale for the small
676: scale features of the eigenfunctions,
677: we find that the timescale for diffusion and for the
678: peak flow oscillation coincide.
679: The magnetic field is then smoothed effectively.
680: %From the peak frequency $\log\omega\approx 2.6$ at this value of $\Rm=150$,
681: %taking $T\sim (\delta/d)^2$
682: %as an approximate timescale for diffusion,
683: %we find $\delta\sim d/8$
684: %as an approximate length scale for
685: %the field configurations, not incompatible with the small-scale features
686: %of eigenfunctions for the steady flows.
687: Above this frequency the growth rate decreases again as the field is
688: unable to respond to rapid changes in the flow. Spatial smoothing is
689: lost and the field is close to steady --- plots at the two times
690: for $\log\omega=2.9$ are almost identical.
691: The field responds as though to a steady flow, retaining the
692: stronger (smaller-scale) features from each eigenfunction.
693:
694: Figure \ref{fig:grBRm} shows growth rates for an orbit about the point
695: $(0.5,-0.15)$, marked B in Fig.~\ref{fig:diamond}, which
696: lies on the lower boundary of the region of successful steady dynamos.
697: $D$ remains constant and $M$ varies to carry the flow outside the
698: dynamo region. The time-dependent flow produces a positive effect on
699: the growth rate. The spatial structure of the eigenvectors on this orbit is
700: similar to that of the previous point considered, with well separated regions
701: of positive and negative azimuthal field. For $R_m=150$ there is dynamo
702: action only for frequencies $\log\omega\approx 2.6$, and the average
703: growth rate around the orbit is negative.
704:
705: We now describe a case where meridional circulation is greater than
706: differential rotation, $(D,M)=(-0.10,-0.45)$,
707: marked C in Fig.~\ref{fig:diamond}.
708: This point is close to where the critical magnetic
709: Reynolds number for steady flows is at a global minimum.
710: Once again time dependence of the flow enhances the growth rate,
711: Fig.~\ref{fig:grGRm}. A small rise can be seen in the
712: growth rate, although less significant relative to the increase
713: associated with an increase in the magnetic Reynolds number.
714: %Figure~\ref{fig:GmerBph90} shows that eigenfunctions on the path
715: %are similar to each other.
716: Being close to the point where $\Rm^c$ is a minimum,
717: the fields are already relatively large scale and the smoothing
718: effect of oscillations therefore has a smaller effect on the growth rate.
719:
720: Growth of the time-dependent solutions is not always found to be
721: better than the mean of the growth rates on the orbit.
722: Figure \ref{fig:grCRm} shows the case $(D,M)=(0.50,0.11)$,
723: marked D in Fig.~\ref{fig:diamond},
724: where oscillations in the flow are initially
725: damaging to the dynamo.
726: If the oscillations are sufficiently rapid, however, the dynamo is
727: again able to perform better. Meridional sections are shown in
728: Fig.\ \ref{fig:CmerBph00}.
729: The regions of strongest flux are located very
730: close together on the equator.
731: Over the cycle radial shifts of the clover-leaf pattern of flux causes
732: considerable overlap of opposite signs. This occurs mostly
733: towards the outer edge of the equatorial region.
734: As the flow oscillates, in the lower plot for
735: $\log\omega=2.13$ it can be seen that there is cancellation of flux in the
736: outer region. Some flux remains at the other regions where the signs
737: for the two eigenfunctions do correlate. This cancellation of fluxes
738: over the cycle leads to reduced growth rates. When the oscillation is
739: much faster, however, the dynamo does not have time to generate
740: flux of opposing sign. The field is more steady for $\log\omega=2.60$
741: and has a larger growth rate.
742: %[[[* Possibly plot $|J|^2$, chk increased at boundries
743: %*]]]
744:
745: The radial field in the above has little structure of interest.
746: It is concentrated mainly on the axis with opposite sign in each
747: hemisphere (see I, Fig. 9{\it a,b}).
748: The structure does not change appreciably over the cycles.
749: In strength, however, it is observed to wax and wane.
750:
751: %----------
752: %
753: %faster osc -> less small scale struture -> steeper spectrum?
754: %steady spectra inside/out -> flatter out.
755: %
756: %opendx plots - field mostly in interior - no surface plots?
757: %
758: %does the (steay) field structure vary significantly? Perhaps it is more
759: %similar at the extremes, but different in the middle (on the bdry) which
760: %may explain the initial decrease.
761:
762:
763:
764: \subsection{Dynamo wave solutions}
765:
766: Meridional circulation has been seen to play a key part in reversals.
767: \citet{sarson99} have studied a system in which irregular reversals
768: are linked to a drop in meridional circulation, leading to a preference
769: for oscillatory fields.
770: \citet{wicht04} have recently studied a reversal mechanism that involves
771: an advection of reversed flux by a large-scale S1 flow.
772: Reversals occur quasi-periodically in their model.
773: This behaviour may be related to the dynamo wave solutions obtained in
774: III. The oscillation has the form of a dynamo wave in which flux migrates
775: %from pole to equator
776: along the longitudes defined by the downwellings of the
777: convective parts of the flow, which could partially explain the observed
778: tendency for virtual geomagnetic poles to track around the Pacific during
779: polarity transition \citep{gubbins94}. The steady flow model can only oscillate
780: periodically, but we can construct a more geophysically realistic reversal by choosing a
781: time-dependent flow that traces an orbit in $(D,M)$-space that strays into the
782: dynamo wave region, depicted by the line E in Fig.~\ref{fig:diamond}, for
783: a fraction $f$ of its period. The field behaviour will depend on the frequency of the
784: dynamo wave, $\omega_D$ and the time spent by the flow in the oscillatory regime.
785: If $\omega\gg f\omega_D$ the flow will only spend a brief time in the oscillatory
786: regime and we expect only a minor change in the magnetic field. If $\omega\ll f\omega_D$
787: the flow spends a long time in the oscillatory regime and we expect the field
788: to oscillate several times before becoming steady again. The interesting
789: case is when $\omega\approx f\omega_D$, when the field may only have time
790: to oscillate for one or a half cycle, producing an excursion or a reversal
791: respectively.
792: %These expectations have been confirmed numerically by time stepping.
793:
794: We now explore reversal behaviour using periodic flows.
795: The structure of the eigenfunctions for steady flows changes appreciably as
796: $M$ crosses zero. It was seen in the previous section that, where this
797: is the case, fluctuations are not necessarily good for the dynamo.
798: Instead, an orbit is chosen to enter the oscillatory range from the
799: negative side. Consider the flow defined
800: by fixed $D=0.7$ and $M$ varying sinusoidally between $-0.0140$ and $-0.0020$
801: (E in Fig.~\ref{fig:diamond}).
802: This orbit spends approximately one third of the time within the band of
803: oscillatory solutions reported in III,
804: which lies between $M=-0.0057$ and $-0.0010$.
805: The dynamo wave frequency for steady flows increases
806: with $R_m$ and appears to saturate
807: at about $\Im(\sigma)=17$ (see III, Figs~2,3);
808: it appears to be limited by the diffusion time.
809: Here, $\omega$ for the time-dependent flow must be chosen comparable
810: with this frequency to give a single reversal, a value which is too low to
811: assist the dynamo action significantly.
812:
813: %Oscillatory dipole solutions are also observed in the
814: %Kumar--Roberts class of flows when the meridional
815: %circulation is very small \citep{gubbins02}.
816:
817: %Figure \ref{fig:grDRm} shows real and imaginary parts of
818: %growth rates for steady flow at a point with low meridional
819: %circulation $(D,M)=(0.7,-0.004)$, marked
820: %D in Figure~\ref{fig:diamond} (see III, Figures~2,3).
821: %Over the range for $\Rm$ the leading mode is oscillatory.
822: %As $\Rm$ increases the reversal period $2\pi/\Im(\sigma)$ does not
823: %continue to decrease, but instead appears to be limited by the
824: %diffusion time.
825: %Comparing $\Im(\sigma)=10$ with peak $\omega$ for the
826: %previous fluctuating flows reversal times correspond to a rather
827: %long period $t\approx\pi/10$.
828: %For a flow that passes into the oscillatory regime for sufficient time
829: %for a field reversal,
830: %the flow oscillation itself is unlikely to improve the associated growth
831: %rates over the mean on the path.
832: %It is therefore
833: %not surprising that the energy of the dipolar field is observed to
834: %decrease significantly during reversals.
835:
836: %At $D=0.7$ and $\Rm=700$ the leading mode is oscillatory for a
837: %narrow range of $M$ between $-0.0057$ and $-0.0010$ \citep{gubbins02}.
838: %The imaginary part of $\sigma$ is approximately $10$ across the range.
839:
840: Figure \ref{fig:grDosc} shows the the growth rate for the time dependent
841: flows as a function of frequency $\omega$.
842: At $\Rm=700$, $\Im(\sigma)$ is approximately $10$
843: for steady flows in the oscillatory range.
844: Reversing solutions may be expected for
845: $Tf\gtrsim\pi/\Im(\sigma)$,
846: or equivalently $\omega\lesssim 20/3$, $\log\omega\lesssim 0.8$
847: as the time in the oscillatory range is approximately one third of the
848: cycle.
849: Reversing solutions (dashed curve) are observed for $\omega$
850: larger than the dynamo wave frequency, although
851: growth rates fall quickly when the
852: period of the flow oscillation is too short to be
853: compatible with the period of the
854: oscillatory solution. If the time within the oscillatory region
855: leads to only a half-complete reversal, the field exiting the
856: region bears little resemblance to the entering field,
857: which is much like the eigenfunction for these low $\omega$,
858: and therefore leads to reduced growth rates.
859: For greater $\omega$
860: the field does not spend sufficient time within the
861: oscillatory region to reverse (solid curve) and at higher $\omega$
862: the growth rates are increased.
863: %
864: The magnetic energy for a typical reversing solution is plotted
865: in Fig.\ \ref{fig:D20mnrg},
866: showing a %sharp
867: drop as the solution passes through the
868: oscillatory region where the reversal occurs.
869: It is possible to vary $M$ so that $f$ is less than a third. However,
870: for the reversing solution in Fig.\ \ref{fig:D20mnrg},
871: the smooth growth rate curve,
872: while outside the oscillatory region, indicates the
873: field quickly becomes independent of the period within the
874: oscillatory region, apart from in sign, due to the slow period of
875: the flow oscillation.
876:
877: The reversal sequence
878: for $B_r$ at the surface is shown in Fig.\ \ref{fig:D20surfBr}.
879: Patches of reversed flux appear at low latitudes, strengthen and
880: migrate polewards replacing the flux at high latitudes.
881: The reversal looks very similar to that reported by \citet{gubbins94}
882: who found the pole paths during the reversal correlate
883: well with the longitudes of these flux patches, located 180$^\circ$
884: apart.
885: If the frequency of the oscillation is too high the field is simply
886: disrupted by the short period in the oscillatory region, as seen in the
887: energy in Fig.\ \ref{fig:D20mnrg}.
888: This may lead to the type behaviour seen in geomagnetic excursions.
889: Figure \ref{fig:D20surfBrf} shows that reversed patches emerge but
890: have insufficient time to migrate polewards before dissipating.
891: They still still weaken the dipole, however.
892:
893: \section{Conclusions}
894: %
895: We have devised a new matrix-free Krylov subspace approach to solving
896: the time-dependent stability problem that is most effective in exploring
897: kinematic dynamo action of periodic flows.
898: It is computationally efficient, uses far less
899: storage than conventional methods, and requires rather little new coding
900: once time-step and eigenvalue routines are available.
901:
902: Time variation of the flow can sometimes, but not always, enhance dynamo
903: action. At low frequency the growth rate of the time dependent flow approaches
904: the average growth rate for the steady flows along the orbit. At moderate
905: frequency the time dependent flow can smooth out any concentrations of
906: magnetic flux generated by the component steady flows. This can produce
907: enhanced dynamo action (higher growth rate than the average) if the flux
908: concentrations are isolated and of one sign. Dynamo action is possible
909: at certain frequencies even when the average growth rate for steady flows
910: around the cycle is negative and the cycle contains mainly steady
911: flows that do not generate magnetic field. The growth rate appears to
912: be capped by the highest growth rate of any steady flow on the cycle.
913: At high frequency the magnetic field does not have time to adjust to time changes
914: in the flow and becomes almost stationary.
915:
916: Time variation does not always enhance dynamo action. When the generated field
917: has flux concentrations of different signs close together, the smoothing
918: effect tends to destroy flux.
919: The dynamo enhancement for these large scale flows is not as dramatic as those
920: reported by \citet{gog99}. Flows with similar eigenfunctions tend to have
921: similar growth rates at the same $R_m$, limiting the effect of the time
922: variation. Where the eigenfunction changes dramatically dynamo action
923: is usually impaired unless the period of the flow is short compared with
924: the diffusion time.
925: %It is however possible to change the flow slightly but get a very different
926: %eigenfunction. This does not tend to be helpful for kinematic
927: %dynamo action, although is not necessarily the case if the oscillations
928: %occur on a timescale very much smaller than the diffusion timescale.
929:
930: %Time dependence results in improved growth rates if the field does
931: %not have regions of opposing flux located very close together.
932: %Although generally capped by the largest growth rate, time
933: %dependence can produce dynamo action where the average growth rate
934: %of flows on the parameter path is negative.
935:
936: Solutions have been found that reverse when $M$ is low and are
937: associated with the steady-flow
938: oscillatory solutions found in II.
939: This result is in common with the reversals studied by
940: \citet{sarson99}, which occur irregularly due to a
941: drop in meridional circulation.
942: %Given there is strictly
943: %no flow across the equator in these flows it is difficult to
944: %communicate reversed flux between the northern and southern hemispheres
945: %in these calculations.
946: %An internal mechanism is required within each hemisphere and the
947: %range in $M$ for which this is the leading solution is narrow.
948: Although flows in their calculations
949: are also predominantly equatorial antisymmetric,
950: a small but increased flow across the equator is observed
951: during a reversal. It is unclear that this results in sufficient
952: advection of flux to influence their reversal mechanism.
953: %
954: A large circulation exterior to the tangent cylinder is required
955: in the reversal mechanism studied by \citet{wicht04}.
956: It is needed to transport reversed flux originating
957: from plumes that protrude the tangent cylinder.
958: %Advection of flux by meridional circulation certainly does play
959: %an important part in the reversal mechanism studied by \citet{wicht04}.
960: %Indeed, the circulation is required to be large and may allow
961: %reversals to occur on a shorter timescale.
962: However, it is difficult
963: to decipher what part fluctuations would play in their model, and in
964: particular to what degree this would affect the quasi-regularity of
965: their reversals.
966: %Given the limited number of configurations for our class of flows,
967: %it seems unlikely that for any real flow the range of $M$ for which
968: %oscillatory solutions occur would be so narrow as in our calculations.
969: %
970: We have shown that for a reversal to occur
971: the drop in $M$ must persist long enough for the field to reverse,
972: which for this class of flows this is approximately a tenth of a
973: diffusion time, or approximately the dipole decay time.
974: A significant drop in magnetic energy is observed
975: during the reversal. This arises because of the change in eigenfunction
976: between the steady and oscillatory modes of the steady solutions.
977: %It is much larger than the factor of 10--100 suggested for the drop in
978: %dipole energy in geomagnetic reversals. \\
979: %
980: %While oscillations have been observed to be damaging to the dynamo,
981: %For the majority of cases considrered oscillations in the flow
982: %are beneficial to the dynamo.
983: %The dynamo has the opportunity to behave like the better
984: %performing dynamos in the nearby space of flows.
985: %
986:
987:
988: {\bf Acknowledgements} \\
989: This work was supported by NERC grant GR3/12825. We thank Johannes Wicht
990: and an anonymous referee for useful comments, and
991: Dr SJ Gibbons for advice on using his time-step code.
992:
993: %========================================================================
994: \bibliographystyle{chicago}
995: %\bibliography{./earth}
996:
997: \begin{thebibliography}{}
998:
999: \bibitem[\protect\citeauthoryear{Arnoldi}{Arnoldi}{1951}]{arnoldi51}
1000: Arnoldi, W.~E. (1951).
1001: \newblock The principle of minimised iterations in the solution of the matrix
1002: eigenvalue problem.
1003: \newblock {\em Q. J. Appl. Math.\/}~{\em 9}, 17--29.
1004:
1005: \bibitem[\protect\citeauthoryear{Backus}{Backus}{1958}]{backus58}
1006: Backus, G. (1958).
1007: \newblock A class of self-sustaining dissipative spherical dynamos.
1008: \newblock {\em Annals of Physics\/}~{\em 4}, 372--447.
1009:
1010: \bibitem[\protect\citeauthoryear{Braginsky}{Braginsky}{1964}]{braginsky64}
1011: Braginsky, S.~I. (1964).
1012: \newblock Kinematic models of the {E}arth's hydrodynamic dynamo.
1013: \newblock {\em Geomagnetism and Aeronomy\/}~{\em 4}, 572--583.
1014:
1015: \bibitem[\protect\citeauthoryear{Brummell, F., and Tobias}{Brummell
1016: et~al.}{1998}]{brummell98}
1017: Brummell, N.~H., C.~F., and S.~M. Tobias (1998).
1018: \newblock Linear and nonlinear dynamo action.
1019: \newblock {\em Physics Letters\/}~{\em 249}, 437--442.
1020:
1021: \bibitem[\protect\citeauthoryear{Bullard and Gellman}{Bullard and
1022: Gellman}{1954}]{bullard54}
1023: Bullard, G.~E. and H.~Gellman (1954).
1024: \newblock Homogeneous dynamos and terrestrial magnetism.
1025: \newblock {\em Phil. Trans. R. Soc. Lond.\/}~{\em 250}, 543--585.
1026:
1027: \bibitem[\protect\citeauthoryear{Christensen, Aubert, Cardin, Dormy, Gibbons,
1028: Glatzmaier, Grote, Honkura, Jones, Kono, Matsushima, Sakuraba, Takahashi,
1029: Tilgner, Wicht, and Zhang}{Christensen et~al.}{2001}]{christensen01}
1030: Christensen, U.~R., J.~Aubert, P.~Cardin, E.~Dormy, S.~Gibbons, G.~A.
1031: Glatzmaier, E.~Grote, Y.~Honkura, C.~Jones, M.~Kono, M.~Matsushima,
1032: A.~Sakuraba, F.~Takahashi, A.~Tilgner, J.~Wicht, and K.~Zhang (2001).
1033: \newblock A numerical dynamo benchmark.
1034: \newblock {\em Phys. Earth Planet. Int.\/}~{\em 128}, 25--34.
1035:
1036: \bibitem[\protect\citeauthoryear{Edwards, Tuckerman, Friesner, and
1037: Sorensen}{Edwards et~al.}{1994}]{edwards94}
1038: Edwards, W.~S., L.~S. Tuckerman, R.~A. Friesner, and D.~C. Sorensen (1994).
1039: \newblock {K}rylov methods for the incompressible {N}avier--{S}tokes equations.
1040: \newblock {\em J. Comput. Phys.\/}~{\em 110}, 82--102.
1041:
1042: \bibitem[\protect\citeauthoryear{Gog, Oprea, Proctor, and Rucklidge}{Gog
1043: et~al.}{1999}]{gog99}
1044: Gog, J.~R., I.~Oprea, M.~R.~E. Proctor, and A.~M. Rucklidge (1999).
1045: \newblock Destabilisation by noise of transverse perturbations to hetroclinic
1046: cycles: a simple model and an example from dynamo theory.
1047: \newblock {\em Proc. R. Soc. Lond.\/}~{\em 455}, 4205--4222.
1048:
1049: \bibitem[\protect\citeauthoryear{Gubbins, Barber, Gibbons, and Love}{Gubbins
1050: et~al.}{2000a}]{gubbins00a}
1051: Gubbins, D., C.~N. Barber, S.~Gibbons, and J.~J. Love (2000a).
1052: \newblock Kinematic dynamo action in a sphere. {I} - {E}ffects of differential
1053: rotation and meridional circulation on solutions with axial dipole symmetry.
1054: \newblock {\em Proc. R. Soc. Lond.\/}~{\em 456}, 1333--1353.
1055:
1056: \bibitem[\protect\citeauthoryear{Gubbins, Barber, Gibbons, and Love}{Gubbins
1057: et~al.}{2000b}]{gubbins00b}
1058: Gubbins, D., C.~N. Barber, S.~Gibbons, and J.~J. Love (2000b).
1059: \newblock Kinematic dynamo action in a sphere. {II} - {S}ymmetry selection.
1060: \newblock {\em Proc. R. Soc. Lond.\/}~{\em 456}, 1669--1683.
1061:
1062: \bibitem[\protect\citeauthoryear{Gubbins and Gibbons}{Gubbins and
1063: Gibbons}{2002}]{gubbins02}
1064: Gubbins, D. and S.~Gibbons (2002).
1065: \newblock Three-dimensional dynamo waves in a sphere.
1066: \newblock {\em Geophys. Astrophys. Fluid Dynamics\/}~{\em 96}, 481--498.
1067:
1068: \bibitem[\protect\citeauthoryear{Gubbins and Kelly}{Gubbins and
1069: Kelly}{1993}]{gubbins93}
1070: Gubbins, D. and P.~Kelly (1993).
1071: \newblock Persistent patterns in the geomagnetic field over the past 2.5{M}yr.
1072: \newblock {\em Nature\/}~{\em 365}, 829--832.
1073:
1074: \bibitem[\protect\citeauthoryear{Gubbins and Sarson}{Gubbins and
1075: Sarson}{1994}]{gubbins94}
1076: Gubbins, D. and G.~R. Sarson (1994).
1077: \newblock Geomagnetic field morphologies from a kinematic dynamo model.
1078: \newblock {\em Nature\/}~{\em 368}, 51--55.
1079:
1080: \bibitem[\protect\citeauthoryear{Kumar and Roberts}{Kumar and
1081: Roberts}{1975}]{kumar75}
1082: Kumar, S. and P.~H. Roberts (1975).
1083: \newblock A three-dimensional kinematic dynamo.
1084: \newblock {\em Proc. R. Soc. Lond.\/}~{\em 244}, 235--258.
1085:
1086: \bibitem[\protect\citeauthoryear{Lehoucq, Sorensen, and Yang}{Lehoucq
1087: et~al.}{1998}]{lehoucq98}
1088: Lehoucq, R.~B., D.~C. Sorensen, and C.~Yang (1998).
1089: \newblock {\em {\sc Arpack} users guide: solution of large scale eigenvalue
1090: problems by implicitly restarted {A}rnoldi methods}.
1091: \newblock Philadelphia, PA: SIAM.
1092:
1093: \bibitem[\protect\citeauthoryear{Livermore and Jackson}{Livermore and
1094: Jackson}{2004}]{livermore04}
1095: Livermore, P.~W. and A.~Jackson (2004).
1096: \newblock On magnetic energy instability in spherical stationary flows.
1097: \newblock {\em Proc. R. Soc. Lond.\/}~{\em 460}, 1453--1476.
1098:
1099: \bibitem[\protect\citeauthoryear{Sarson and Gubbins}{Sarson and
1100: Gubbins}{1996}]{sarson96}
1101: Sarson, G.~R. and D.~Gubbins (1996).
1102: \newblock Three-dimensional kinematic dynamos dominated by strong differential
1103: rotation.
1104: \newblock {\em J. Fluid Mech.\/}~{\em 306}, 223--265.
1105:
1106: \bibitem[\protect\citeauthoryear{Sarson and Jones}{Sarson and
1107: Jones}{1999}]{sarson99}
1108: Sarson, G.~R. and C.~A. Jones (1999).
1109: \newblock A convection driven geodynamo reversal model.
1110: \newblock {\em Phys. Earth Planet. Int.\/}~{\em 111}, 3--20.
1111:
1112: \bibitem[\protect\citeauthoryear{Verhulst}{Verhulst}{1996}]{verhulst96}
1113: Verhulst, F. (1996).
1114: \newblock {\em Nonlinear differential equations and dynamical systems}.
1115: \newblock Springer, Berlin.
1116:
1117: \bibitem[\protect\citeauthoryear{Wicht and Olsen}{Wicht and
1118: Olsen}{2004}]{wicht04}
1119: Wicht, J. and P.~Olsen (2004).
1120: \newblock A detailed study of the polarity reversal mechanism in a numerical
1121: dynamo model.
1122: \newblock {\em Geochem. Geophys. Geosyst.\/}~{\em 5}, No. 3, Q03H10.
1123:
1124: \end{thebibliography}
1125:
1126: \newpage
1127: \linespread{1.0}
1128: %========================================================================
1129: \begin{figure}
1130: \centerline{\epsfig{file=diamond.eps,width=12.0cm}}
1131: \caption{ \label{fig:diamond}
1132: Regions of dynamo solutions for steady flows and axial-dipole symmetry for the
1133: flow parameterisation: $D$, differential rotation; $M$
1134: meridional circulation. The thick line running into the right
1135: hand apex contains oscillatory solutions related to the dynamo waves of
1136: the Braginsky limit. Other lines and circles describe periodic flows mentioned
1137: in the text.
1138: }
1139: \end{figure}
1140:
1141: %------------------------------------------------------------------------
1142: \begin{table}
1143: \begin{center}
1144: \begin{tabular}{rccc}
1145: %matrix
1146: ({\it a}) \\ & $N_r$=50 & 100 & 150 \\[2pt]
1147: \hline
1148: $L$=8 & 0.86410 & 0.95152 & 0.95593 \\
1149: 12 & 0.86705 & 0.95388 & 0.95826 \\
1150: 16 & 0.86717 & 0.95400 & 0.95838 \\[2pt]
1151: 8 & -36.137 & -36.406 & -36.420 \\
1152: 12 & -35.717 & -35.959 & -35.971 \\
1153: 16 & -35.704 & -35.954 & -35.958 \\[12pt]
1154: %tstep+arnoldi
1155: %T=0.001, dt=0.000001,0.0000007
1156: ({\it b}) \\ & $N_r$=25 & 50 & 75 \\[2pt]
1157: \hline
1158: $L$=8 & 0.98156 & 0.94976 & 0.95479 \\
1159: 12 & 0.99434 & 0.95211 & 0.95709 \\
1160: 16 & 0.99446 & 0.95223 & 0.95721 \\[2pt]
1161: 8 & -36.239 & -36.452 & -36.431 \\
1162: 12 & -35.788 & -36.005 & -35.984 \\
1163: 16 & -35.776 & -35.991 & -35.971
1164: \end{tabular}
1165: \,
1166: \begin{tabular}{rrr}
1167: %nr=75, l=16
1168: ({\it c}) \\%& $t_k=kT, T=0.01$
1169: \multicolumn{1}{c}{$k$} &
1170: \multicolumn{1}{c}{$\sigma_1$} &
1171: \multicolumn{1}{c}{$\sigma(t_k)$} \\[2pt]
1172: \hline
1173: 10 & -22.29762 & 81.56889 \\
1174: 20 & -6.12153 & 27.14633 \\
1175: 30 & 0.79446 & 4.61364 \\
1176: 40 & 2.24163 & -3.58609 \\
1177: 50 & 0.98918 & -3.14851 \\
1178: 60 & 0.91102 & 0.43870 \\
1179: 70 & 0.95688 & 3.19484 \\
1180: 80 & 0.95722 & 4.12172 \\
1181: 90 & 0.95721 & 3.87464 \\
1182: 100 & 0.95721 & 3.19592 \\
1183: 120 & & 1.91438 \\
1184: 150 & & 1.09479 \\
1185: 200 & & 0.98221 \\
1186: 250 & & 0.96445 \\
1187: 300 & & 0.95774 \\
1188: 400 & & 0.95735 \\
1189: 500 & & 0.95733
1190: \end{tabular}
1191: \end{center}
1192: \caption{ \label{tbl:grtest}
1193: Comparison of computed growth rates for the
1194: K--R flow, $(0.9834915,0.0001632689)$, at $\Rm=1000$;
1195: ({\it a}) leading two eigenvalues computed using the matrix
1196: and its LU factorisation;
1197: ({\it b}) eigenvalues calculated by the matrix-free method;
1198: ({\it c}) comparison with simple timestepping,
1199: $T=0.001$, $t_k=kT$ and $N_r=75$, $L=16$.
1200: }
1201: \end{table}
1202:
1203: %------------------------------------------------------------------------
1204: \begin{figure}
1205: \epsfig{figure=grAamp.eps}
1206: \caption{ \label{fig:grAamp}
1207: Growth rates $\sigma_1$ {\it vs}. $\omega$ for different amplitudes
1208: $A_D=A_M=A$ about the point $(0.25,-0.14)$.
1209: }
1210: \end{figure}
1211:
1212:
1213: %------------------------------------------------------------------------
1214: \begin{figure}
1215: \epsfig{figure=grARm.eps}
1216: \caption{ \label{fig:grARm}
1217: As Fig.\ \ref{fig:grAamp} for various $\Rm$; $A_D=A_M=0.1$,
1218: A in Fig.\ \ref{fig:diamond}.
1219: }
1220: \end{figure}
1221:
1222:
1223: %------------------------------------------------------------------------
1224: \begin{figure}
1225: \begin{tabular}{cccc}
1226: & $\omega\to 0$ & $\log\omega=2.6$ & $\log\omega=2.9$ \\
1227: $\frac{8}{16}T$ &
1228: \epsfig{figure=A8eigmerBph90.eps, scale=0.35} &
1229: \epsfig{figure=A8400merBph90s.eps, scale=0.35} &
1230: \epsfig{figure=A8800merBph90s.eps, scale=0.35} \\[8pt]
1231: $\frac{13}{16}T$ &
1232: \epsfig{figure=A13eigmerBph90.eps, scale=0.35} &
1233: \epsfig{figure=A13400merBph90.eps, scale=0.35} &
1234: \epsfig{figure=A13800merBph90.eps, scale=0.35}
1235: \end{tabular}
1236: \caption{ \label{fig:AmerBph90}
1237: Meridional sections, $B_\phi$, $\phi=\pi/2$;
1238: $\Rm=150$, $(D,M)=(0.25,-0.14)$, $A_D=A_M=0.1$.
1239: Eigenfunctions were calculated for steady flows at the appropriate
1240: point and are plotted with independent contour values;
1241: for $\omega\ne 0$, contours are plotted for the
1242: same values at different times.
1243: At $\log\omega=2.6$ %$\omega=400$
1244: the field structure looks smoothed. At $\log\omega=2.9$ %$\omega=800$
1245: the flow changes too quickly for the field to respond and
1246: the field is almost steady.
1247: }
1248: \end{figure}
1249:
1250:
1251: %------------------------------------------------------------------------
1252: \begin{figure}
1253: \epsfig{figure=grBRm.eps}
1254: \caption{ \label{fig:grBRm}
1255: Growth rates $\sigma_1$ {\it vs}.\
1256: $\omega$ for an orbit about the point
1257: $(0.5,-0.15)$; $A_D=0$, $A_M=0.1$ (Fig.\ \ref{fig:diamond}, B).
1258: }
1259: \end{figure}
1260:
1261:
1262: %------------------------------------------------------------------------
1263: \begin{figure}
1264: \epsfig{figure=grGRm.eps}
1265: \caption{ \label{fig:grGRm}
1266: Rising growth rates for an orbit about $(-0.1,-0.45)$;
1267: $A_D=A_M=0.1$ (Fig.\ \ref{fig:diamond}, C).
1268: }
1269: \end{figure}
1270:
1271:
1272: %------------------------------------------------------------------------
1273: %\begin{figure}
1274: % \begin{tabular}{cc}
1275: % \epsfig{figure=G2eigmerBph90.eps, scale=0.35} &
1276: % \epsfig{figure=G10eigmerBph90.eps, scale=0.35}
1277: % \end{tabular}
1278: % \caption{ \label{fig:GmerBph90}
1279: % Meridional sections, $B_\phi$, $\phi=\pi/2$;
1280: % $\Rm=75$, $(D,M)=(-0.1,0.45)$ near the global minimum for
1281: % $\Rm^c$. The field is generally large scale.
1282: % $A_D=A_M=0.1$;
1283: % $t=\frac{2}{16}\,T$, $\frac{10}{16}\,T$
1284: % corresponding approximately to $\max\sigma$, $\min\sigma$.
1285: % }
1286: %\end{figure}
1287: %
1288: %------------------------------------------------------------------------
1289: \begin{figure}
1290: \epsfig{figure=grCRm.eps}
1291: \caption{ \label{fig:grCRm}
1292: Growth rates for increasing $\omega$ about the point $(0.5,0.11)$;
1293: $A_D=0$, $A_M=0.1$ (Fig.\ \ref{fig:diamond}, D).
1294: }
1295: \end{figure}
1296:
1297:
1298: %------------------------------------------------------------------------
1299: \begin{figure}
1300: \begin{tabular}{cccc}
1301: & $\omega\to 0$ & $\log\omega=2.13$ & $\log\omega=2.60$ \\
1302: $\frac{4}{16}T$ &
1303: \epsfig{figure=C4eigmerBph00.eps, scale=0.35} &
1304: \epsfig{figure=C4135merBph00.eps, scale=0.35} &
1305: \epsfig{figure=C4400merBph00.eps, scale=0.35} \\[8pt]
1306: $\frac{12}{16}T$ &
1307: \epsfig{figure=C12eigmerBph00.eps, scale=0.35} &
1308: \epsfig{figure=C12135merBph00.eps, scale=0.35} &
1309: \epsfig{figure=C12400merBph00.eps, scale=0.35}
1310: \end{tabular}
1311: \caption{ \label{fig:CmerBph00}
1312: Meridional sections, $B_\phi$, $\phi=0$;
1313: $\Rm=150$, $(D,M)=(0.5,0.11)$, $A_D=0$, $A_M=0.1$.
1314: The times $t=\frac{4}{16}\,T$ and $\frac{12}{16}\,T$ correspond to
1315: maximum and minimum $M(t)$ respectively. There is a radial
1316: shift of the `clover' pattern near the equator for the eigenfunctions
1317: ($\omega\to 0$).
1318: Closely proximity of opposing flux leads to cancellation seen
1319: towards the outer boundary, lower panel with $\log\omega=2.13$.
1320: At $\log\omega=2.60$ the field is more steady.
1321: }
1322: \end{figure}
1323:
1324:
1325: %------------------------------------------------------------------------
1326: %\begin{figure}
1327: % \vspace{-10mm}
1328: % \epsfig{figure=grDRm.eps}
1329: % \caption{ \label{fig:grDRm}
1330: % $\Re(\sigma)$ (solid) and $\pm\Im(\sigma)$ (dashed)
1331: % at $(D,M)=(0.7,-0.004)$ for increasing $\Rm$.
1332: % Steady flow, $\omega=0$.
1333: % }
1334: %\end{figure}
1335: %
1336: %
1337: %------------------------------------------------------------------------
1338: \begin{figure}
1339: \epsfig{figure=grDosc.eps, scale=0.95}
1340: \caption{ \label{fig:grDosc}
1341: $\Re(\sigma)$, $\Rm=700$ at
1342: $(D,M)=(0.7,-0.008)$, $A_D=0$, $A_M=0.006$
1343: (Fig.\ \ref{fig:diamond}, E). For the dashed curve
1344: $\Im(\sigma_1)=\omega/2$.
1345: % Solutions emerge from the oscillatory region
1346: % with reversed sign for the dashed branch.
1347: }
1348: \end{figure}
1349:
1350: %------------------------------------------------------------------------
1351: \begin{figure}
1352: \epsfig{figure=D20mnrg.eps, scale=0.95}
1353: \caption{ \label{fig:D20mnrg}
1354: Magnetic energy (minus net growth) for parameters as
1355: Fig.\ \ref{fig:grDosc}. Reversing solution $\log\omega=1.30$ (solid);
1356: failed reversal $\log\omega=2.00$ (dashed).
1357: Vertical bars represent the period in the oscillatory regime.
1358: }
1359: \end{figure}
1360:
1361: %------------------------------------------------------------------------
1362: \begin{figure}
1363: % \begin{center}
1364: \begin{tabular}{cc}
1365: $0.05\,T$
1366: \epsfig{figure=005surfBr.eps, scale=0.35} &
1367: \epsfig{figure=015surfBr.eps, scale=0.35}
1368: $0.15\,T$ \\
1369: % $t=0.0157$ & $t=0.0471$ \\[8pt]
1370: $0.25\,T$
1371: \epsfig{figure=025surfBr.eps, scale=0.35} &
1372: \epsfig{figure=035surfBr.eps, scale=0.35}
1373: $0.35\,T$
1374: % $t=0.0785$ & $t=0.1100$
1375: \end{tabular}
1376: % \end{center}
1377: \caption{ \label{fig:D20surfBr}
1378: $B_r$ at the surface during the reversal in Fig.\ \ref{fig:D20mnrg}.
1379: Patches of reversed flux near the equator
1380: migrate polewards, replacing the flux at high latitudes
1381: with reversed field.
1382: }
1383: \end{figure}
1384:
1385: %------------------------------------------------------------------------
1386: \begin{figure}
1387: % \begin{center}
1388: \begin{tabular}{cc}
1389: $0.20\,T$
1390: \epsfig{figure=020surfBrf.eps, scale=0.35} &
1391: \epsfig{figure=035surfBrf.eps, scale=0.35}
1392: $0.35\,T$ \\
1393: % $t=0.0157$ & $t=0.0471$ \\[8pt]
1394: $0.50\,T$
1395: \epsfig{figure=050surfBrf.eps, scale=0.35} &
1396: \epsfig{figure=065surfBrf.eps, scale=0.35}
1397: $0.65\,T$
1398: % $t=0.0785$ & $t=0.1100$
1399: \end{tabular}
1400: % \end{center}
1401: \caption{ \label{fig:D20surfBrf}
1402: $B_r$ at the surface during the failed reversal
1403: in Fig.\ \ref{fig:D20mnrg}.
1404: Reversed flux patches have insufficient time to migrate polewards
1405: before dissipating.
1406: }
1407: \end{figure}
1408:
1409: \end{document}
1410:
1411: