1: \documentclass[12pt,preprint]{aastex}
2:
3: %\usepackage{amsmath,amssymb}
4:
5: \shorttitle{Predicting the Solar Cycle}
6: \shortauthors{Bushby \& Tobias}
7:
8: % ....................................................
9: \newcommand{\ovr}[2]{\ensuremath{\frac{{#1}}{r^{#2}}}}
10: \newcommand{\dprime}{\ensuremath{{\prime\prime}}}
11: \newcommand{\etal}{{\it et al }}
12: % ....................................................
13:
14:
15: %=============================================================================
16: \begin{document}
17:
18: %% \bibliographystyle{apj}
19:
20:
21: \title{On Predicting the Solar Cycle using Mean-Field Models}
22:
23:
24: \author{Paul J. Bushby\altaffilmark{1}}
25: \email{P.J.Bushby@damtp.cam.ac.uk}
26:
27:
28: \author{Steven M. Tobias\altaffilmark{2}}
29: \email{smt@maths.leeds.ac.uk}
30:
31:
32: \altaffiltext{1}{DAMTP, Centre for Mathematical Sciences, University of
33: Cambridge, Wilberforce Road, Cambridge CB3 0WA, U.K.}
34:
35: \altaffiltext{2}{Department of Applied Mathematics, University of Leeds,
36: Leeds LS2 9JT, U.K.}
37:
38: \begin{abstract}
39: We discuss the difficulties of predicting the solar cycle using
40: mean-field models. Here we argue that these difficulties arise owing
41: to the significant modulation of the solar activity cycle, and that
42: this modulation arises owing to either stochastic or deterministic
43: processes. We analyse the implications for predictability in both of
44: these situations by considering two separate solar dynamo models. The
45: first model represents a stochastically-perturbed flux transport
46: dynamo. Here even very weak stochastic perturbations can give rise to
47: significant modulation in the activity cycle. This modulation leads to
48: a loss of predictability. In the second model, we neglect stochastic
49: effects and assume that generation of magnetic field in the Sun can be
50: described by a fully deterministic nonlinear mean-field model --- this
51: is a best case scenario for prediction. We designate the output from
52: this deterministic model (with parameters chosen to produce
53: chaotically modulated cycles) as a target timeseries that subsequent
54: deterministic mean-field models are required to predict. Long-term
55: prediction is impossible even if a model that is correct in all
56: details is utilised in the prediction. Furthermore, we show that even
57: short-term prediction is impossible if there is a small discrepancy in
58: the input parameters from the fiducial model. This is the case even if
59: the predicting model has been tuned to reproduce the output of
60: previous cycles. Given the inherent uncertainties in determining the
61: transport coefficients and nonlinear responses for mean-field models,
62: we argue that this makes predicting the solar cycle using the output
63: from such models impossible.
64: \end{abstract}
65:
66: \keywords{(magnetohydrodynamics:) MHD -- Sun: activity -- Sun: magnetic fields}
67:
68: %=============================================================================
69: \section{Introduction}
70:
71: Magnetic activity in the Sun is known to play a central role in
72: driving both long-term and short-term dynamics (Tobias 2002; Weiss
73: 2002). The magnetic field is responsible for spectacular
74: events such as sunspots, solar flares, and coronal mass ejections, and for
75: heating the solar corona to high temperatures. Large-scale magnetic
76: activity is known to be dominated by the eleven year activity
77: cycle. This cycle has been systematically observed since the early
78: seventeenth century and its properties are well documented (see
79: e.g. Ossendrijver 2003). Of particular current interest is the impact
80: of magnetic activity on solar irradiance that might have significant
81: implications for the terrestrial climate (see Solanki \etal 2004).
82:
83: Given the importance of solar activity, it is not surprising that
84: there has been a continued interest in understanding the mechanisms
85: responsible for generating the solar magnetic field. The Sun's
86: magnetic field is believed to be generated by a hydromagnetic dynamo
87: in which motion of the solar plasma (advection) is able to sustain a
88: magnetic field against the continued action of ohmic dissipation (see
89: e.g Moffatt 1978; Charbonneau 2005). Progress in understanding this
90: fundamental problem of solar magnetohydrodynamics is slow owing to the
91: difficulties of the dynamo problem. The extreme parameters of the
92: solar interior and the inherent three-dimensionality of the dynamo
93: problem make it impossible to solve the equations accurately on a
94: computer. Much effort has therefore focused on {\it mean-field} dynamo
95: models (Steenbeck, Krause \& R\"adler 1966; Krause \& R\"adler 1980),
96: which describe the evolution of the mean magnetic field,
97: parameterising the effects of the small-scale fields and flows in
98: terms of tensor transport coefficients. These transport coefficients
99: include $\alpha_{ij}$ (which leads to a regenerative term in the
100: mean-field equations --- the so-called $\alpha$-effect) and the
101: turbulent diffusivity ($\beta_{ijk}$). We stress here that there is no
102: mechanism within the theory for determining the form of these
103: coefficients, except for flows at low magnetic Reynolds number or with
104: short correlation time, and in solar models these are usually chosen
105: in a plausible but ad-hoc manner (often, for simplicity, adopting
106: isotropic representations in which $\alpha_{ij}=\alpha \delta_{ij}$
107: and $\beta_{ijk}=\beta\epsilon_{ijk}$). Much attention has been
108: focused upon determining these transport coefficients in both the
109: linear and nonlinear dynamo regimes from numerical simulations
110: (Cattaneo \& Hughes 1996; Brandenburg \& Subramanian 2005) but there
111: is still no consensus over the nature of these, even to within an order of
112: magnitude (see Courvoisier, Hughes \& Tobias 2006). Mean-field
113: models have, however, proved successful in providing illustrations of
114: the type of behaviour that might be expected to occur in the Sun (and
115: other stars). It is often argued that, {\it although these models have
116: no predictive power}, understanding the underlying mathematical form
117: of the equations can lead to the identification of robust patterns of
118: behaviour.
119:
120: Many different models have been proposed for the solar dynamo. In the
121: distributed dynamo model, the $\alpha$-effect operates throughout the
122: convection zone and interacts with the latitudinal shear (or the
123: sub-surface shear layer, see Brandenburg 2005) to generate magnetic
124: field. Alternatively, the dynamo could be operating near the
125: tachocline, where an $\alpha$-effect might be driven either by a
126: tachocline-based instability or by turbulent convection. This, in
127: conjunction with the strong shear, could drive an ``interface''
128: dynamo (Parker 1993). Finally, there are flux transport models, in
129: which the (so-called) Babcock-Leighton mechanism produces an
130: $\alpha$-effect (or source term) at the surface. This surface
131: $\alpha$-effect is coupled to the radial shear in the tachocline
132: (where another $\alpha$-effect may be operating) via a meridional flow
133: (Choudhuri, Sch\"ussler \& Dikpati 1995; Dikpati \& Charbonneau
134: 1999). The relative merits of these models are discussed elsewhere in
135: the literature (see, e.g. Charbonneau 2005) --- the only comment we
136: make here is that this plethora of models arises because of the lack
137: of available constraints on the form of the transport coefficients in
138: the mean-field formalism. We note further that it is not clear that
139: any of the above scenarios capture the essential dynamo processes
140: correctly or that these processes can ever be captured by a mean-field
141: model.
142:
143: It is also possible to construct predictions of solar activity without
144: using dynamo theory, and there is a long literature describing these
145: predictive methods (see e.g. Zhang 1996; Hathaway, Wilson \& Reichmann
146: 1999; Sello 2003; Zhao \etal 2004; Saba, Strong \& Slater 2005). One class
147: of prediction techniques uses statistical and timeseries analysis
148: methods (see e.g. Tong 1995 for details). These methods, which are
149: also applicable in many other areas of physics, vary in complexity
150: from simple linear methods to methods that use dynamical systems
151: theory to reconstruct nonlinear attractors in phase space. However,
152: these methods have the drawback that they do not utilise any of the
153: ``physics'' of the problem. Predictions can also be made by using
154: precursor methods (see e.g. Schatten 2002), which do utilise some of
155: the physical features of the system in addition to the timeseries data.
156:
157: In recent papers (Dikpati, de Toma \& Gilman 2006; Dikpati \& Gilman
158: 2006), an attempt has been made to unify these two approaches by
159: utilising a mean-field model in order to make predictions about the
160: future activity of the Sun. These papers describe an axisymmetric,
161: mean-field model of a flux transport dynamo. Here the authors make use
162: of the observations of magnetic flux at the solar surface to feed into
163: a model of solar activity. The flux that is observed at the solar
164: surface is advected by a parameterised meridional flow (which can be
165: observed down to a certain depth) and interacts with a differential
166: rotation profile that has been inferred from helioseismology. The
167: magnetic flux also interacts with turbulence, the effects of which are
168: parameterised by certain turbulent transport coefficients
169: (representing the turbulent diffusivity and the $\alpha$-effect). As
170: with all current mean field models these turbulent transport effects
171: have been parameterised in a plausible but ad-hoc manner, and are
172: unconstrained by observations and indeed theory. The simplest predictive scheme
173: proposed by Dikpati {\it et al} (2006) therefore takes the form of a
174: parameterised linear system forced by boundary observations. The
175: implicit underlying philosophy here is that by reducing the correct
176: physics for the generation of the solar activity cycle (i.e. a
177: nonlinear self-excited dynamo) to such a scheme, predictions about
178: future solar activity can be made.
179:
180: In this paper we shall investigate the predictability of various
181: dynamo models. We demonstrate that even when all the nonlinear
182: physics of the solar dynamo is removed, problems remain for prediction
183: owing to the increased importance of stochastic effects --- even very
184: weak stochastic perturbations can produce significant modulation in
185: these linear-type models. We also discuss the best-case
186: scenario for prediction where stochastic effects can be ignored, and
187: demonstrate that in these cases prediction is still difficult owing to
188: uncertainties in the input parameters of these parameterised
189: mean-field models.
190:
191: The paper is organised as follows. In the next section we describe (in
192: a general way) the importance of modulation and the role of
193: stochasticity and nonlinearity in solar dynamo models. In section 3 we
194: investigate a flux transport model and demonstrate how the presence of
195: even extremely weak noise can render predictions useless. In section 4
196: we consider the ``best-case" scenario for prediction where noise does
197: not play a role in the modulation --- we demonstrate that more
198: accurate prediction schemes may arise by using basic timeseries
199: analysis techniques rather than from constructing mean-field models of
200: the solar cycle. Finally, in section 5 we discuss the implications of
201: our work for predictions of the solar cycle.
202:
203:
204: \section{Problems for prediction and mechanisms for modulation}
205:
206: In this section, we discuss the problems that must be overcome by
207: schemes designed to yield a prediction of future solar magnetic
208: activity. Some of these problems arise owing to the nature of solar
209: magnetic activity whilst others arise from the lack of a detailed
210: theory that is capable of describing solar magnetic activity in such
211: extreme conditions as those that exist in the solar interior.
212:
213: It is clear that if the solar cycle were strictly periodic, with a
214: constant amplitude, then it would be straightforward to predict future
215: behaviour. However, all measurements of solar magnetic activity (both
216: direct observations and evidence from proxy data) indicate that the
217: variations in the magnetic activity do not follow a periodic
218: pattern. Departures from periodicity may be driven either by
219: perturbations or by modulation. For the case of a weakly perturbed
220: periodic system, the dynamics is essentially captured by the periodic
221: signal, with the small perturbations playing a secondary role. We
222: distinguish this behaviour from a modulated signal in which there are
223: significant departures from periodicity (often occurring on longer
224: timescales), with large variations in the observed amplitude of the
225: signal. All the evidence from direct observations indicates that the
226: solar cycle is strongly modulated. The amplitude of the solar cycle
227: varies enormously over long timescales, an extreme example of this
228: modulation was a period of severely reduced activity in the
229: seventeenth century known as the Maunder Minimum. Proxy data from
230: records of terrestrial isotopes, such as $^{10}$Be and $^{14}$C (see
231: e.g. Beer 2000, Weiss \& Tobias 2000, Wagner {\it et al} 2001),
232: demonstrate that this modulation has been a characteristic feature of
233: the solar magnetic activity over (at least) the last 20,000 years.
234:
235: Mathematically there are only two possible sources for this strong
236: modulation of the basic solar cycle (Tobias 2002). The modulation may
237: arise either as a result of stochastic effects (see e.g. Ossendrijver
238: \& Hoyng 1996) or by deterministic processes (see e.g. Tobias, Weiss
239: \& Kirk 1995). In this context we define deterministic processes to be
240: those that {\it are} captured by the differential equations of dynamo
241: theory, with no random elements. Stochastic processes are those
242: that occur on an unresolved length or timescale, and so can not be
243: described by the differential equations without including a random
244: element into the model.
245:
246: It is well known that stochastic modulation can arise even if the
247: deterministic physics that leads to the production of the basic cycle
248: is essentially linear. This parameter regime is generally considered to be a
249: good one for prediction, since any nonlinear effects are only playing
250: a secondary role. However, in this stochastically-perturbed case, the
251: small random fluctuations that lead to the modulation will have large
252: short-term effects and render prediction extremely difficult, if not
253: impossible. Conversely, if the modulation arises purely as a result of
254: deterministic processes, then the underlying physics is nonlinear (or
255: potentially non-autonomous) and this leads to difficulty in prediction
256: owing to the possible presence of deterministic chaos and (more
257: importantly) the difficulty of constructing accurate nonlinear models
258: with large numbers of degrees of freedom.
259:
260: In the next two sections we demonstrate the problems for prediction
261: for dynamo models in both of the classes described above. In the next
262: section we describe a flux transport model of the same type as the one
263: used in the prediction scheme of Dikpati {\it et al} (2006) and we
264: demonstrate that even very small random fluctuations can produce
265: significant modulation, leading to extreme difficulties for
266: prediction. We then, in section 4, go on to describe a model where the
267: modulation arises owing to the presence of deterministic chaos and show that in
268: this case, prediction using model fitting is a poor way to proceed, but some
269: prediction is possible if it is possible to reconstruct the attractor
270: for activity.
271:
272: \section{Prediction using a stochastically-perturbed flux transport
273: dynamo model}
274:
275: \subsection{The dynamo model}
276:
277: We assume initially that the modulated solar magnetic activity can be
278: described by a stochastically-perturbed mean-field
279: dynamo model. In this model, nonlinear effects are playing a secondary
280: role, and all the modulation is being driven by the stochastic
281: effects. The aim of this section is to assess whether or not models
282: of this type can be used to make meaningful predictions of the solar
283: magnetic activity. In these models, the evolution of the large-scale
284: magnetic field is described by the standard mean-field equation (see,
285: for example, Moffatt 1978),
286:
287: \begin{equation}
288: \frac{\partial \mathbf{B}}{\partial t} = \nabla \times \left(
289: \alpha\mathbf{B}+\mathbf{U} \times \mathbf{B} - \beta \nabla \times \mathbf{B} \right).\label{eqn:1}
290: \end{equation}
291:
292: \noindent Here, $\mathbf{B}$ represents the large-scale magnetic
293: field and $\mathbf{U}$ corresponds to the mean velocity field, $\beta$ is
294: the (turbulent) magnetic diffusivity, and the $\alpha\mathbf{B}$ term
295: corresponds to the mean-field $\alpha$-effect. Using the well-known
296: $\alpha\omega$ approximation, we solve this equation numerically in an
297: axisymmetric spherical shell ($0.6R_{\odot} \le r \le
298: R_{\odot}$ and $0 \le \theta \le \pi$). In solving
299: Equation~(\ref{eqn:1}) we need to ensure that $\mathbf{B}$ remains
300: solenoidal (i.e. $\nabla \cdot \mathbf{B}=0$). To achieve
301: this, we decompose the magnetic field into its poloidal and toroidal
302: components,
303:
304: \begin{equation}
305: \mathbf{B}= B(r,\theta,t)\mathbf{e_{\phi}} + \nabla \times \left(
306: A(r,\theta,t)\mathbf{e_{\phi}}\right),\label{eqn:2}
307: \end{equation}
308:
309: \noindent where $B(r,\theta,t)$ denotes the toroidal (azimuthal) field
310: component and the scalar potential $A(r,\theta,t)$ relates to the
311: poloidal component of the magnetic field. So, rather than solving
312: Equation~(\ref{eqn:1}) directly, the problem has been reduced to solving two
313: coupled partial differential equations for the scalar quantities
314: $A(r,\theta,t)$ and $B(r,\theta,t)$. We adopt idealised boundary
315: conditions, in which $A=B=0$ at $\theta=0$ and $\theta=\pi$ and
316: $r=0.6R_{\odot}$ and $A$ and $B$ are smoothly matched to a potential
317: field at $r=R_{\odot}$.
318:
319: \par This particular dynamo model is closely related to the flux
320: transport model described by Dikpati \& Charbonneau (1999). The
321: large-scale velocity field, $\mathbf{U}$, is given by
322:
323: \begin{equation}
324: \mathbf{U}= u_r(r,\theta)\mathbf{e_{r}} +
325: u_{\theta}(r,\theta)\mathbf{e_{\theta}} +
326: \Omega(r,\theta)r\sin\theta\mathbf{e_{\phi}},\label{eqn:3}
327: \end{equation}
328:
329: \noindent where $\Omega(r,\theta)$ is a prescribed analytic fit to the
330: helioseismologically-determined solar rotation profile (see, for
331: example, Bushby 2006) and $u_r$ and $u_{\theta}$ correspond to a
332: prescribed meridional circulation. We assume that the meridional
333: circulation pattern in each hemisphere consists of a single cell, with
334: a polewards flow at the surface and an (unobservable) equatorwards
335: flow at the base of the convection zone --- the flow is confined to the
336: region $R_b \le r \le R_{\odot}$. The functional form that we adopt
337: for this flow is similar in form to the one described by Dikpati \&
338: Charbonneau (1999),
339:
340: \begin{eqnarray}
341: u_r(r,\theta)&=&U_o \left( \frac{R_{\odot}}{r} \right)^2 \left[
342: -\frac{2}{3} + \frac{1}{2}c_1\xi^{0.5}-
343: \frac{4}{9}c_2\xi^{0.75}\right] \xi \sin \theta \left( 3\cos^2
344: \theta - \sin^2 \theta \right),\label{eqn:4}\\
345: u_{\theta}(r,\theta)&=&U_o \left( \frac{R_{\odot}}{r} \right)^3 \left[
346: -1 + c_1\xi^{0.5}-c_2\xi^{0.75}\right] \cos \theta \sin^2
347: \theta,\label{eqn:5}
348: \end{eqnarray}
349:
350: \noindent where $\xi(r)=[(R_{\odot}/r)-1]$, $c_1=4[\xi(R_b)]^{-0.5}$,
351: $c_1=3[\xi(R_b)]^{-0.75}$, and $U_o$ is some characteristic flow
352: speed. This flow pattern can be stochastically perturbed by setting
353: $R_b=0.7R_{\odot}+\epsilon(t)$, where $\epsilon(t)$ is a time-dependent,
354: randomly fluctuating variable in the range $-0.005R_{\odot} \le
355: \epsilon \le 0.005R_{\odot}$. The aim here is to assess whether or not
356: such weak stochastic variations in the flow pattern could give rise to
357: significant modulation in the activity cycle, and if so what are the
358: consequences for prediction.
359:
360: \par In order to complete the specification of the model, we need to
361: choose plausible functional forms for the $\alpha$-effect and the
362: turbulent magnetic diffusivity. It should be emphasised again that
363: these mean-field coefficients are poorly constrained by theory and
364: observations, although plausible assumptions can be made. Defining
365: $\beta_o$ to be a characteristic value of the turbulent magnetic
366: diffusivity within the solar convection zone, we adopt a similar
367: spherically-symmetric profile to that adopted by Dikpati \&
368: Charbonneau (1999),
369:
370: \begin{equation}
371: \beta(r) = \frac{1}{2} (\beta_o-\beta_c)\left[ 1 + \mbox{erf}\left(\frac{r-0.7R_{\odot}}{0.025R_{\odot}} \right)\right] +
372: \beta_c, \label{eqn:6}
373: \end{equation}
374:
375: \noindent where erf corresponds to the error function and $\beta_c$
376: (here taken to be $1\%$ of $\beta_o$) represents the magnetic
377: diffusivity below the turbulent convection zone. Following Dikpati
378: \& Charbonneau (1999), rather than prescribing a simple functional
379: form for $\alpha$ we neglect the $\alpha$-effect term in the toroidal
380: ($B$) field equation and replace the corresponding $\alpha B$ term in
381: the poloidal ($A$) equation by a non-local, nonlinear source of
382: poloidal flux,
383:
384: \begin{eqnarray}
385: S(r,\theta,t)&=&\frac{S_o}{2}\left[ 1 +
386: \mbox{erf}\left(\frac{r-0.95R_{\odot}}{0.01R_{\odot}} \right)\right]
387: \left[ 1 - \mbox{erf}\left(\frac{r-R_{\odot}}{0.01R_{\odot}}
388: \right)\right] \\ \nonumber &&\left[
389: 1+\left(\frac{B(0.7R_{\odot},\theta,t)}{B_o}\right)^2\right]^{-1}\sin
390: \theta \cos \theta B(0.7R_{\odot},\theta,t).\label{eqn:7}
391: \end{eqnarray}
392:
393: \noindent Here, $S_o$ is a characteristic value of this poloidal
394: source and $B_o$ represents the (somewhat arbitrarily chosen) field
395: strength at which this non-local source becomes suppressed by the
396: magnetic field. This source term parameterises the contribution to the
397: poloidal magnetic flux due to the decay of active regions --- the
398: non-locality reflects the fact that active regions are believed to
399: form as the result of buoyant magnetic flux rising from the base of
400: the convection zone to the solar photosphere. See Dikpati \&
401: Charbonneau (1999) for a more detailed discussion of this source term, though
402: again it must be stressed that the functional form and the nonlinear dependence
403: are chosen in a plausible yet ad-hoc manner.
404:
405: \subsection{Numerical results}
406:
407: In order to carry out numerical simulations, we first
408: non-dimensionalise this flux transport model. By using scalings similar
409: to those described by Dikpati \& Charbonneau (1999), it can be shown
410: that the model solutions are fully determined by two non-dimensional
411: parameters (once other parameters such as $B_o$ have been selected).
412: Denoting the equatorial angular velocity at the solar
413: surface by $\Omega_{eq}$, these non-dimensional parameters are the
414: Dynamo number, $D=S_o\Omega_{eq}R_{\odot}^3/\beta_o^2$, and the
415: magnetic Reynolds number corresponding to the meridional flow,
416: $Re=U_oR_{\odot}/\beta_o$. Here, we set $D=7 \times 10^6$ and
417: $Re=5600$. In the absence of stochastic noise, this set
418: of parameters produces a strong circulation-dominated dynamo in which the
419: magnetic energy is a periodic function of time. Although the dynamo
420: number is not weakly supercritical, nonlinear effects are not strong
421: enough here to produce a modulated activity cycle --- the primary role
422: of the nonlinearity is to prevent the unstable dynamo mode from
423: growing exponentially. We term such a model a ``linear-type'' model.
424:
425: \par When weak stochastic effects are included in the model, the
426: resulting activity cycle is indeed weakly modulated. This is
427: illustrated in Figure~1, which shows the time-dependence of this
428: solution. The time-series clearly illustrates that, although the
429: amplitude of the ``cycle minimum'' only appears to be weakly
430: time-dependent, there are significant variations in the peak amplitude
431: of the magnetic energy time-series. These variations are qualitatively
432: similar to those observed by Charbonneau \& Dikpati (2000), who
433: considered large amplitude random fluctuations in the flow pattern
434: within the solar convection zone --- the peak amplitude of these
435: fluctuations was comparable with the peak amplitude of the flow. In this
436: particular model, we have shown that even very weak stochastic
437: variations in the centre of mass of the flow pattern can still produce
438: significantly modulated behaviour. These stochastic effects are
439: expected to become increasingly significant for dynamo numbers
440: approaching critical. So, these models are obviously highly sensitive
441: to the addition of stochastic noise.
442:
443: \par In the absence of stochastic noise, the attractor (in phase
444: space) for this solution is two-dimensional, and the future behaviour
445: of the solution at any instant in time is entirely determined by the
446: current position of the system on the attractor. The same is not true
447: when this system is perturbed by stochastic effects, and it clearly
448: becomes much more difficult to predict the future behaviour of the
449: system. Since the attractor of this stochastically perturbed solution
450: cannot be unambiguously defined, another possible way of assessing the
451: ``predictability'' of this solution is to look for a correlation
452: between successive cycle maxima. Defining $T_n$ to be the magnitude of
453: the $n^{th}$ cycle maximum, Figure~2 shows $T_{n+1}$ as a function of
454: $T_n$. It is clear from this scatter plot that there is no obvious
455: correlation between the amplitudes of successive cycle maxima in this
456: case. Since the modulation is being driven entirely by random
457: stochastic forcing, this result is not surprising. This lack of
458: correlation suggests that the behaviour of previous cycles cannot be
459: used to infer the magnitude of the following one. This implies that
460: even weak stochastic effects may seriously reduce the possibilities
461: for solar cycle prediction in this linear-type regime.
462:
463: \section{Predictions using a deterministic dynamo model}
464:
465: \subsection{The dynamo model}
466:
467: In the previous section, we demonstrated that even very weak
468: stochastic perturbations to the meridional flow pattern can lead to a
469: loss of predictability in a linear-type flux transport dynamo
470: model. In that model, the modulation of the activity cycle was driven
471: entirely by stochastic effects. As discussed in Section~2, the only other
472: possible scenario is that the observed modulation is driven by
473: nonlinear effects. This scenario, where the observed modulation is
474: deterministic in origin, is the ``best-case'' scenario for prediction,
475: as in this case the entirely unpredictable stochastic elements may be
476: ignored. We stress again that, given that solar magnetic activity is
477: significantly modulated, either deterministic or stochastic modulation must be
478: considered in any realistic model (predictive or otherwise) of the
479: solar cycle. So, in this section, we completely neglect stochastic effects and
480: assume that the observed (chaotic) modulation in the solar magnetic
481: activity can be described by a fully deterministic model in which any activity
482: modulation (e.g. solar-like ``Grand minima'') is driven entirely by
483: nonlinear effects. The model that we use was described in detail in
484: two recent papers (Bushby 2005, 2006), so we only present a brief
485: description here. The exact details of the model are unimportant for
486: our main conclusions.
487:
488: \par Like the flux transport dynamo model from the previous section,
489: this model describes an axisymmetric, mean-field,
490: $\alpha\omega$-dynamo in a spherical shell. Unlike the previous model,
491: this model represents an ``interface-like'' dynamo that is operating
492: primarily in the region around the base of the solar convection
493: zone. It is worth mentioning again that (as discussed in
494: the introduction) there is still no general consensus regarding which of
495: these dynamo scenarios is more likely to be an accurate representation
496: of the solar dynamo. For this interface-like dynamo model, we neglect
497: meridional motions, since they are poorly determined near the base of the solar
498: convection zone. Like several earlier models (e.g. Tobias 1997;
499: Moss \& Brooke 2000; Covas \etal 2000), this dynamo model includes the
500: feedback (via the azimuthal component of the Lorentz force) of the
501: mean magnetic field upon the differential rotation (Malkus \& Proctor
502: 1975). This nonlinear feedback is a crucial element of the model and, in the
503: absence of stochastic effects, is the sole driver of modulation in the
504: magnetic activity cycle. Denoting this magnetically-driven velocity
505: perturbation by $V(r,\theta,t)$, the large-scale velocity field is
506: given by
507:
508: \begin{equation}
509: \mathbf{U} = \left[\Omega(r,\theta)r\sin\theta +
510: V(r,\theta,t)\right]\mathbf{e_{\phi}},
511: \label{eqn:8}
512: \end{equation}
513:
514: \noindent where (as in the previous model) $\Omega(r,\theta)$
515: represents an analytic fit to the solar differential
516: rotation. Whilst the evolution of the large-scale magnetic field is
517: again governed by Equation~(1), an additional evolution equation is
518: required for the velocity perturbation, $V$. This equation is given by
519:
520: \begin{eqnarray}
521: \frac{\partial V}{\partial t}&=& \frac{1}{\mu_o
522: \rho} \left[(\nabla \times \mathbf{B}) \times \mathbf{B} \right]\cdot
523: \mathbf{e_{\phi}} +
524: \frac{1}{r^3}\frac{\partial}{\partial r}\left[\nu r^4
525: \frac{\partial}{\partial r} \left(\frac{V}{r}\right)\right] \\
526: \nonumber &&+
527: \frac{1}{r^2 \sin^2 \theta}\frac{\partial}{\partial \theta}\left[\nu
528: \sin^3 \theta \frac{\partial}{\partial \theta} \left(\frac{V}{\sin
529: \theta}\right)\right],\label{eqn:9}
530: \end{eqnarray}
531:
532: \noindent where $\rho$ represents the fluid density (here taken to be
533: constant), $\mu_o$ is the permeability of free space and $\nu$
534: represents the (turbulent) fluid viscosity.
535:
536:
537: \par In order to complete the model, the spatial dependence of the
538: transport coefficients ($\alpha$, $\beta$ and $\nu$) must also be
539: specified. Again, we emphasise that there are no direct
540: observational constraints relating to these coefficients --- as noted
541: in the introduction, there is no consensus as to their form and there
542: is still a debate as to their order of magnitude (and even their
543: sign). Having said that, it is possible to make some plausible
544: assumptions for an ``interface-like'' dynamo model (see Bushby 2006
545: for more details). The precise choices of these parameters are unimportant for
546: our main conclusions.
547:
548: \par Having set up this model, it is possible to choose a set of
549: parameters so that the solutions {\it do} reproduce some salient
550: features of the solar dynamo (Bushby 2005, 2006). We stress here that,
551: although the parameters have been chosen in a plausible manner, this
552: dynamo model should not be regarded as an accurate representation of
553: the solar interior and is subject to many uncertainties. Furthermore
554: we stress again that this is the case with {\it all} mean-field solar
555: dynamo models. However we use this model as a useful tool to analyse
556: the possibility of producing predictive models of the solar cycle. We
557: proceed by choosing fiducial parameters and profiles for the turbulent
558: transport coefficients that lead to ``solar-type" magnetic activity,
559: with chaotically modulated cycles and recurrent ``Grand Minima''. We
560: then integrate this model forward in time to produce a timeseries and
561: designate this timeseries as the ``target" run, which any subsequent
562: model should be able to predict. This target run is shown in Figure~3,
563: which shows a timeseries of the activity together with a
564: reconstruction of the dynamo attractor in phase space. Although this
565: solution is chaotically modulated, it is certainly no more chaotic than
566: the equivalent attractor for the $^{10}$Be data, which is a well-known
567: proxy for solar magnetic activity (e.g. Beer 2000). Whilst the
568: nonlinear effects are significant enough to drive the modulation, they
569: are actually very difficult to detect. In this model, the cyclic
570: component of the fluctuations in the differential rotation (which are
571: driven by the nonlinear Lorentz force) are small compared with the
572: mean differential rotation. This is consistent with observations of
573: the (so called) torsional oscillations in the solar convection zone. Finally,
574: note once more that, since the modulation is driven entirely by
575: nonlinear effects, this model is specified exactly.
576:
577: \subsection{Numerical results}
578:
579: The question is then posed as to whether {\it any} mean-field model
580: can be constructed that leads to meaningful predictions of the future
581: behaviour of the target run. Clearly the best chance for a mean-field
582: model being capable of predicting the future behaviour of the target
583: run is to use {\it the exact model} that led to the target run
584: data. Hence we test this model first, as all subsequent models will be
585: inferior to this. We proceed by setting the model parameters to be
586: those that generated the long test run, and consider the behaviour of
587: solutions that are started from very similar points on the
588: attractor. Some of the solutions are shown in Figure~4. This figure
589: shows clearly that although the predictor solutions are able to track
590: the target solution for a couple of activity cycles, the nature of the
591: solutions means that the predictors and target solution diverge
592: quickly after this time. This is not surprising behaviour. It is
593: well-known that chaotic solutions have a sensitive dependence on
594: initial conditions and that long-term prediction of such solutions is
595: fraught with problems (see e.g. Tong 1995). What is clear is that
596: simply using a model that is based upon mean-field theory will not
597: work in the long term {\it even if the model is correct in every
598: detail}. One might be able to predict one or two cycles ahead {\it if
599: one has solved the problem of constructing an exact representation of
600: the solar dynamo} but as noted above this is not an easy task.
601:
602: We now turn to the related problem of short-term prediction. As
603: discussed in the introduction there are large uncertainties in the
604: form and amplitude of the input parameters for all mean-field models.
605: What we investigate here is whether these uncertainties lead to
606: significant difficulties in prediction even in the short term. Again
607: we examine the best case scenario and consider a mean-field model for
608: the predictor runs that has correctly parameterised the {\it form} of
609: all the input variables (differential rotation, $\alpha$-effect,
610: turbulent diffusivity and nonlinear response). In addition these
611: predictors have been given the correct input {\it values} for
612: all-but-one of the parameters. Hence the predictor models are exactly
613: the same as the target model with the exception of one input parameter
614: that has been altered by $5\%$. This would be a staggeringly good
615: representation should it be possible to achieve this for solar
616: activity. Furthermore we increase the chances of the predictor being
617: able to predict the future behaviour of the target solution by
618: matching the two timeseries over a number of cycles. This is analogous
619: to the procedure employed by Dikpati \etal (2006) who cite support for
620: their forecasting model by assuring that their model agrees with the
621: solar cycle data for eight solar cycles --- in reality this is not
622: difficult to achieve with enough model parameters at one's
623: disposal. Figure~5 shows the results of integrating the predictor
624: models for two different choices of incorrect parameter. Note that
625: even though the predictor has been designed to reproduce the target
626: over a number of cycles and that the predictor is very closely related
627: to the target, there is still a good chance that it can get the next
628: cycle incorrect, with significant errors in (particularly) the cycle
629: amplitude. There are also clear variations in the cycle period, which
630: obviously implies that the exact time between successive cycle maxima
631: is also an unpredictable feature of the system.
632:
633: We stress again that any mean-field model of solar activity includes
634: transport coefficients that are still uncertain possibly to an order of
635: magnitude (and certainly not to $5\%$ accuracy). Although the
636: incredible success of global and local helioseismology is placing
637: restrictions on the form of the differential rotation and the
638: meridional flows, it is unlikely in the foreseeable future that
639: significant constraints will be put on the transport coefficients or their
640: nonlinear response to the mean magnetic field.
641:
642: \section{Predictions using a reconstruction of the attractor}
643:
644: Having established that there are difficulties in obtaining reliable
645: predictions by fitting mean-field models (even if the modulation is
646: deterministic in origin), it is of interest to determine
647: whether or not more reliable predictions could be obtained by
648: utilising more general timeseries analysis techniques. In order to
649: reconstruct an attractor from a given timeseries, it is necessary to
650: define a corresponding phase space. There are various ways of doing
651: this, but given (any) discrete timeseries, $x(t)$, in which the data
652: is sampled at intervals of $\Delta t$, the vector
653:
654: \begin{equation}
655: \mathbf{X}(t) = \left[ x(t), x(t-\Delta t)..., x(t-(d-1)\Delta t)\right]
656: \label{eqn:10}
657: \end{equation}
658:
659: \noindent defines a
660: point in a $d$-dimensional ``embedded'' phase space (see, e.g., Farmer
661: \& Sidorowich 1987; Casdagli 1989). Given a time $T$, the idea of a
662: prediction algorithm is to find a mapping $f$ such that $f
663: \left(\mathbf{X}(T)\right)$ gives a good approximation to $x(T+\Delta
664: t)$. The predictive mapping technique that is used here uses a local
665: approximation method (see, e.g., Casdagli 1989), which considers the
666: behaviour of the nearest neighbours, in phase space, to
667: $\mathbf{X}(T)$. By using a least squares fit, the subsequent
668: evolution of each of these neighbouring points in phase space is used
669: to construct a piecewise-linear approximation to the predictive map,
670: $f$. This approximate mapping can then be applied to $\mathbf{X}(T)$
671: to obtain an estimate for $x(T+\Delta t)$. This algorithm can then be
672: repeated to find estimates for $x(T+2 \Delta t)$ and subsequent
673: points. The optimal value for $d$ can be determined by minimising the
674: error of this predictive algorithm over the known segment of the
675: timeseries.
676:
677: The results of applying this predictor algorithm to the target
678: solution are also shown in Figure~5, where the timeseries
679: predictions are shown as crosses. The prediction is started from the
680: cycle maximum before the mean-field predictor diverges from the
681: target. Longer training timeseries lead to a more densely-populated
682: reconstructed attractor, which increases the probability of making
683: more accurate predictions. However, rather than using the
684: entire target run, these predictions are based upon (approximately)
685: $50$ cycles --- this will give a fairer comparison between these
686: results and timeseries predictions that are based upon the real sunspot
687: data. The application of the algorithm to earlier segments of
688: the timeseries suggests that a value of $d \ge 5$ is required in
689: order to minimise predictive errors. As can be seen from Figure~5,
690: this algorithm appears to predict the magnitude of the maximum of the
691: following cycle to a reasonable degree of accuracy, although the
692: predictions subsequently diverge from the target. Whilst neither of
693: these techniques are capable of producing reliable long-term predictions,
694: these results do suggest that for the short-term prediction of solar
695: magnetic activity, timeseries analysis techniques may provide a
696: viable alternative to predictions based simply upon mean-field dynamo
697: models (provided stochastic effects can be neglected).
698:
699: \section{Conclusions}
700:
701: Solar magnetic activity arises as a result of a hydromagnetic dynamo
702: --- that much we believe to be true. As yet, there is no consensus on
703: the location of the dynamo, the dominant nonlinear or stochastic
704: effects, or even the fundamental processes that are responsible for
705: the operation of such a dynamo. Although plausible mechanisms have
706: been proposed, as yet none of these are entirely satisfactory. Against
707: this background, there is a drive to be able to predict solar activity
708: with greater accuracy, due to the importance of this activity in
709: driving solar events.
710:
711: What we have demonstrated here is that no meaningful predictions can
712: be made from illustrative mean-field models, no matter how they are
713: constructed. If the mean-field model is constructed to be a driven linear
714: oscillator then the small stochastic effects that lead to the modulation
715: will have an extremely large
716: effect on the basic cycle and make even short-term prediction
717: extremely difficult. The second scenario, where the modulation arises
718: as a result of nonlinear processes rather than stochastic
719: fluctuations, is clearly a better one for prediction --- though here
720: too, prediction is fraught with difficulties. Owing to the inherent
721: nonlinearity of the dynamo system, long-term predictions are
722: impossible (even if the form of the model is completely correctly
723: determined). Furthermore, even short-term prediction from mean-field
724: models is meaningless because of fundamental uncertainties in the form
725: and amplitude of the transport coefficients and nonlinear
726: response. Any deterministic nonlinear model that produces chaotically
727: modulated activity cycles will be faced with the same difficulties.
728:
729: The equations that describe dynamo action in the solar interior are
730: known to be nonlinear partial differential equations --- the momentum
731: equation is nonlinear in both the velocity and the magnetic field. One
732: indication of the role played by nonlinear effects in the solar dynamo
733: is the presence of cyclic variations in the solar differential
734: rotation (the ``torsional oscillations''). Furthermore estimates of
735: the field strength at the base of the convection zone consistent with
736: the observed formation of active regions yield fields of sufficient
737: strength ($10^4-10^5$G) for the nonlinear Lorentz force to be
738: extremely significant, whilst the flows are vigorously nonlinear and
739: turbulent. It therefore seems extremely unlikely that the dynamics of
740: the solar interior can be described by a forced linear system without
741: throwing away much (if not all) of the important physics. In this case
742: it must be argued {\it not only} that this discarded physics is
743: irrelevant to the dynamo process but also that the parameterisation of
744: the unresolved physics should not include a stochastic component, as
745: this would have an extremely large effect on such a relinearised system.
746:
747: It is certainly tempting to try to use the observed magnetic flux at
748: the solar surface as an input to a model for prediction (whether
749: nonlinear or stochastic, mean-field or full MHD). Certainly any fully
750: consistent solar activity model constructed in the future should be
751: capable of reproducing the observed pattern of magnetic activity at
752: the solar surface, although this will require a complete understanding
753: not only of the generation process via dynamo action, but also the
754: processes which lead to the formation and subsequent rise of
755: concentrated magnetic structures from the solar interior to the
756: surface. However it is not clear what role the flux at the solar
757: surface plays in the basic dynamo process. Is it inherent to the
758: process (as modelled by flux transport dynamos) or simply a by-product
759: of the dynamo process that is occurring deep within the sun? Estimates
760: suggest that between $5$ and $10 \%$ of the solar flux generated in
761: the deep interior makes it to the solar surface (e.g. Galloway \&
762: Weiss 1981). For the flux at the solar surface to be the key for
763: dynamo action, it must be explained why the majority of the magnetic
764: flux that resides in the solar interior plays such a little part in
765: the dynamics (to such an extent that it does not even appear as a
766: small stochastic perturbation to the large-scale flux transport
767: dynamo).
768:
769: Finally it is important to stress that {\it even if} a model has been
770: tuned so as to reproduce results over a number of solar activity
771: cycles, then there is a good chance of error in the prediction for the
772: next cycle. Any advection-diffusion system in which one is free to
773: specify not only the sources and the sinks but also the transport
774: processes can be tuned to reproduce any required features of activity.
775: Moreover, the formulation of a prediction in terms of a parameterised
776: mean-field model does not inherently put the prediction on a sounder
777: scientific basis than a prediction based on methods of timeseries
778: analysis alone (some of which use very sophisticated mathematical
779: techniques). This, of course, is not to say that any given prediction
780: from such a model will be incorrect, just that the basis for making
781: the prediction has no strong scientific support.
782:
783: \acknowledgments
784:
785: We would like to thank Nigel Weiss for useful discussions and for
786: providing helpful comments and suggestions. PJB would like to
787: acknowledge the support of PPARC.
788:
789: \clearpage
790: %=============================================================================
791: %% \bibliography{apj-jour,ms}
792:
793: \begin{thebibliography}{34}
794: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
795:
796: \bibitem[{{Beer}(2000)}]{Beer00}
797: Beer, J. 2000, Space Science Rev., 94, 53
798:
799: \bibitem[{{Brandenburg}(2005)}]{Brand05}
800: Brandenburg, A. 2005, ApJ, 625, 539
801:
802: \bibitem[{{Brandenburg \& Subramanian}(2005)}]{BrandS05}
803: Brandenburg, A., Subramanian, K. 2005, Physics Reports, 417, 1
804:
805: \bibitem[{{Bushby}(2005)}]{Bush05}
806: Bushby, P. J. 2005, Astron. Nachr., 326, 218
807:
808: \bibitem[{{Bushby}(2006)}]{Bush06}
809: Bushby, P. J. 2006, MNRAS, 371, 772
810:
811: \bibitem[{{Casdagli}(1989)}]{Casd89}
812: Casdagli, M. 1989, Physica D, 35, 335
813:
814: \bibitem[{{Cattaneo \& Hughes}(1996)}]{Catt96}
815: Cattaneo, F., Hughes, D. W. 1996, Phys. Rev. E, 54, 4532
816:
817: \bibitem[{{Charbonneau \& Dikpati}(2000)}]{Charb00}
818: Charbonneau, P., Dikpati, M. 2000, ApJ, 543, 1027
819:
820: \bibitem[{{Charbonneau}(2005)}]{Charb05}
821: Charbonneau, P. 2005, Living Reviews in Solar Physics, Vol. 2, No. 2
822:
823: \bibitem[{{Choudhuri, Sch\"ussler \& Dikpati}(1995)}]{Choud95}
824: Choudhuri, A. R., Sch\"ussler, M., Dikpati, M. 1995, A\&A, 303, L29
825:
826: \bibitem[{{Courvoisier, Hughes \& Tobias}(2006)}]{Courv06}
827: Courvoisier, A., Hughes, D. W., Tobias, S. M. 2006, Phys. Rev. Lett.,
828: 96, 034503
829:
830: \bibitem[{{Covas \etal}(2000)}]{Covas2000}
831: Covas, E., Tavakol, R., Moss, D., Tworkowski, A. 2000, A\&A, 360, L21
832:
833: \bibitem[{{Dikpati \& Charbonneau}(1999)}]{Dikp99}
834: Dikpati, M., Charbonneau, P. 1999, ApJ, 518, 508
835:
836: \bibitem[{{Dikpati \& Gilman}(2006)}]{Dikgil06}
837: Dikpati, M., Gilman, P. A. 2006, ApJ, 649, 498
838:
839: \bibitem[{{Dikpati, de Toma \& Gilman}(2006)}]{Dikp06}
840: Dikpati, M., de Toma, G., Gilman, P. A. 2006, Geophys. Research Lett., 33,
841: L05102
842:
843: \bibitem[{{Farmer \& Sidorowich}(1987)}]{Farm87}
844: Farmer, J. D., Sidorowich, J. J. 1987, Phys Rev. Lett., 59, 845
845:
846: \bibitem[{{Galloway \& Weiss}(1981)}]{GallWeiss81}
847: Galloway, D.~J., Weiss, N.~O. 1981, ApJ, 243, 945
848:
849: \bibitem[{{Hathaway, Wilson \& Reichmann}(1999)}]{Hath99}
850: Hathaway, D. H., Wilson, R. M., Reichmann, E. J. 1999, Journal of
851: Geophys. Research, 104, 22,375
852:
853: \bibitem[{{Krause \& R\"adler}(1980)}]{Krause80}
854: Krause, F., R\"adler, K. H. 1980, Mean-field Magnetohydrodynamics and
855: Dynamo Theory (Oxford: Pergamon Press)
856:
857: \bibitem[{{Malkus \& Proctor}(1975)}]{Malk75}
858: Malkus, W. V. R., Proctor, M. R. E. 1975, J. Fluid Mech., 67, 417
859:
860: \bibitem[{{Moffatt}(1978)}]{Moff78}
861: Moffatt, H. K. 1978, Magnetic Field Generation in Electrically
862: Conducting Fluids (Cambridge: Cambridge University Press)
863:
864: \bibitem[{{Moss \& Brooke}(2000)}]{Moss00}
865: Moss, D., Brooke, J. 2000, MNRAS, 315, 521
866:
867: \bibitem[{{Ossendrijver}(2003)}]{Oss03}
868: Ossendrijver, M. 2003, Astron. and Astrophys. Rev., 11, 287
869:
870: \bibitem[{{Ossendrijver \& Hoyng}(1996)}]{OsHo96}
871: Ossendrijver, A. J. H., Hoyng, P. 1996, A\&A, 313, 959
872:
873: \bibitem[{{Parker}(1993)}]{Park93}
874: Parker, E. N. 1993, ApJ, 408, 707
875:
876: \bibitem[{{Saba, Strong \& Slater}(2005)}]{Saba05}
877: Saba, J. L. R., Strong, K. T., Slater, G. L. 2005, Memorie della
878: Societa Astronomica Italiana, 76, 1034
879:
880: \bibitem[{{Schatten}(2002)}]{Schat02}
881: Schatten, K. 2002, J. Geophys. Res., 107(A11), 1377
882:
883: \bibitem[{{Sello}(2003)}]{Sello03}
884: Sello, S. 2003, A\&A, 410, 691
885:
886: \bibitem[{{Solanki et al.}(2004)}]{Sol04}
887: Solanki, S. K., Usoskin, I. G., Kromer, B., Sch\"ussler, M., Beer,
888: J. 2004, Nature, 431 (7012), 1084
889:
890: \bibitem[{{Steenbeck, Krause \& R\"adler}(1966)}]{Steen66}
891: Steenbeck, M., Krause, F., R\"adler, K. H. 1966, Z. Naturforsch., 21a,
892: 369
893:
894: \bibitem[{{Tobias, Weiss \& Kirk}(1995)}]{TWK95}
895: Tobias, S.~M., Weiss, N.~O., Kirk, V. 1995, MNRAS, 273, 1150
896:
897: \bibitem[{{Tobias}(1997)}]{Tobias97}
898: {Tobias}, S.~M. 1997, A\&A, 322, 1007
899:
900: \bibitem[{{Tobias}(2002)}]{Tobias02}
901: {Tobias}, S.~M. 2002, Phil. Trans. Roy. Soc. Lond. A, 360, 2741
902:
903: \bibitem[{{Tong}(1995)}]{Tong95}
904: Tong, H. 1995, Chaos and Forecasting (Singapore: World Scientific
905: Publications)
906:
907: \bibitem[{{Wagner \etal}{2001}}]{Wagner}
908: Wagner, G., Beer, J., Masarik, J., Muscheler, R., Kubik, P.~W., Mende,
909: W., Laj, C., Raisbeck, G.~M., Yiou, F. 2001, Geophys. Res. Lett., 28,
910: 303
911:
912: \bibitem[{{Weiss}(2002)}]{Weiss02}
913: {Weiss}, N.~O. 2002, A\&G, 43, 3.09
914:
915: \bibitem[{{Weiss \& Tobias}{2000}}]{WeissTobias}
916: {Weiss}, N.~O., Tobias, S.~M. 2000, Space Science Rev., 94, 99
917:
918: \bibitem[{{Zhang(1996)}}]{Zhang96}
919: Zhang, Q. 1996, A\&A, 310, 646
920:
921: \bibitem[{{Zhao et al.}(2004)}]{Zhaoetal04}
922: Zhao, H., Liang, H., Zhan, L., Zhong, S. 2004, ChA\&A, 28, 67
923:
924: \end{thebibliography}
925:
926: \clearpage
927:
928: \begin{figure}
929: \epsscale{0.68}
930: \plotone{f1a.eps}
931: \epsscale{0.68}
932: \plotone{f1b.eps}
933: \caption{The time evolution of the stochastically-perturbed flux
934: transport dynamo. Top: Timeseries
935: for the mean of the squared toroidal field ($B^2$) at the base of
936: the convection zone. Bottom: In this figure, $B^2$ at the base of
937: the convection zone is plotted
938: against the mean of the squared values of the poloidal magnetic
939: potential ($A^2$) at the surface of the domain.\label{fig1}}
940: \end{figure}
941:
942: \begin{figure}
943: \plotone{f2.eps}
944: \caption{The lack of correlation between successive maxima in the
945: stochastically perturbed timeseries (as shown in Figure~1). Defining
946: $T_n$ to be magnitude of the $n^{th}$ maximum, this plot shows the
947: sequential behaviour of these maxima, plotting $T_{n+1}$ as a
948: function of $T_n$.\label{fig2}}
949: \end{figure}
950:
951: \clearpage
952:
953: \begin{figure}
954: \plotone{f3a.eps}
955: \plotone{f3b.eps}
956: \caption{The time evolution of the target solution. Top: Timeseries
957: for the mean of the squared toroidal field ($B^2$) in the dynamo
958: region. Bottom: An attractor for the target solution,
959: in which $B^2$ is plotted against the mean of the squared values of the
960: poloidal magnetic potential ($A^2$) and the velocity
961: perturbation ($V^2$).\label{fig3}}
962: \end{figure}
963:
964: \clearpage
965:
966: \begin{figure}
967: \plotone{f4.eps}
968: \caption{Three timeseries showing the evolution of the mean squared
969: toroidal field in the dynamo region. The solid line shows a
970: segment of the target solution timeseries; the dashed and dotted
971: lines show the time-evolution of solutions that are started from nearby
972: points on the same attractor. Although all solutions have the same
973: model parameters, the timeseries rapidly diverge after a couple of
974: cycles.\label{fig4}}
975: \end{figure}
976:
977: \clearpage
978:
979: \begin{figure}
980: \begin{center}
981: \epsscale{0.99}
982: \plotone{f5a.eps}
983: \vspace{0.1cm}
984: \plotone{f5b.eps}
985: \caption{Attempts to predict two different segments of the target
986: solution timeseries: In each plot, the solid line shows a segment of
987: the target solution timeseries, the dashed line shows the behaviour of
988: the chosen mean-field predictor (different predictors are used for
989: each plot), whilst the crosses show the predictions that are obtained
990: by reconstructing the nonlinear attractor for the target solution. In each
991: case, the mean-field predictors have been optimised by ensuring that
992: the chosen predictor closely matches the target timeseries segment for
993: a large number of cycles (6 cycles in the upper plot, 9 in the
994: lower).\label{fig5}}
995: \end{center}
996: \end{figure}
997:
998:
999: \end{document}
1000: