0806.2551/ms.tex
1: % mn2esample.tex
2: %
3: % v2.1 released 22nd May 2002 (G.Hutton)
4: %
5: % The mnsample.tex file has been amended to highlight
6: % the proper use of LaTeX2e code with the class file
7: % and using natbib cross-referencing. These changes
8: % do not reflect the original paper by A. V. Raveendran.
9: % v1.1 released 18th July 1994
10: % v1.0 released 28th January 1994
11: 
12: %\documentclass[useAMS,usenatbib,referee]{mn2e}
13: \documentclass[useAMS,usenatbib]{mn2e}
14: 
15: \usepackage{epsfig}
16: \usepackage{amsmath}
17: 
18: \addtolength{\topmargin}{-1cm}
19: 
20: 
21: % If your system does not have the AMS fonts version 2.0 installed, then
22: % remove the useAMS option.
23: %
24: % useAMS allows you to obtain upright Greek characters.
25: % e.g. \umu, \upi etc.  See the section on "Upright Greek characters" in
26: % this guide for further information.
27: %
28: % If you are using AMS 2.0 fonts, bold math letters/symbols are available
29: % at a larger range of sizes for NFSS release 1 and 2 (using \boldmath or
30: % preferably \bmath).
31: %
32: % The usenatbib command allows the use of Patrick Daly's natbib.sty for
33: % cross-referencing.
34: %
35: % If you wish to typeset the paper in Times font (if you do not have the
36: % PostScript Type 1 Computer Modern fonts you will need to do this to get
37: % smoother fonts in a PDF file) then uncomment the next line
38: % \usepackage{Times}
39: 
40: \newcommand{\bu}{\mbox{\boldmath $u$}}
41: \newcommand{\bg}{\mbox{\boldmath $g$}}
42: \newcommand{\be}{\mbox{\boldmath $e$}}
43: \newcommand{\br}{\mbox{\boldmath $r$}}
44: \newcommand{\bB}{\mbox{\boldmath $B$}}
45: \newcommand{\bj}{\mbox{\boldmath $j$}}
46: \newcommand{\bO}{\mbox{\boldmath $\Omega$}}
47: 
48: 
49: \newcommand{\ptl}{\partial}
50: \newcommand{\dd}{{\rm d}}
51: \def\div{{\mathbf \nabla \cdot}}
52: \def\curl{{\mathbf \nabla \times}}
53: \def\grad{{\mathbf \nabla}}
54: 
55: \def\ephi{{ \hat{\be}_{\phi}}}
56: \def\er{{ \hat{\be}_r}}
57: \def\etheta{{ \hat{\be}_{\theta}}}
58: \def\es{{ \hat{\be}_s}}
59: \def\ez{{ \hat{\be}_z}}
60: \def\ex{{ \hat{\be}_x}}
61: \def\ey{{ \hat{\be}_y}}
62: \def\rc{r_{\rm c}}
63: \def\rt{r_{\rm t}}
64: \def\rin{r_{\rm in}}
65: \def\rout{r_{\rm out}}
66: \def\rsun{R_{\odot}}
67: 
68: \def\ur{u_r}
69: \def\ut{u_{\theta}}
70: \def\up{u_{\phi}}
71: \def\Br{B_r}
72: \def\Bt{B_{\theta}}
73: \def\Bp{B_{\phi}}
74: \def\Pr{{\rm Pr}}
75: 
76: 
77: \def\st{\sin\theta}
78: \def\ct{\cos\theta}
79: \def\s2t{\sin^2\theta}
80: \def\c2t{\cos^2\theta}
81: \def\Oc{\Omega_{\rm c}}
82: \def\Oeq{\Omega_{\rm eq}}
83: \def\Oin{\Omega_{\rm in}}
84: 
85: \def\Ot{\tilde{\Omega}}
86: \def\Tt{\tilde{T}}
87: \def\Pt{\tilde{P}}
88: \def\rhot{\tilde{\rho}}
89: 
90: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91: 
92: \title[Dynamics of the solar tachocline]{Dynamics of the solar tachocline -- II: the stratified case}
93: \author[P. Garaud \& J.-D. Garaud]{P. Garaud$^{1}$\thanks{E-mail:
94: pgaraud@ams.ucsc.edu} \& J.-D. Garaud$^{2}$ \\
95: $^{1}$Department of Applied Mathematics and Statistics, Baskin School of Engineering, University of California Santa Cruz \\
96: $^{2}$Centre des Mat\'eriaux, Mines Paristech/CNRS UMR 7633, Evry, France}
97: 
98: \begin{document}
99: 
100: \date{}
101: 
102: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2007}
103: 
104: \maketitle
105: 
106: \label{firstpage}
107: 
108: \begin{abstract}
109: We present a detailed numerical study of the Gough \& McIntyre 
110: model for the solar tachocline. This model explains the uniformity 
111: of the rotation profile observed in the bulk of the radiative zone 
112: by the presence of a large-scale primordial magnetic 
113: field confined below the tachocline 
114: by flows originating from within the convection zone. 
115: We attribute the failure of previous 
116: numerical attempts at reproducing even qualitatively Gough \& McIntyre's 
117: idea to the use of boundary conditions which inappropriately model
118: the radiative--convective interface. We emphasize 
119: the key role of flows downwelling from the convection zone in confining the 
120: assumed internal field. We carefully select the range of 
121: parameters used in the simulations to guarantee a faithful representation 
122: of the hierarchy of expected lengthscales. We then present, for the first
123: time, a fully nonlinear and self-consistent 
124: numerical solution of the Gough \& McIntyre model which 
125: qualitatively satisfies the following set of observational constraints:
126: (i) the quenching of the large-scale differential rotation below the tachocline
127:  -- including in the polar regions -- as seen by helioseismology (ii) 
128: the confinement of the large-scale meridional flows to the uppermost 
129: layers of the radiative zone as required by observed light element abundances and suggested by helioseismic sound-speed data.
130: \end{abstract}
131: 
132: \begin{keywords}
133: MHD -- Sun:magnetic fields -- Sun:interior -- Sun:rotation
134: \end{keywords}
135: 
136: \section{Introduction}
137: 
138: The presence of the tachocline, a thin shear layer located at the interface 
139: between the radiative and convective regions of the Sun, was established two 
140: decades ago (Christensen-Dalsgaard \& Schou, 1988; 
141: Kosovichev, 1988; Brown {\it et al.} 1989; Dziembowski {\it et al.} 1989) 
142: but its {\it modus operandi} still remains mysterious. 
143: 
144: Anisotropic turbulent stresses associated with rotationally 
145: constrained eddies are thought to maintain 
146: the differential rotation profile observed within the convection zone:
147: \begin{equation}
148: \Omega_{\rm cz}(\theta,r) \simeq \Omega_{\rm eq}(r) (1-a_2(r) \cos^2\theta 
149: - a_4(r) \cos^4\theta) \mbox{   ,   }
150: \label{eq:ocz}
151: \end{equation}
152: where for example at $r = 0.75\rsun$ $\Omega_{\rm eq}/2\pi  
153: = 463 {\rm nHz}$, $a_2 = 0.17$ and $a_4  = 0.08$ (Schou {\it et al.} 
154: 1998; Gough 2007). However, as shown by Spiegel \& Zahn (1992) 
155: (SZ92 hereafter), 
156: the mere reduction in the amplitude of these stresses naturally 
157: expected to occur across the radiative--convective interface 
158: (at $r_{\rm cz} = 0.713 \rsun$) cannot explain the transition to 
159: near-uniform rotation below.
160: It was later argued by Gough \& McIntyre (1998) (GM98 hereafter) 
161: that only {\it long-range} stresses can explain the suppression of 
162: the rotational shear in the bulk of the radiative zone 
163: and in addition maintain the angular velocity of the interior (observed to be 
164: $\Omega_{\rm rz}/2\pi \simeq 430$nHz) close to that of the surface despite 
165: the global spin-down induced by magnetic-braking.
166: 
167: Two competing theories for these presumed long-range stresses have been 
168: investigated: purely hydrodynamic stresses in the form of
169: gravity waves (see the review by 
170: Zahn, 2007) and hydromagnetic stresses (see the review by 
171: Garaud, 2007). This paper focuses on the latter mechanism only, although the 
172: dynamical balance in the Sun could arguably involve a combination of the two. 
173: 
174: It has long been known that even a very small primordial magnetic field 
175: embedded in the 
176: radiative zone could in principle impose uniform rotation 
177: (Ferraro, 1937; Mestel, 1953; Mestel \& Weiss, 1987). Ferraro's 
178: isorotation law,
179: \begin{equation}
180: \bB \cdot \grad \Omega = 0 \mbox{   ,   }
181: \end{equation} 
182: valid in the limit of negligible dissipation and for steady-state, 
183: axisymmetric flows, is usually stated as {\it the angular velocity 
184: must be constant on magnetic field lines}. Thus in its simplest form, 
185: Ferraro's law predicts the possibility of uniform rotation for the 
186: radiative zone provided the magnetic field is entirely confined beneath the
187: radiative--convective interface (R\"udiger \& Kitchatinov 1997). 
188: On the other hand, any field line directly connected with 
189: the convection zone promotes the propagation of the rotational shear into
190: the radiative zone (MacGregor \& Charbonneau 1999), 
191: inducing what will be referred to from 
192: here on as a ``differentially rotating Ferraro state''. 
193: Hence, field confinement is the key to the existence of a tachocline.
194: 
195: GM98 were the first to address the question of how
196: the presumed primordial field could indeed be confined within 
197: the radiative zone, and proposed
198: that meridional flows driven by Coriolis forces in the convection zone 
199: and burrowing downward would interact nonlinearly with the underlying
200: magnetic field lines,  bending them towards the horizontal in the
201: tachocline region, thus effectively suppressing direct radial
202: Alfv\'enic transport between the  convection zone and the radiative
203: zone. Conveniently, the same meridional flows can also be held
204: responsible for mixing light elements such as  Li and Be between the
205: convection zone and their respective nuclear-burning  regions, 
206: reducing He settling (Elliott \& Gough, 1999) as well
207: as providing weak but sufficient angular-momentum  transport to
208: adjust the mean rotation rates of the convective and
209: radiative zones continuously throughout the spin-down phase. The seminal
210: boundary-layer  analysis of the dynamics of the tachocline presented
211: by GM98 validated the plausibility of this theory, although
212: many of their simplifying scaling assumptions remain to be confirmed
213: through the direct numerical solution of the governing equations.
214: 
215: This paper first briefly reviews existing work and then 
216: presents new results on the laminar tachocline dynamics according to GM98. 
217: We begin by discussing past attempts at implementing 
218: their model numerically in Section \ref{sec:prevfail}. 
219: In particular, we argue that the failure of previous numerical studies 
220: in reproducing field confinement can be explained by 
221: the selection of inappropriate boundary conditions. A new numerical 
222: algorithm is then 
223: presented in Section \ref{sec:model}, and used in Section \ref{sec:numexp} 
224: to revisit the idea proposed by GM98 with much more success. 
225: We discuss future prospects in Section \ref{sec:disconcl}.
226: 
227: \section{Discussion of previous numerical models in the light of 
228: Gough \& McIntyre's original idea}
229: \label{sec:prevfail}
230: 
231: GM98 argue that the radiative interior should be
232: divided into three dynamically distinct zones including (from the base of 
233: the convection zone downward) (1) a more-or-less magnetic-free region in 
234: thermal-wind balance, well-ventilated by meridional flows originating 
235: from the convection zone, which can be 
236: thought of as the bulk of the tachocline, (2) a very thin magnetic 
237: advection-diffusion layer where the tachocline flows and the underlying 
238: field interact nonlinearly to confine one another and (3) a magnetically 
239: dominated, near-uniformly rotating interior. 
240: 
241: The nonlinear nature and geometric complexity of the problem precludes 
242: analytical solutions, and all existing studies since the 
243: original work of GM98 have been numerical (see the review by Garaud, 2007). 
244: Among these, only two include all the nonlinear terms required 
245: (i.e. the advection terms 
246: in the momentum and the magnetic induction equations, and the Lorentz 
247: force in the momentum equation)
248: to represent the nonlinear magnetohydrodynamics of the tachocline 
249: ``correctly'': Garaud (2002) -- hereafter 
250: Paper I -- and Brun \& Zahn (2006) -- hereafter BZ06. Surprisingly, neither 
251: have been able to find evidence for the kind of dynamical balance proposed by 
252: GM98, and the time is therefore ripe to take a step back and discuss why.
253: 
254: \subsection{``Failure'' of previous numerical models}
255: 
256: Paper I and BZ06 are two studies which model magnetohydrodynamic
257: perturbations induced in the solar radiative zone by 
258: the differentially rotating convection zone and by an assumed 
259: primordial magnetic field. Both papers focus on the 
260: issues of field confinement and the suppression of differential 
261: rotation below the tachocline. 
262: 
263: Paper I presents steady-state, axially symmetric solutions of 
264: an incompressible and
265: unstratified version of the GM98 model. BZ06 on the other
266: hand use a time-dependent, three-dimensional 
267: algorithm based on the ASH code (Glatzmaier 1984; Clune {\it et al.} 1999; 
268: Miesch {\it et al.} 2000; Brun {\it et al.} 2004), and solve 
269: the complete set of anelastic MHD equations. 
270: 
271: Both studies otherwise consider 
272: a similar computational domain, namely a spherical shell which 
273: spans the region between the base of the 
274: convection zone (at $r = r_{\rm out}\simeq r_{\rm cz}$) and 
275: an inner sphere (at $r = r_{\rm in}$). The boundary conditions are also 
276: essentially similar: the ``outer'' boundary, 
277: which effectively models the radiative--convective interface,
278: is in both cases assumed to be impermeable and rotating differentially with 
279: an angular velocity profile given by equation (\ref{eq:ocz}); the 
280: magnetic field is matched onto a potential field. 
281: 
282: In Paper I, the numerical solutions only show confinement of the 
283: magnetic field lines in the equatorial
284: regions for low enough diffusivities,  occasionally at mid-latitudes
285: for lower field strength, but never in the  polar regions 
286: (see Figure 11 of Paper I for example). 
287: It is commonly argued that the failure of Paper I to find 
288: fully-confined magnetic field solutions stems entirely from the simplified 
289: nature of the model equations (incompressible, unstratified): indeed, 
290: within these assumptions the meridional flows are generated only by
291: Ekman-Hartmann pumping on the boundaries, and their amplitude scales with
292: the diffusivities in a way which always maintains the magnetic Reynolds number 
293: below unity. As a result, these 
294: flows are unable to confine the magnetic field and the
295: differential rotation imposed at the upper boundary of the domain 
296: persists within a large part of the radiative
297: zone.
298: 
299: The three-dimensional, time-dependent solutions of BZ06 
300: naturally depend on the initial magnetic field 
301: configuration selected. Various cases with an initial field more-or-less
302: deeply confined are discussed. Against expectations, BZ06 find that 
303: regardless of the initial conditions, the field lines always
304: spread out and eventually overlap with the convection zone, permitting the 
305: propagation of the differential rotation into the radiative interior. 
306: Interestingly, the transient state of the system -- prior to any field line
307: connecting with the convection zone -- qualitatively 
308: looks in many ways similar to the 
309: well-known GM98 picture, although it is clearly not a 
310: steady state. The ``tachocline'' thus formed slowly narrows with time until 
311: field lines spread into the convection zone, at which point a 
312: differentially rotating Ferraro state is rapidly established. 
313: One is forced to conclude in one of three ways: (1) for unspecified
314: primordial reasons, the initial magnetic field was more deeply embedded in 
315: the radiative zone and the currently observed tachocline is merely a transient
316: phase; (2) the tachocline is indeed in a steady-state and some of the 
317: assumptions made in the way have been wrong; (3) the parameter regime
318: studied by BZ06 (where diffusion coefficients are artificially 
319: increased by many orders of magnitude to satisfy numerical constraints) 
320: does not appropriately reproduce the solar interior dynamics. 
321: 
322: One should be uncomfortable in selecting the first option, as
323: it would place very strong and unlikely constraints on the initial field 
324: conditions to provide just the right structure for today's tachocline. One 
325: could naively tend to favour the third option, but BZ06 {\it a priori} 
326: took care to select a range of parameters for which all expected boundary layer
327: thicknesses are small enough and in the same hierarchical order as those
328: of the model proposed by GM98. 
329: The plot thickens while we are left with the uneasy task of reconsidering
330: the key features of either the numerical models or of the 
331: GM98 model (both, perhaps). 
332: 
333: \subsection{The source of the problem}
334: \label{subsec:sourceprob}
335: 
336: At this point it is worth discussing one of the more delicate aspects
337: of the GM98 model, namely the exact mechanism by which
338: these pivotal meridional flows are thought to be generated. This point
339: has been  a source of confusion and debate, but is clearly
340: crucial to a better  understanding of the tachocline dynamics. In 
341: the original work of GM98, the principal clue to the nature of
342: the flows can be found in the sentence:  {\it ``Turbulent stresses in
343: the convection zone induce (through Coriolis  effects) a meridional
344: circulation, causing the gas from the convection zone  to burrow
345: downwards ...''}. The flows considered by GM98 do not
346: originate from within the tachocline and can therefore  not be
347: appropriately modelled  by a numerical scheme in which the
348: radiative--convective interface is assumed to be impermeable (as in Paper I 
349: and BZ06). While this conclusion seems obvious
350: in hindsight, the physics of the problem are actually rather subtle and 
351: deserve clarification. In what follows, we discuss the issue in more detail
352: and present a unified view of the results of previous works on the subject. 
353: 
354: Spiegel \& Zahn were the first to study the
355: dynamics of the newly-discovered tachocline (SZ92). They considered
356: a non-magnetic radiative zone only, and imposed a 
357: latitudinally-varying rotation profile at the radiative--convective interface. 
358: They performed two distinct calculations. The first looked at the 
359: time-dependent evolution of the angular velocity profile in the 
360: radiative zone under such forcing, assuming isotropic viscous stresses. 
361: The second sought steady-state ``tachocline'' solutions assuming anisotropic 
362: turbulent stresses. Since the latter did not specifically address 
363: the question of the meridional flows, we focus here on the results 
364: of the time-dependent calculation, which can be interpreted 
365: in the following way. The differential rotation imposed by the 
366: convection zone to the top of the radiative zone inevitably induces
367: some degree of shear along the rotation axis if the radiative zone 
368: is originally in a state of uniform rotation. This generates a thermal
369: gradient in the latitudinal direction by way of the thermal-wind equation 
370: (see GM98, equation (1) for example). Meridional flows are then 
371: required to balance this thermal gradient when the system is in 
372: thermal equilibrium.
373: These flows burrow into the radiative zone, advecting angular momentum 
374: thereby helping the propagation of the shear further down. 
375: The results of SZ92 imply that the system continues evolving in this 
376: fashion until the radiative zone has achieved complete thermal and 
377: dynamical equilibrium.
378: 
379: The characteristic amplitude of the time-dependent flows associated with this
380: thermo-dynamical relaxation process is, by way of the assumptions 
381: listed above, only dependent on the (evolving) local differential rotation, 
382: stratification and thermal conductivity. Their turnover 
383: timescale is calculated to be of the order of the local Eddington-Sweet
384: timescale, which is short in the initial relaxation stages and steadily 
385: increases as the system evolves. 
386: Crucially, this result was derived independently of any
387: boundary condition on the meridional flows. It is therefore correct 
388: to think of the induced transient flows as being driven by baroclinic stresses 
389: from {\it within} the tachocline rather than downwelling from 
390: the convection zone. In fact, they exist {\it even if} the 
391: radiative--convective interface is assumed to be 
392: impermeable. This fact is probably at the origin of the impermeable boundary 
393: conditions selected by Paper I and by BZ06, in spite the obvious contradiction 
394: with GM98's intent. 
395: 
396: However, it is vital to remember that this transient phase and its associated
397: flows both end once the system achieves thermal and dynamical equilibrium. 
398: The only flows remaining in the final steady-state 
399: are driven, within tiny boundary layers, by the thermo(-magneto)-viscous 
400: stresses required to match the bulk equilibrium 
401: solutions to the applied boundary conditions. 
402: Thus, as expected from any elliptic problem, the nature of the boundary 
403: conditions selected entirely controls the steady-state solutions. 
404: This is clearly illustrated by the work of Garaud 
405: \& Brummell (2008) (GB08 hereafter), which complements that of SZ92 by
406: calculating the spatial properties as well as characteristic 
407: amplitudes of {\it steady-state} meridional flows in the radiative zone, 
408: as induced by various kinds of forcing applied at the radiative--convective interface. 
409: 
410: GB08 showed that in the non-magnetic case, 
411: the induced steady-state flows can always be viewed -- at least in the linear 
412: sense, and for solar parameters -- 
413: as the sum of two ``modes'': a viscously-dominated 
414: ``Ekman mode'', which 
415: very rapidly decays away from the interface (on an Ekman lengthscale), 
416: and a ``thermo-viscous mode'' which can essentially span the 
417: entire radiative zone when not hindered by a magnetic field. 
418: Since the Ekman mode decays too rapidly to have
419: any effect on the tachocline dynamics, the flows discussed by 
420: GM98, which are thought to ventilate the bulk of the tachocline,
421: should be identified with the thermo-viscous mode. This view
422: is perfectly consistent with GM98's analysis, since the 
423: thermo-viscous mode is in thermal-wind balance (GB08).
424: 
425: The respective amplitudes with which the Ekman and thermo-viscous 
426: modes are driven
427: depend on the nature, amplitude and spatial structure of the forcing.
428: GB08 showed that when the interface is 
429: impermeable, as assumed in Paper I and by BZ06, then the amplitude 
430: of the thermo-viscous mode is negligible for microscopic solar values 
431: of the diffusivities. As a result, the magnetic Reynolds number of these
432: tachocline flows is much smaller than unity, which straightforwardly 
433: explains why field confinement eluded these two previous studies.
434: 
435: When flows are pumped directly through the interface by stresses 
436: within the convection zone (i.e. the radial component of the flows 
437: is non-zero at the radiative--convective interface), GB08
438: find that the amplitude of the thermo-viscous mode can be much higher, 
439: in which case the tachocline flows may be expected to confine the field 
440: in a scenario qualitatively similar to the one proposed by GM98. 
441: The quantitative detailed analysis must be done numerically; 
442: this is the purpose of the present study.
443: 
444: \section{The numerical model}
445: \label{sec:model}
446: 
447: We now revisit the idea proposed by GM98 in the light
448: of the previous discussion. A new numerical algorithm has been constructed
449: specifically for this purpose. We first present, for the sake of completeness, 
450: a derivation of our model equations in Section \ref{subsec:numeqs}. 
451: Since our computational domain is limited 
452: to a spherical shell within the radiative zone, we then present and 
453: discuss in detail all the boundary conditions applied in Section 
454: \ref{subsec:numbcs}. The relevant non-dimensional parameters, and 
455: expected boundary layers of the problem are discussed in Section 
456: \ref{subsec:bls}. The numerical method, as well as other numerical constraints, 
457: are finally outlined in Section \ref{subsec:numalgo}.
458: 
459: \subsection{The governing equations}
460: \label{subsec:numeqs}
461: 
462: The Standard Solar Model (SSM hereafter) provides an accurate representation 
463: of the thermal structure and composition of the interior of a hypothetical 
464: spherically symmetric, non-rotating, non-magnetic Sun. The excellent
465: match between the SSM and helioseismic observations  
466: confirms that the likely internal dynamics of the Sun only induce weak 
467: deviations in the background state thermodynamical quantities. 
468: The large-scale magnetohydrodynamics 
469: of the solar interior, thought to be responsible for the maintenance of 
470: the peculiar observed rotation profile, can therefore be thought of 
471: as perturbations upon the SSM. 
472: 
473: The quasi-steady SSM equations (which are valid on timescales
474: much shorter than the nuclear burning timescale) reduce to the following
475: set of four equations within the solar radiative zone (but outside of the
476: nuclear burning core):
477: \begin{eqnarray}
478: && - \grad \overline{p} = \overline{\rho} \grad \overline{\Phi} \mbox{   ,   } \nonumber  \\
479: && \div( \overline{k} \grad \overline{T}) = 0   \mbox{   ,   }\nonumber  \\
480: && \overline{p} = \overline{p}(\overline{\rho},\overline{T}) \mbox{   ,   } \nonumber  \\
481: && \grad^2 \overline{\Phi} = 4\pi G \overline{\rho}  \mbox{   .   } 
482: \label{eq:backg} 
483: \end{eqnarray}
484: These equations describe hydrostatic equilibrium, thermal 
485: equilibrium, the equation of state and finally the Poisson equation for the 
486: gravitational potential respectively. Here, $\rho$, $T$, and $p$ 
487: are the standard
488: thermodynamical variables, $k$ is the thermal conductivity 
489: (which typically depends on $T$ and $\rho$), $G$ is the gravitational constant 
490: and $\Phi$ is the gravitational potential. The equation of state itself 
491: is well-approximated by that of a perfect gas in this region of the Sun. 
492: The solution to the set of equations (\ref{eq:backg}) for the present-day Sun
493: can be inferred from model S  of Christensen-Dalsgaard {\it et al.} (1991) for example 
494: (see Appendix A).
495: 
496: 
497: We then consider perturbations on this background equilibrium 
498: caused by meridional and azimuthal flows, as well as magnetic fields.  
499: For simplicity, we restrict our study to axisymmetric systems. Moreover, 
500: since the 
501: observed internal rotation profile of the Sun has remained approximately 
502: constant over the past decade, we boldly 
503: postulate that the Sun is in fact in a 
504: state of quasi-steady dynamical and thermal balance. 
505: The general set of equations governing the 
506: axisymmetric, quasi-steady perturbations are then 
507: \begin{eqnarray}
508: && \rho \bu \cdot \grad \bu 
509: = - \grad p - \rho \grad \Phi + \bj \times \bB + \div \Pi \mbox{   ,   } \nonumber \\
510: &&  \div(\rho \bu) = 0 \mbox{   ,   } \nonumber  \\
511: &&  \rho T \bu\cdot \grad s  
512: = \div( k \grad T)\mbox{   ,   } \nonumber  \\
513: && p = p(\rho,T) \mbox{   ,   }\nonumber \\
514: && \grad^2 \Phi = 4\pi G \rho \mbox{   ,   } \nonumber \\
515: && \curl(\bu \times \bB) = \curl( \eta \curl \bB)\mbox{   ,   } 
516: \nonumber  \\
517: && \div \bB = 0\mbox{   ,   }
518: \label{eq:global}
519: \end{eqnarray}
520: where $\bu$ represents the axisymmetric flow velocity, 
521: $\bB$ the magnetic field and $\bj = \curl \bB / 4\pi $ is the electric current, $\Pi$ is the 
522: viscous stress tensor, $s$ is the specific entropy
523: and $\eta$ is the magnetic diffusivity. 
524: These equations respectively characterise the conservation of
525:  momentum, mass, and thermal
526: energy, the equation of state and the Poisson equation, conservation of 
527: magnetic flux and finally, the solenoidal condition. 
528: 
529: Thermodynamical perturbations are from here on denoted with tildes, 
530: as for example in $p = \overline{p} 
531: + \tilde{p}$. The system of equations (\ref{eq:global}) can be 
532: linearised in the thermodynamical quantities 
533: around the background state equations (\ref{eq:backg}) to yield
534: \begin{eqnarray}
535: && \overline{\rho} \bu \cdot \grad \bu 
536: = - \grad \tilde{p} - \overline{\rho} \grad \tilde{\Phi}  - \tilde{\rho} \grad \overline{\Phi}  + \bj \times \bB + \div \Pi \mbox{   ,   } \nonumber \\
537: &&  \div(\overline{\rho} \bu) = 0 \mbox{   ,   } \nonumber  \\
538: && \overline{\rho} \overline{T} \bu\cdot \grad \overline{s}  
539: = \div( \overline{k} \grad \tilde{T} ) \mbox{   ,   }
540: %= \div( \overline{k} \grad \tilde{T} + \tilde{k} \grad \overline{T}) 
541: \nonumber  \\
542: && \frac{\tilde{p}}{\overline{p}} = \frac{\tilde{\rho}}{\overline{\rho}} +  \frac{\tilde{T}}{\overline{T}}\mbox{   ,   }\nonumber \\
543: && \grad^2 \tilde{\Phi} = 4\pi G \tilde{\rho} \mbox{   ,   }\nonumber \\
544: && \curl(\bu \times \bB) = \curl( \overline{\eta} \curl \bB) \mbox{   ,   }
545: \nonumber  \\
546: && \div \bB = 0\mbox{   ,   }
547: \label{eq:globallin}
548: \end{eqnarray}
549: where we approximated the equation of state with that of a perfect gas, 
550: and neglected perturbations to the chemical composition. Note that the 
551: latter approximation was chosen for the sake of simplicity: Wood \& McIntyre 
552: (2007) showed that the effects of compositional gradients (notably Helium) 
553: could play an important role in the tachocline dynamics. The 
554: background magnetic diffusivity $\overline{\eta}$, kinematic viscosity 
555: $\overline{\nu}$ and thermal conductivity $\overline{k}$ are calculated
556: from model S using the expressions provided by Gough (2007) (see Appendix A).
557: 
558: Note that the nonlinearities in the quantities relating to the 
559: magnetohydrodynamics of the interior (flow velocities and magnetic field) 
560: have not yet been tampered with\footnote{Also note that in principle, the thermal 
561: energy equation should contain an additional term, namely 
562: $\tilde{k} \grad \overline{T}$. This term is included in the numerical 
563: algorithm for the sake of completeness, but in practise has little influence 
564: on the result (at least in the case of the solar radiative zone). 
565: It will not be discussed in this paper for the sake of simplicity.}.
566: The next step is to use a spherical coordinate system ($r,\theta,\phi$) and 
567: move to a rotating frame of reference, by considering 
568: that $ \bu = \overline{\bu} + \tilde{\bu}$, with $\overline{\bu} = 
569: r \sin\theta \overline{\Omega} \ephi$ and $\tilde{\bu} = (\tilde{u}_r,\tilde{u}_\theta,\tilde{u}_\phi$). 
570: The value of $\overline{\Omega}$ 
571: adopted is uniquely defined by requiring that the total angular momentum of 
572: the convection zone be null in the rotating frame (Gilman {\it et al.}, 
573: 1989): if $\Omega_{\rm cz}(\theta,r)$ is assumed to be independent of radius, 
574: and given by equation (\ref{eq:ocz}), then 
575: \begin{equation}
576: \overline{\Omega} = \Omega_{\rm eq} \left( 1 - \frac{a_2}{5} - \frac{3a_4}{35} \right) \mbox{  . }
577: \end{equation}
578: The equation for the conservation of momentum then becomes
579: \begin{eqnarray}
580: && \overline{\rho} \tilde{\bu} \cdot \grad \tilde{\bu} +  2\overline{\rho} 
581: \overline{\bO} \times \tilde{\bu} + \overline{\rho} \overline{\bO} 
582: \times \overline{\bO} \times \br  \nonumber \\ 
583: = && 
584: - \grad \tilde{p} - \overline{\rho} \grad \tilde{\Phi}  - \tilde{\rho} 
585: \grad \overline{\Phi}  + \bj \times \bB + \div \Pi \mbox{   ,   }
586: \label{eq:rotating}
587: \end{eqnarray}
588: while the assumption of axial symmetry implies that the equations for the conservations of mass, 
589: thermal energy and magnetic flux are merely changed by replacing $\bu$ with $\tilde{\bu}$.
590: 
591: It is well known that the baroclinic deformation of the background state 
592: caused by the centrifugal force drives meridional flows, albeit very slow 
593: ones since the Sun is a slow rotator (Eddington, 1925; Sweet, 1950). Since 
594: the global turnover timescale of this Eddington-Sweet circulation is 
595: calculated to be orders of magnitude longer than the age of the Sun, 
596: it cannot play any role in the angular momentum or chemical mixing processes. 
597: In a steady-state calculation, however, these flows play an artificially 
598: important role and it is important to suppress 
599: them. This can be done by using the following momentum conservation 
600: equation instead (SZ92):
601: \begin{equation}
602: \overline{\rho} \tilde{\bu} \cdot \grad \tilde{\bu} + 
603: 2\overline{\rho} \overline{\bO} \times \tilde{\bu}   =
604: - \grad \tilde{p} - \tilde{\rho} 
605: \grad \overline{\Phi}  + \bj \times \bB + \div \Pi \mbox{   ,   }
606: \label{eq:rotating2}
607: \end{equation}
608: where Cowling's approximation was also 
609: used to justify neglecting the perturbation 
610: of the gravitational potential related to density and temperature perturbations
611: caused by the remaining Coriolis and Lorentz forces.
612: 
613: Finally, we assume that the meridional flow velocities remain 
614: small (SZ92; GM98): we neglect quadratic terms in $\tilde{u}_r$ 
615: and $\tilde{u}_\theta$, but retain nonlinearities in 
616: $\tilde{u}_{\phi}$ to allow for significant differential rotation. 
617: 
618: To summarise, the model equations we consider are 
619: \begin{eqnarray}
620: && \overline{\rho} \tilde{\bu} \cdot \grad \tilde{\bu} + 2 \overline{\rho} \overline{\bO} \times \tilde{\bu}   =
621: - \grad \tilde{p} - \tilde{\rho} \grad \overline{\Phi}  
622: + \bj \times \bB + \div \Pi \mbox{   ,   }\nonumber  \\
623: &&  \div(\overline{\rho} \tilde{\bu}) = 0 \mbox{   ,   } \nonumber  \\
624: && \overline{\rho} \overline{T} \tilde{\bu}\cdot \grad \overline{s}  
625: = \div( \overline{k} \grad \tilde{T})\mbox{   ,   } \nonumber  \\
626: %= \div( \overline{k} \grad \tilde{T} + \tilde{k} \grad \overline{T}) \nonumber  \\
627: %&& \mbox{   where   } \frac{\tilde{k}}{\overline{k}} = 6.5 \frac{\tilde{T}}{\overline{T}} - 2\frac{\tilde{\rho}}{\overline{\rho}} \nonumber \\
628: && \frac{\tilde{p}}{\overline{p}} = \frac{\tilde{\rho}}{\overline{\rho}} +  \frac{\tilde{T}}{\overline{T}}\mbox{   ,   }\nonumber \\
629: && \curl(\tilde{\bu} \times \bB) = \curl( \overline{\eta} \curl \bB) \mbox{   ,   }
630: \nonumber  \\
631: && \div \bB = 0\mbox{   ,   }
632: \label{eq:global2}
633: \end{eqnarray}
634: where the quantities solved for are the three components of $\tilde{\bu}$ (as defined above), the three 
635: components of $\bB = (B_r,B_\theta,B_\phi)$, and the three thermodynamical variables $\tilde{p}$, 
636: $\tilde{\rho}$ and $\tilde{T}$, and where quadratic terms in $\tilde{u}_r^2$, 
637: $\tilde{u}_\theta^2$ and $\tilde{u}_r\tilde{u}_\theta$ are neglected (see Appendix B for the full set of 
638: equations expanded in a spherical coordinate system). 
639: 
640: This set of equations is equivalent to the one used by GM98 before 
641: they apply their boundary-layer analysis. It also 
642: reduces to the one used by SZ92 in the non-magnetic case,
643: provided the following further approximations are applied: the Boussinesq 
644: approximation, assuming that $\tilde{\Omega} \ll \overline{\Omega}$ and 
645: that $\tilde{u}_\theta \gg \tilde{u}_r$. Finally, it is a steady-state and
646: axisymmetric version of the equations used by BZ06.
647: 
648: For comparison, note that R\"udiger \& Kitchatinov (1997), 
649: MacGregor \& Charbonneau (1999), and Kitchatinov \& R\"udiger (2006) neglect 
650: meridional flows entirely in their calculations of the structure of the 
651: radiative interior\footnote{although in work of Kitchatinov \& R\"udiger (2006), the effect 
652: of the flows does influence the boundary conditions applied to the magnetic field.} Sule, 
653: Arlt \& R\"udiger (2005) include 
654: meridional flows but assume that the poloidal field is fixed, so that the
655: interaction between the field and the flows through the Lorentz force in their 
656: paper is fundamentally linear (and therefore inappropriate).  
657: 
658: \subsection{Domain boundaries and boundary conditions}
659: \label{subsec:numbcs}
660: 
661: \subsubsection{Computational domain}
662: 
663: The geometry of the
664: computational domain is shown in Figure \ref{fig:modelfig}. 
665: The upper boundary is selected to be slightly below the base of 
666: the convection zone, at $r_{\rm out} = 0.7\rsun$. 
667: The lower boundary is at $r_{\rm in} =
668: 0.35\rsun$ as in BZ06. The inner core (for $r < r_{\rm in}$) 
669: can be viewed as a solid metallic sphere, within which a
670: permanent magnetic dipole is maintained. The selection of the position of the
671: lower boundary was found to have little influence on the tachocline dynamics 
672: for low enough values of the diffusivities.
673: \begin{figure}
674: \epsfig{file=f1.eps,width=6cm}
675: \caption{The computational domain is limited to the spherical shell
676: shown  here in grey.}
677: \label{fig:modelfig}
678: \end{figure}
679: 
680: 
681: \subsubsection{Model boundary conditions}
682: \label{subsec:modelbc}
683: 
684: The selection of adequate boundary conditions for this model 
685: is a very delicate task. Given the elliptic and nonlinear nature of the system 
686: considered, the existence of solutions is not guaranteed, and solutions
687: (when they exist) are entirely controlled by the boundary conditions applied. 
688: 
689: Our aim is to propose a set of boundary conditions at the upper
690: boundary of the computational domain which reproduces most faithfully the
691: influence of the convection zone on the radiative zone, while those 
692: near the lower boundary are selected in such a way as to impose the 
693: required dipolar field, but otherwise be as inconspicuous as possible.
694: 
695: The set of coupled partial differential equations described in 
696: (\ref{eq:global2}) and expanded in Appendix B represents a 
697: 12th order system, and therefore requires 12 independent boundary conditions.
698: \\ 
699: \\
700: {\it Bottom boundary conditions.} The bottom boundary of the computational domain is located at 
701: $r_{\rm in}$, and should be thought of as the interface between the modelled 
702: fluid and an electrically and thermally conducting solid sphere 
703: (hereafter ``the core''). Impermeability implies
704: \begin{equation}
705: \tilde{u}_r = 0  \mbox{   at   } r = r_{\rm in} \mbox{   .   }
706: \end{equation}
707: We generally impose no-slip boundary conditions at $r_{\rm in}$ (except when 
708: specifically mentioned), so that 
709: $\tilde{u}_\theta = 0$ and $\tilde{u}_\phi = r_{\rm in} \sin \theta \tilde{\Omega}_{\rm in}$, 
710: where $\tilde{\Omega}_{\rm in}$ is the constant angular velocity of the core in the rotating 
711: frame. The value of $\tilde{\Omega}_{\rm in}$ is selected in such a way as to guarantee that 
712: the total torque applied to the core be zero:
713: \begin{equation}
714: \int_{-\pi/2}^{\pi/2} \left( \overline{\rho} \overline{\nu} r^2_{\rm in} \sin^2 \theta \frac{\partial \tilde{\Omega}}{\partial r} + r_{\rm in} \sin\theta \frac{B_r B_\phi}{4\pi} \right) \sin\theta \dd \theta = 0 \mbox{   .  }
715: \end{equation} 
716: 
717: Assuming that the core is a homogeneous thermally 
718: conducting solid implies that the temperature fluctuations within 
719: satisfy Laplace's equation
720: \begin{equation}
721: \grad^2 \tilde{T}^{\rm in} = 0 \mbox{   .  }
722: \end{equation}
723: The regular solution can be expanded upon standard 
724: eigenfunctions, namely
725: \begin{equation}
726: \tilde{T}^{\rm in}(r,\theta) = \sum_{n = 0}^{\infty} a^{\rm in}_n P_n\left(\cos \theta\right) r^n \mbox{   ,}
727: \label{eq:Tin}
728: \end{equation}
729: where $P_n$ is the $n$-th order Legendre polynomial, and the set $\{a_n^{\rm in}\}$ are integration constants.
730: Requiring the simultaneous continuity of $\tilde{T}$ and its derivative 
731: at $r = r_{\rm in}$ while satisfying equation (\ref{eq:Tin}) constrains the 
732:  $\{a_n^{\rm in}\}$ and yields a unique relationship between $\tilde{T}$ and 
733: $\partial \tilde{T}/ \partial r$, which is used as a boundary condition 
734: on temperature.  
735: 
736: Two boundary conditions for the magnetic field are required. They are 
737: obtained in a similar fashion to that of the temperature fluctuations, by 
738: assuming that the core
739: is a homogeneous electrically conducting solid. In that case 
740: the magnetic field within satisfies
741: \begin{equation}
742: \grad^2 \bB^{\rm in} = 0 \mbox{   ,   }
743: \end{equation}
744: which, by axial symmetry, is equivalent to 
745: \begin{equation} 
746: (\grad^2 \bB^{\rm in})_\phi = 0 \mbox{   and   } (\curl \bB^{\rm in})_\phi = 0 \mbox{   .  }
747: \end{equation}
748: Defining a flux function $\chi$ as 
749: \begin{equation}
750: \bB^{\rm in}_{\rm p} = \curl\left( \chi^{\rm in} \ephi \right) 
751: \end{equation}
752: implies that both $B^{\rm in}_\phi$ and $\chi^{\rm in}$ satisfy the same equation, 
753: namely
754: \begin{equation}
755: \grad^2 B^{\rm in}_\phi - \frac{B^{\rm in}_\phi}{r^2 \sin^2\theta} = 0 \mbox{  , similarly for   } \chi^{\rm in}\mbox{   .   }
756: \end{equation}
757: The regular solution for $B^{\rm in}_\phi$ can be expanded as
758: \begin{equation}
759: B_\phi^{\rm in}(r,\theta) = \sum_{n=1}^\infty b^{\rm in}_n \frac{\partial}{\partial \theta} \left( P_n(\cos\theta)\right) r^n\mbox{   ,   }
760: \label{eq:Bphiin}
761: \end{equation} 
762: where as before the $\{b_n^{\rm in}\}$ are integration constants.
763: Requiring the simultaneous continuity of $B_\phi$ and its derivative 
764: at $r = r_{\rm in}$ while satisfying equation (\ref{eq:Bphiin}) 
765: yields a unique relationship between $B_\phi$ and 
766: $\partial B_\phi/ \partial r$, which is used as one of the two required 
767: boundary conditions on the magnetic field.
768: 
769: A point-dipole located at $r=0$ is required 
770: to maintain the primordial magnetic field in this steady-state study. Note that
771: this implicitly assumes that the dynamics of the tachocline occur
772: on timescales much shorter than the global magnetic diffusion time 
773: (which is a reasonable assumption since the global magnetic diffusion 
774: timescale is of the order of the age of the Sun). To match 
775: $\bB^{\rm in}_{\rm p}$ to a point dipole of 
776: amplitude $B_0$ at $r=r_{\rm in}$ and $\theta =0$ (on the polar axis), we 
777: require that
778: \begin{equation}
779: \chi^{\rm in} = B_0 \frac{r^3_{\rm in}}{2} \frac{\st}{r^2} + \sum_{n=1}^\infty c^{\rm in}_n \frac{\partial}{\partial \theta} \left( P_n(\cos\theta)\right) r^n\mbox{   .   }
780: \label{eq:chiin}
781: \end{equation}
782: Requiring the simultaneous continuity of $B_r$ and $B_\theta$  
783: at $r = r_{\rm in}$ provides a relationship between these
784: two quantities, which is used as the second of the two required
785: boundary conditions on the magnetic field.  \\ 
786: \\ 
787: {\it Top boundary conditions.} The choice of boundary conditions near 
788: the top boundary should model  the effects of the convective zone dynamics 
789: on the radiative--convective interface.
790: 
791: Requiring the continuity of the angular velocity profile across the interface 
792: is a natural choice for the boundary condition on $\tilde{u}_{\phi}$, 
793: although one must bear in mind that it is intrinsically equivalent to 
794: assuming that the interface is a no-slip boundary. For consistency, 
795: a similar condition should then also be imposed on the other 
796: component of the meridional flow parallel to the boundary, $\tilde{u}_\theta$. 
797: Hence we select
798: \begin{eqnarray}
799: \tilde{u}_\theta(r_{\rm out}, \theta)  &=& u^{\rm out}_\theta(\theta) \mbox{   ,   }\nonumber \\
800: \tilde{u}_\phi(r_{\rm out}, \theta) &=& r_{\rm out}\st \left(\Omega_{\rm cz}(\theta)-\overline{\Omega}\right)\mbox{   ,   }
801: \end{eqnarray}
802: where $u^{\rm out}_\theta(\theta)$ could be any desired latitudinal velocity
803: profile, and $\Omega_{\rm cz}$ is given by equation (\ref{eq:ocz}).
804: 
805: Boundary conditions are also needed on the vertical velocity $\tilde{u}_r$. 
806: In previous models of the radiative zone (Paper I; BZ06), the 
807: radiative--convective interface was assumed to be impermeable. However, 
808: this boundary condition is inadequate as a model of a (permeable) 
809: fluid interface (see Section \ref{subsec:sourceprob}). Instead, we 
810: consider the more general distribution: 
811: \begin{equation}
812: \tilde{u}_r (r_{\rm out}, \theta)  = u^{\rm out}_r(\theta) \mbox{   ,   }
813: \end{equation}
814: where the profile selected for $u^{\rm out}_r(\theta)$ is discussed in 
815: more detail in Section \ref{subsubsec:urforcing}.
816: 
817: The temperature boundary condition near the upper boundary is selected 
818: to model a very efficiently conducting fluid, by requiring that 
819: %in such a way as to maintain a null perturbed heat flux at every point of the
820: %boundary (where advection and diffusion of heat cancel out): 
821: %\begin{equation}
822: %\overline{k} \frac{\partial \tilde{T}}{\partial r} = \overline{\rho} \overline{h} u_r
823: %\end{equation} 
824: %where $\overline{h}$ is the background enthalpy. This boundary conditions is discussed in detail by GB08.
825: \begin{equation}
826: \nabla^2 \tilde{T}^{\rm out} = 0
827: \end{equation}
828: outward of the computational domain. Matching $\tilde{T}$ to $\tilde{T}^{\rm out}$ 
829: provides a unique relationship between $\tilde{T}$ and $\partial \tilde{T}/\partial r$ 
830: at the outer boundary. 
831: 
832: The boundary conditions for the magnetic field are obtained by assuming that 
833: the convection zone is an infinitely diffusive medium (so that $\nabla^2 \bB^{\rm out} = 0$ 
834: outward of $r_{\rm out}$) and selecting the solution which decays as $r \rightarrow \infty$: 
835: \begin{eqnarray}
836: && B_\phi^{\rm out}(r,\theta) = \sum_{n=1}^\infty b^{\rm out}_n \frac{\partial}{\partial \theta} \left( P_n(\cos\theta)\right) r^{-(n+1)} \mbox{   ,   }\nonumber \\
837: && \chi^{\rm out}(r,\theta) = \sum_{n=1}^\infty c^{\rm out}_n \frac{\partial}{\partial \theta} \left( P_n(\cos\theta)\right) r^{-(n+1)} \mbox{   .   }
838: \end{eqnarray}
839: Matching these solutions at $r = r_{\rm out}$ with the solution for $\bB$ within the domain 
840: provides relationships between $B_\phi$ and its derivative, and between $B_r$ and $B_\theta$ 
841: respectively.
842: 
843: \subsection{Non-dimensional parameters, and the nature of boundary layers}
844: \label{subsec:bls}
845: 
846: \subsubsection{Enhanced diffusivities}
847: \label{subsubsec:diffs}
848: 
849: The typical solutions of the set of equations and boundary conditions 
850: described in Sections \ref{subsec:numeqs} and \ref{subsec:numbcs} 
851: are known to contain thin boundary layers and internal layers of 
852: various kinds, typically scaling as some (positive) 
853: power of the diffusivities. 
854: The limited achievable numerical resolution of the algorithm used implies 
855: that the set of equations (\ref{eq:global2}) 
856: cannot be solved for solar values of these quantities. To address the problem
857: we multiply the diffusivities $\overline{\nu}(r)$, $\overline{\eta}(r)$ and 
858: \begin{equation}
859: \overline{\kappa}(r) = \frac{\overline{k}}{\overline{\rho} \overline{c}_{\rm p}} 
860: \end{equation}
861: by the enhancement factors $f_\nu$, $f_\eta$ 
862: and $f_{\kappa}$ respectively. When these factors are all selected equal 
863: to each other the Prandtl and magnetic Prandlt numbers
864: (Pr = $\overline{\nu}/\overline{\kappa}$ and Pr$_m = \overline{\nu}
865: /\overline{\eta}$ respectively) 
866: are solar everywhere. Note that we specifically do not view the numerical model
867: diffusivities as ``turbulent diffusivities''. We strive instead to derive
868: characteristic scalings of the solutions with each of the $f-$factors, and 
869: understand the asymptotic behaviour of the solutions as they are simultaneously 
870: reduced towards unity. 
871: 
872: \subsubsection{Non-dimensional parameters}
873: \label{subsubsec:params}
874: 
875: A better understanding of the scaling behaviour of the numerical solutions 
876: as $f_\nu$, $f_\eta$ and $f_\kappa$ 
877: are varied can be achieved by working with 
878: non-dimensional parameters. In what follows, we scale all distances with 
879: the solar radius $\rsun$, and all velocities with $\rsun\Omega_{\rm eq}$ 
880: (see Appendix C for details). The magnetic field strength is scaled by $B_0$, 
881: which is the amplitude of the imposed field on the inner boundary on the 
882: polar axis 
883: (see equation (\ref{eq:chiin})). Within this framework, the following
884: non-dimensional numbers naturally emerge: 
885: \begin{eqnarray}
886: && E_{\nu} = \frac{f_\nu \overline{\nu}}{\rsun^2 \Omega_{\rm eq}} = f_\nu E_\nu^{\odot} \mbox{   ,   } \nonumber \\
887: && E_{\eta} = \frac{f_\eta \overline{\eta}}{\rsun^2 \Omega_{\rm eq}} = f_\eta E_\eta^{\odot} \mbox{   ,   } \nonumber \\
888: && E_{\kappa} = \frac{f_\kappa \overline{\kappa}}{\rsun^2 \Omega_{\rm eq}} = f_\kappa E_\kappa^{\odot} \mbox{   ,   }\nonumber \\
889: && \Pr = \frac{E_\nu}{E_\kappa} = \frac{f_\nu}{f_\kappa} \Pr^{\odot} \mbox{   ,   }\nonumber \\
890: && {\rm H} = \sqrt{\frac{E_\eta E_\nu}{\Lambda}} = \sqrt{f_\nu f_\eta} {\rm H}^\odot \mbox{  where   }  \Lambda  = \frac{B^2}{ 4\pi \rho_0 \rsun^2 \Omega_{\rm eq}^2} \mbox{   ,   }\nonumber \\  
891: && {\rm Bu} = \left(\frac{ND_\rho}{\rsun \Omega_{\rm eq}}\right)^2 = {\rm Bu}^{\odot} \mbox{   ,   }
892: \end{eqnarray} 
893: where $D_\rho$ is the local density scaleheight ($D_\rho = 8.6 \times 10^9$ cm in the tachocline region) and $N$ is the buoyancy frequency ($N = 8 \times 10^{-4}$ s$^{-1}$ in the tachocline region). All of the numbers defined above 
894: actually vary with position within 
895: the solar interior. The first three are Ekman numbers, 
896: defined as the ratios of the rotation timescale to the respective 
897: diffusion timescales. 
898: The solar Ekman numbers calculated using the microscopic diffusivities 
899: at the base of the convection zone are $E_\nu^\odot \simeq 2 \times 10^{-15}$, 
900: $E_\eta^\odot = 3 \times 10^{-14}$, and
901: $E_\kappa^\odot = 10^{-9}$ respectively, all well below unity. The fourth is the 
902: Prandtl number, equal to $2\times 10^{-6}$ near the base of the convection zone in the Sun. 
903: The Hartmann number H depends naturally on the magnetic field strength $B$ selected. 
904: For $B = 1$G, ${\rm H}^\odot = 5 \times 10^{-9}$ at the base of the convection zone, 
905: and H scales as the inverse of the magnetic field strength. The Burger number Bu 
906: is of the order of $10^3$.
907: 
908: \subsubsection{Expected boundary layers and desirable parameter range}
909: \label{subsubsec:paramsreg}
910: 
911: To run simulations in a parameter regime that is at least in the same
912: asymptotic limit as the Sun, one should bear in mind the following 
913: constraints. To be in the non-diffusive regime, one must ensure that all three
914: Ekman numbers in the simulations are well-below unity. The 
915: $f-$parameters selected for the numerical calculations should therefore
916: remain below $(E_\nu^{\odot})^{-1}$, $(E_\eta^{\odot})^{-1}$ and 
917: $(E_\kappa^{\odot})^{-1}$ respectively, and also ensure that the hierarchy
918: $E_\nu \ll E_\eta \ll E_\kappa$ is respected. 
919: 
920: In the non-magnetic case, GB08 showed that the
921: characteristic lengthscales of the dynamics of the radiative zone 
922: (for solar parameter values) can be summarised as $l_1 \sim \rsun$ 
923: (i.e. the entire interior), $l_2 \sim \frac{D_\rho}{\sqrt{\Pr {\rm Bu}}}$ 
924: (i.e. $l_2 \sim 3 \sqrt{\frac{f_\kappa}{f_\nu}} \rsun$) and $l_3 \sim 
925: E_\nu^{1/2} \rsun$. Hence, to retain the same ordering of lengthscales as 
926: in the Sun ($l_2 > l_1 \gg l_3$), the factor $\sqrt{f_\kappa/f_\nu}$ must 
927: not drop below 1/3, placing 
928: very strong constraints on the Prandtl number used in the simulations 
929: (i.e. Pr should not exceed $\sim$10Pr$^{\odot}$).
930: 
931: The typical magnetic boundary layers are reviewed by Acheson \& Hide (1973). 
932: In the case of a magnetic field 
933: parallel to the boundary considered, the nature of the boundary layer depends on 
934: the respective sizes of $E_\nu$ and H: if $E_\nu >$ H then 
935: the boundary layer thickness is of the order of 
936: H$^{1/2} \rsun$, while if $E_\nu  < H$ then the boundary layer thickness scales as 
937: $E_\nu^{1/2} \rsun$ (i.e. the boundary layer is actually of Ekman type).
938: We conclude that for realistic field strengths, the boundary layers 
939: in the horizontal field case (e.g. near the equator) will always be of 
940: Ekman type. When the magnetic field is oblique, the situation changes: 
941: the Ekman regime is recovered when $E_\nu \ll {\rm H}^2$, while the boundary 
942: layer has a magnetic nature when $E_\nu \gg {\rm H}^2$, in which case its thickness
943: is of the order of H$\rsun$. We therefore conclude that unless the magnetic field 
944: strength (in the tachocline) has amplitudes lower than a mG, one
945: may actually expect a Hartmann layer near the radiative--convective interface.
946: 
947: To guarantee that $E_\nu/ {\rm H}^2 \gg 1$ in the simulations we
948: need $f_\eta \ll E_\nu^\odot / ({\rm H}^\odot)^2 \sim 10^2$ if 
949: the magnetic field is indeed 
950: of the order of 1G as suggested by GM98. This is clearly
951: not an achievable goal, so we must inflate the magnetic field strength 
952: artificially to be of the order of 1T near the outer boundary;  in that case 
953: the constraint on $f_\eta$ is much less stringent, with
954: $f_\eta \le 10^{10}$. 
955: 
956: 
957: Finally, if we rely on flows downwelling from the convection zone to confine the 
958: magnetic field, then the amplitude of the flows assumed at the radiative--convective 
959: interface must be large enough for their magnetic Reynolds number ${\rm R}_m$ to be 
960: larger than unity. Since ${\rm R}_m$ is very loosely defined as 
961: \begin{equation}
962: {\rm R}_m = \frac{|u_r| \Delta}{f_\eta \overline{\eta}}\mbox{   ,   }
963: \label{eq:Rm}
964: \end{equation}
965: where $|u_r|$ is the typical amplitude of the radial flows and $\Delta \sim 0.03 \rsun$ 
966: is the observed thickness of the tachocline (both in cgs units), then equation (\ref{eq:Rm}) 
967: implies that we need a priori
968: flows with amplitude of the order of 
969: $f_\eta \overline{\eta} / \Delta \sim 2000 (f_\eta / 10^{10}) $cm/s 
970: to have a magnetic Reynolds number of order one. For values of $f_\eta \sim 10^{10}$ 
971: (typically achieved in the simulations), this naturally requires unphysically large 
972: flow amplitudes. The consequences of this choice are discussed in Sections 
973: \ref{subsec:explor} and \ref{sec:disconcl}.
974: 
975: \subsection{Numerical method}
976: \label{subsec:numalgo}
977: 
978: \subsubsection{Method and tests.}
979: 
980: The numerical method selected for the solution of this elliptic system
981: is based on the expansion of the governing equations and boundary conditions 
982: upon a truncated set of orthogonal functions spanning the $\theta-$direction 
983: (Chebishev polynomials to be precise), followed by the solution of the 
984: associated set of ordinary differential equations (ODEs) in the $r$-direction 
985: using a second-order 
986: Newton-Raphson-Kantorovich (NRK) relaxation 
987: algorithm. The numerical algorithm 
988: is globally similar to the one used by Garaud (2001) and 
989: in Paper I, and is summarised in Appendix C. The high latitudinal resolution 
990: required to capture the detailed 
991: structure of the interior dynamics demands a high-order expansion in $\theta$. 
992: This in turn requires a sophisticated parallelized version of NRK
993: An outline of RELAX -- the parallel NRK algorithm freely available 
994: upon request to P. Garaud -- 
995: is given in Appendix D (see also Garaud \& Garaud, 2007). 
996: 
997: The complexity of the full set of equations (\ref{eq:global2}) 
998: precludes the derivation of analytical solutions in the general case, 
999: and hence direct tests of the full numerical algorithm. However,
1000: we have extensively and successfully tested the code 
1001: through direct comparisons of the numerical solutions with 
1002: analytical solutions of selected {\it subsets} of the system: Paper I 
1003: tested the un-stratified magnetic version of the code, while
1004: GB08 tested the stratified non-magnetic version of the 
1005: code.
1006: 
1007: \subsubsection{Note on the convergence of the solutions}
1008: \label{subsubsec:convsol}
1009: 
1010: In the NRK/RELAX algorithm, a guess is required for the solution of the
1011: ODE system, and refined by successive iterations until the desired accuracy 
1012: is reached. The convergence of the algorithm is guaranteed and immediate 
1013: for linear systems regardless of the initial guess, but requires 
1014: more accurate initial guesses for increasingly dominant 
1015: nonlinearities. As a result, 
1016: solutions are found in practise, first for high-values of the 
1017: diffusivities (i.e. high $f$) and with a low number of latitudinal modes,
1018: and then progressively followed into the nonlinear regime by lowering $f$
1019: and increasing the number of modes.  
1020: 
1021: Convergence of the solutions usually becomes painfully difficult as $f$ 
1022: is lowered below a certain threshold. The causes of the problem can be varied: 
1023: in some cases, the nonlinearities begin to 
1024: dominate the solution and bifurcations occur which reduce the basin of 
1025: attraction of the solution followed; in others, the spatial structure of the 
1026: solution becomes finer and finer, requiring more latitudinal modes and 
1027: more radial meshpoints to be fully resolved.  
1028: 
1029: The only way to address the first problem is to follow the solution 
1030: carefully, and reduce $f$ only by a very small fraction between each run. 
1031: This has the disadvantage of being a frustratingly slow process, with 
1032: little to be gained in practise. As a matter of fact, 
1033: the likely change of stability of the solution tracked (suggested by the 
1034: change in convergence properties) is an interesting dynamical information 
1035: in itself (see Section \ref{subsubsec:algoselec}).  
1036: 
1037: To address the second problem typically requires increasing the number of modes
1038: and meshpoints, which leads to a very rapid increase in computational costs.  
1039: Nonetheless, some acceptable trade-offs are found with typical simulations
1040:  using a few thousand meshpoints (mostly concentrated in the boundary layers) 
1041: and about 60-100 latitudinal modes. These require no more than a few hours 
1042: per iteration on about $2^7$-$2^8$ processors (see Appendix D).
1043: 
1044: \subsubsection{Discussion of the selection of the algorithm}
1045: \label{subsubsec:algoselec}
1046: 
1047: At this point it may be worth discussing an oft-raised question about the 
1048: selection of this specific numerical method of solution, as an 
1049: alternative to the time-dependent ASH code 
1050: used by BZ06. Both methods, when taken individually, 
1051: are equally useful tools for studying the problem, each with their
1052: own advantages and limitations. 
1053: 
1054: For example, the algorithm of BZ06 studies the general properties 
1055: of the radiative zone under the forcing considered, and in addition 
1056: provides information on the stability of the system, 
1057: naturally taking into account 
1058: transport by smaller-scale three-dimensional flows when 
1059: necessary. However, 
1060: the very lengthy integration time required 
1061: implies that only very few different parameter 
1062: values can realistically be explored. Moreover, 
1063: the time-dependent nature of the algorithm makes it more difficult 
1064: to identify quasi-steady states (in particular 
1065: when the typical timescales in the numerical model are far less 
1066: obviously separated than in the Sun).
1067: 
1068: By contrast, the simplified axisymmetric and steady-state nature of the 
1069: equations solved in Paper I and here implies that each solution 
1070: can be found in at most a day of real-time computations (more typically 
1071: an hour). As a result, we are able to explore a 
1072: wide range of parameters and study in detail the behaviour of the solutions
1073: as a function of the imposed forcing (boundary flows, 
1074: differential rotation, magnetic field strength, ...) and as a function 
1075: of the background model (diffusivities, thermo-dynamical profile, ...). 
1076: This feature is invaluable when trying to understand how solutions scale 
1077: with the various parameters -- see GB08 for example.
1078: In addition, we always know that the solutions found
1079: are indeed the steady-state solutions of the system, and not transients. 
1080: On the other hand, taking into account transport
1081: by three-dimensional turbulence is not possible (without artificial
1082: parameterization of its effects). In addition, we cannot 
1083: {\it a priori} know whether the steady-states found are stable, or 
1084: whether they may be unstable to either 2D or 3D perturbations.
1085: 
1086: Interestingly, however, we can make use of the fact that 
1087: the emergence of unstable solutions implies a reduction of the 
1088: basin of attraction of the stable ones (see Section \ref{subsubsec:convsol}), 
1089: and (tentatively) 
1090: identify sudden convergence difficulties 
1091: with the presence of a bifurcation. For example, 
1092: the simulations of BZ06 can be compared with ours for 
1093: \begin{equation}
1094: f_\nu = 2.6\times 10^8 \mbox{  , } f_\eta = 8 \times 10^7 \mbox{   and } f_\kappa = 8 \times 10^5\mbox{   ,   }
1095: \label{eq:bzf}
1096: \end{equation}
1097: (see Section \ref{subsec:bz}). Our numerical algorithm has little difficulty 
1098: converging to a solution 
1099: for parameters $f_\nu$, $f_\eta$ and $f_\kappa$ respectively about
1100: 2 times larger than the ones quoted above. To go from there 
1101: down to the values used by BZ06 on the other hand requires much more care,
1102: a problem which is very likely related to the fact that the 
1103: BZ06 solution is seen to be mildly unstable to 2D and 3D perturbations.
1104: 
1105: % \subsection{Inner core size}
1106: 
1107: %Selecting the size of the artificial inner core requires a tradeoff between 
1108: %the desire to minimize its influence on the modelled 
1109: %radiative zone dynamics, and the need to ease the numerical convergence of the solutions. 
1110: %Indeed, while the NRK algorithm used converges in one iteration from any 
1111: %initial guess for linear problems, nonlinearities reduce the convergence 
1112: %speed and increase the need for a more accurate initial guess. Given the 
1113: %boundary conditions selected, the magnetic field strength in the absence 
1114: %of flows would vary as $r^{-3}$, and thus the relative effect of the 
1115: %nonlinear Lorentz force term in the momentum equation can be expected to vary  
1116: %essentially as $r^{-6}$. Hence, selecting a very small inner
1117: %core implies that the computational domain contains regions of dominant 
1118: %nonlinearities in which convergence is painfully slow when and where 
1119: %the magnetic field strength is high. 
1120: 
1121: %Selecting the size of the core to be $r_{\rm in} = 0.2\rsun$ was found to
1122: %be a good compromise. The convergence of the algorithm is adequate for all
1123: %reasonable parameter regime envisaged, and 
1124: %comparison between numerical solutions with 
1125: %various core sizes but otherwise identical characteristics reveal that the 
1126: %influence of the core size is negligible on the solution in the tachocline
1127: %region. Figure \ref{fig:exp0BCR} compares the outcome of simulations of 
1128: %various core sizes in an extreme case when the selected magnetic field strength
1129: %is very high ($\sim$ 5000 G near the outer boundary). In that case it is found that 
1130: %the influence of the position of the inner boundary is negligible outside 
1131: %of about $0.5\rsun$. When the magnetic field strength is lowered, the influence
1132: %of the position of the inner core weakens even more, which is expected since 
1133: %the Lorentz forces are the only long-range forces present in the system. 
1134: %\begin{figure}
1135: %\includegraphics[width=84mm]{exp0BCR.eps}
1136: %\caption{Variation of the absolute value of the latitudinal velocity 
1137: %(at a latitude of $70^{\circ}$) in three
1138: %simulations with different inner-core sizes ($r_{\rm in} = 0.05\rsun$ 
1139: %for the dotted line, $r_{\rm in } = 0.1\rsun$ for the dashed line and 
1140: %$r_{\rm in } = 0.2\rsun$ for the solid line) and otherwise identical setups. 
1141: %In this simulation, $f=10^{10}$. The seeminly large flow velocities near the 
1142: %outer boundary are consistent with the relatively high Ekman number of this
1143: %set of simulations.}
1144: %\label{fig:exp0BCR}
1145: %\end{figure}
1146: 
1147: 
1148: \section{Numerical experiments}
1149: \label{sec:numexp}
1150: 
1151: The versatility of the numerical algorithm constructed combined with its rapid
1152: real-time execution (see Appendix D) enables us to perform an extensive 
1153: range of simulations, both in terms of the parameter regime studied and in 
1154: terms of the forcing applied. In this section we 
1155: use this tool in a number numerical experiments.
1156: 
1157: In Section \ref{subsec:bz}, we first prove our diagnostic for the 
1158: failure of previous numerical attempts (Paper I, BZ06) to reach 
1159: a dynamical equilibrium qualitatively similar to the one studied by GM98.
1160: We reproduce a quasi-steady state similar to the one achieved 
1161: by BZ06, and study its scaling properties. We show that the 
1162: meridional flow velocities induced in the tachocline, when the
1163: radiative--convective interface is modelled as an impermeable boundary, 
1164: are always too low to confine the field.
1165: 
1166: We contrast this result in Section \ref{subsec:pump} with a similar 
1167: calculation in which meridional flows are directly pumped through the 
1168: boundary (through some unspecified radial forcing mechanism). We show that 
1169: their amplitude remains sufficiently high in the tachocline to provide 
1170: significant field confinement.
1171: 
1172: We then perform a suit of simulations in Section
1173: \ref{subsec:explor},  in the ``desirable'' parameter regime specifically 
1174: selected in Section \ref{subsubsec:paramsreg}, 
1175: and discuss the numerically derived solutions both qualitatively and 
1176: quantitatively. We show that the system can, in some situations and for 
1177: low-enough diffusivities, achieve a balance qualitatively similar to that of 
1178: the GM98 model, but that the solutions also reveal a number of previously 
1179: unaccounted-for dynamical subtelties (see Sections \ref{subsec:fieldconf}, 
1180: \ref{subsec:omvsf} and \ref{subsec:ekmanrole}).
1181: 
1182: \subsection{Failure of ``impermeable'' simulations}
1183: \label{subsec:bz}
1184: 
1185: \subsubsection{The calculation}
1186: 
1187: In this first numerical experiment, we compare our quasi-steady solutions 
1188: with those integrated by BZ06. 
1189: Note that since the overall interior magnetic field 
1190: strength decays slightly 
1191: in their time-dependent calculations, while our 
1192: steady-state solutions fix the value of the field amplitude near 
1193: the inner boundary, the two simulations can only be compared qualitatively 
1194: at best. Nonetheless, the comparison is meaningful 
1195: since the system is thought to relax to a quasi-steady 
1196: equilibrium on a timescale comparable with the meridional 
1197: circulation timescale, which is argued (by BZ06) to be 
1198: much shorter than the magnetic diffusion timescale.
1199: 
1200: To achieve the same parameter regime as the one selected 
1201: by BZ06 in terms of the diffusivities, we set 
1202: \begin{equation}
1203: f_\nu = 2.6\times 10^8  \tilde{f} \mbox{  , } f_\eta = 8 \times 10^7  \tilde{f} \mbox{   and } f_\kappa = 8 \times 10^5  \tilde{f}
1204: \label{eq:bzf2}
1205: \end{equation}
1206: and progressively reduce $\tilde{f}$ from about $10^5$ to 1. When $\tilde{f} = 1$, 
1207: our diffusivities are the same in the tachocline region as 
1208: theirs\footnote{Also note that the BZ06 simulations actually have constant 
1209: diffusivities throughout the radiative interior, while ours vary with radius
1210: as in the Sun. Since we only seek a qualitative comparison of the 
1211: two simulations, this difference in the diffusivities near the inner 
1212: boundary is irrelevant.}. Naturally, we select $u_r^{\rm out} = 0$ to mimic
1213: an impermeable boundary as in BZ06. We also modify our bottom boundary 
1214: conditions to be ``stress-free'', again to be consistent with their simulations. 
1215: Finally, the radial component of the magnetic field on the inner boundary 
1216: is set to be same as the initial condition selected by BZ06 (i.e. $B_r = 500$G 
1217: at the poles at $r = r_{\rm in}$).
1218: 
1219: \begin{figure}
1220: \epsfig{file=f2_color.epsf,width=8cm}
1221: %\epsfig{file=f2_bw.eps,width=8cm}
1222: \caption{Quasi-steady state results for a numerical simulation similar to that of 
1223: Brun \& Zahn (see Section \ref{subsec:bz}). In each quadrant, the dotted circle 
1224: marks the position of the base of the convection zone at $r = r_{\rm cz}$, the 
1225: outer solid circle marks the solar surface and the inner solid circle is at 
1226: $r = r_{\rm in}$. {\it Top left:} Streamlines of the meridional flow. Clockwise 
1227: flows are shown as solid lines, and anticlockwise flows are shown as dotted lines. 
1228: {\it Top right:} Toroidal field (in Gauss). {\it Bottom left:} Magnetic field lines. 
1229: {\it Bottom right:} Angular velocity profile. Note how the magnetic field lines 
1230: connect to the bottom of the convection zone and establish a differentially 
1231: rotating Ferraro state throughout the domain.}
1232: \label{fig:bzsimu}
1233: \end{figure}
1234: 
1235: \subsubsection{Qualitative comparison of the results}
1236: 
1237: The results for $\tilde{f} = 1$  (see equation
1238: (\ref{eq:bzf2})) are shown in Figure \ref{fig:bzsimu}.  The overall
1239: structure of the streamlines and of the magnetic fields  lines,
1240: as well as the amplitude of the toroidal field and of the differential
1241: rotation compare qualitatively well with
1242: the simulations of BZ06 (see the $t=4.7$Gyr solution in Figures 4 and
1243: 5 of BZ06). Small discrepancies in  the magnetic field geometry can be
1244: attributed to the fact that their magnetic  field  is diffusing out of
1245: their inner core, while it is maintained in our simulation by a
1246: permanent dipole.
1247: 
1248: This steady-state calculation therefore confirms the main conclusion of the 
1249: study carried out by BZ06, namely that in the case where the radiative--convective 
1250: interface is modelled as an impermeable boundary, the only possible quasi-steady 
1251: solution has an unconfined field structure leading to a 
1252: differentially rotating Ferraro state.
1253: 
1254: 
1255: \subsubsection{The nature and amplitude of the flows}
1256: 
1257: \begin{figure}
1258: \epsfig{file=f3.eps,width=8cm}
1259: \caption{Absolute value of the radial component 
1260: of the meridional flow  at $r = 0.68 \rsun$ as a function of latitude, when 
1261: $\tilde{f} =1$ (plus symbols), $\tilde{f} =  10$ (diamonds), and $\tilde{f} = 100$ 
1262: (triangles), for $\tilde{f}$ defined in equation (\ref{eq:bzf2}). 
1263: The radial position 
1264: was selected to be safely below the Ekman layer for all three calculations. The 
1265: solid part of the curve denotes $u_r > 0$ and the dotted part of the curve denotes 
1266: $u_r < 0$. Note how $|u_r|$ roughly scales with $\tilde{f}$.}
1267: \label{fig:bzur}
1268: \end{figure}
1269: 
1270: We now justify our diagnostic of the failure of ``impermeable simulations'' 
1271: to find a confined-field solution. The induced meridional flows in 
1272: this set of simulations are generated primarily by Ekman pumping on 
1273: the boundary. Indeed, the magnetic field strength selected by BZ06 implies 
1274: a field of the order of a few tens of Gauss at the most near the outer 
1275: boundary, and therefore according to the analysis of 
1276: Section \ref{subsubsec:paramsreg} a boundary layer in the Ekman regime rather 
1277: than in the Hartmann regime. In that case, the flow amplitudes are in fact
1278: well-estimated by the (non-magnetic) work of GB08 and 
1279: scale as $E_\nu / \sqrt{\Pr}$. This is verified in Figure \ref{fig:bzur}, 
1280: which 
1281: shows the radial velocity profile of the flows as a function of latitude, well below 
1282: the Ekman layer (at $r = 0.68\rsun$), 
1283: in three different simulations: for $\tilde{f} = 1$, $\tilde{f} = 10$ and 
1284: $\tilde{f} = 100$. It is clear from the figure that
1285: the meridional flow velocity decreases linearly with $\tilde{f}$, 
1286: and can in fact be shown to scale as predicted above. 
1287: 
1288: As a result, one can straightforwardly see that the magnetic Reynolds number 
1289: does not increase with decreasing $E_\eta$ when the respective diffusivity 
1290: ratios are held constant: with R$_m$ defined as in equation (\ref{eq:Rm}),  
1291: if $|u_r|$ is proportional to $\sqrt{f_\kappa f_\nu}$ then 
1292: the magnetic Reynolds number is in fact proportional to 
1293: $\sqrt{f_\nu f_\kappa}/f_\eta$ and remains roughly constant as 
1294: $\tilde{f}$ decreases. The effective magnetic Reynolds number 
1295: of the BZ06 simulations can be estimated from the flow amplitudes calculated 
1296: in  Figure \ref{fig:bzur} to be ${\rm R}_m \sim 5 \times 10^{-3}$ -- 
1297: far too low to influence the magnetic field.
1298: 
1299: Using the true solar diffusivities would not change the situation 
1300: since in that case it can be shown that the expected magnetic Reynolds number 
1301: is of the order of ${\rm R}_m \sim 4 \times 10^{-2}$. 
1302: We have therefore confirmed 
1303: our original diagnostic, namely that the reason for the failure of 
1304: numerical simulations to find confined-field solutions is not related to 
1305: the fact that 
1306: the diffusivities in the simulations are too high (i.e. selecting solar 
1307: diffusivities would yield a similar result), but rather to
1308: a fundamental mis-representation of the radiative--convective interface. 
1309: 
1310: \subsection{The possibility of magnetic field confinement by the thermo-viscous mode}
1311: \label{subsec:pump}
1312: 
1313: Following the original idea proposed by GM98, we now explore 
1314: the possibility of confining the interior field through flows which 
1315: are directly downwelling from the convection zone. As shown by GB08, 
1316: flows which are pumped into and out of the radiative 
1317: zone (by stresses in the convection zone for example) can retain significant 
1318: amplitudes throughout much of the tachocline (and below), and could therefore
1319: result in a magnetic Reynolds number larger than unity for low enough 
1320: diffusivities. We illustrate this idea with a qualitative example, which,  
1321: for ease of comparison with the results of the previous section,
1322: is selected to have the same diffusivity ratios and magnetic field 
1323: strength as those of BZ06. 
1324: 
1325: \subsubsection{Structure of the imposed meridional flow}
1326: \label{subsubsec:urforcing}
1327: 
1328: We now select a radial velocity profile $u^{\rm out}_r(\theta)$ which
1329: satisfies the following properties: (i) $u^{\rm out}_r(0) = 0$  (the
1330: radial velocity is  zero on the polar axis), and the flow is symmetric with
1331: respect to the equator (ii) the total mass flux through the boundary (at $r =
1332: r_{\rm cz}$) is zero:
1333: \begin{equation}
1334: \int_{0}^{\pi/2} u^{\rm out}_r(\theta) \sin\theta {\rm d}\theta = 0\mbox{   ,   }
1335: \end{equation}
1336:  and (iii) the total angular momentum flux carried by the meridional circulation  
1337: through the boundary is zero:
1338: \begin{equation}
1339: \int_0^{\pi/2} r^2 \sin^2\theta \tilde{\Omega}_{\rm cz}(\theta) u^{\rm out}_r(\theta) {\rm d} \theta  = 0\mbox{   .   }
1340: \end{equation}
1341: The simplest solution which
1342:  satisfies all three constraints is a 6-th order polynomial in  $\mu = \cos(\theta)$:
1343: \begin{equation}
1344: u^{\rm out}_r(\mu) = U_0 \left( 1 + \sum_{n=1}^3 \alpha_n \mu^{2n}\right) 
1345: \label{eq:urout}
1346: \end{equation}
1347:  where 
1348: \begin{eqnarray}
1349: \alpha_1 = - \frac{3 ( 715 + 169 a_2 + 141 a_4)}{ 143 + 13 a_2 + 25 a_4} \mbox{   ,   }\nonumber \\
1350: \alpha_2 = \frac{5 ( 1001 + 299 a_2 + 207 a_4)}{ 143 + 13 a_2 + 25 a_4}\mbox{   ,   } \nonumber \\
1351: \alpha_3 = - \frac{91 ( 33 + 11 a_2 + 7 a_4)}{ 143 + 13 a_2 + 25 a_4} \mbox{   ,   }
1352: \end{eqnarray}
1353: and $U_0$ is the radial velocity at the equator.
1354: 
1355: \begin{figure}
1356: \epsfig{file=f4.eps,width=7cm}
1357: \caption{Example of imposed radial flow $u^{\rm out}_r$ as a function of 
1358: latitude, for the values of $a_2$ and $a_4$ described in equation 
1359: (\ref{eq:ocz}), satisfying the conditions listed in the main text. 
1360: The velocity of the flow ($U_0$) 
1361: here is arbitrarily selected to be $-1$ cm/s at the equator. 
1362: The width of the upwelling region, as can be seen in the 
1363: figure, spans about 30$^\circ$ in latitude and is centred on 30$^\circ$.}
1364: \label{fig:uforcing}
1365: \end{figure}
1366: 
1367: An example of this profile can be seen in Figure \ref{fig:uforcing}: for the 
1368: values of $a_2$ and $a_4$ given in (\ref{eq:ocz}) the upwelling region is 
1369: centered at $30^{\circ}$ latitude, and ranges from $15^{\circ}$ to $45^{\circ}$. 
1370: Note that the selection of the ``simplest'' profile is fairly arbitrary -- had 
1371: we chosen a different set of orthogonal functions, we would have obtained a 
1372: slightly different flow profile. One may also wish instead to select a profile 
1373: where the width and position of the upwelling region is prescribed, or instead 
1374: of selecting the polar velocity to be zero, to select the equatorial velocity 
1375: to be zero, etc... The only 
1376: necessary constraints, however, are zero mass and angular momentum fluxes.
1377:  These conditions are often ignored, but are crucial to obtaining
1378: meaningful results; failure to satisfy them yields unphysical 
1379: angular velocity profiles (see Garaud, 2007).
1380: 
1381: \subsubsection{Qualitative structure of the solutions}
1382: 
1383: In the simulation presented here, the diffusivities are selected as in 
1384: equation (\ref{eq:bzf2}) with $\tilde{f} = 1$. 
1385: A radial meridional flow profile is imposed at the outer boundary (see 
1386: equation (\ref{eq:urout})), with an amplitude $U_0$ which is progressively 
1387: increased from 0 to about -5.5 cm/s. 
1388: The resulting numerical solution is presented in Figure 
1389: \ref{fig:bzwithu}. It clearly shows that the field lines are 
1390: strongly distorted by the meridional flows, and seem to be confined both 
1391: near the equator, and more importantly also near the poles. In  fact, our results
1392: appear to be at least qualitatively similar to the analytical 
1393: solutions of Wood \& McIntyre (2007), which focussed on the polar regions only.
1394: The angular velocity profile is very different from the one observed in the 
1395: simulations shown in Figure \ref{fig:bzsimu} and reveals uniform rotation 
1396: from the equator up to about 60$^{\circ}$ in latitude. The rotation rate of 
1397: the inner core is found to be 0.86$\Omega_{\rm eq}$. 
1398: \begin{figure}
1399: \epsfig{file=f5_color.epsf,width=8cm}
1400: %\epsfig{file=f5_bw.eps,width=8cm}
1401: \caption{Same as \ref{fig:bzsimu}, but with an imposed meridional flow at 
1402: $r = r_{\rm out}$ (see Figure \ref{fig:uforcing}).}
1403: \label{fig:bzwithu}
1404: \end{figure}
1405: 
1406: We have therefore shown that, at least on a qualitative basis, magnetic field
1407: confinement is possible provided a radial flow is imposed near the radiative--convective interface. 
1408: 
1409: \subsection{Exploration of parameter space}
1410: \label{subsec:explor}
1411: 
1412: Having proved on a qualitative basis that the GM98 model 
1413: may indeed lead to global field confinement, we
1414: now turn to a more quantitative analysis of the resulting tachocline
1415: dynamics. For this purpose, we return to using the boundary conditions
1416: discussed in Section \ref{subsec:numbcs}.
1417: 
1418: In what will be referred to from here on as the ``fiducial model'',  
1419: the enhancements factors
1420: $f_\nu$, $f_\eta$ and $f_\kappa$ are now selected as follows to be in the 
1421: ``desirable'' parameter range discussed in Section \ref{subsubsec:paramsreg}: 
1422: \begin{equation}
1423: f_\nu = \frac{\tilde{f}}{10} \mbox{   ,   } f_\eta =  \tilde{f} \mbox{
1424: ,  } f_\kappa= \frac{\tilde{f}}{100}\mbox{   ,   }
1425: \end{equation}
1426: and the factor $\tilde{f}$ is progressively reduced from $10^{15}$ down
1427: to $10^{10}$ typically ($8\times 10^9$ for the lowest case). For
1428: $\tilde{f} =  10^{10}$, and in the tachocline region, $E_\nu  \sim  2
1429: \times 10^{-6}$, $E_\eta \sim 3 \times 10^{-4}$ and $E_k \sim 10^{-1}$. In
1430: that case, and for all simulations, $\Pr = 10 \Pr^{\odot}$ (guaranteeing
1431: that the thermo-viscous mode lengthscale is indeed of the order of
1432: the entire radiative zone), and $\Pr_m = 0.1 \Pr_m^{\odot}$. Note that the 
1433: simulations of BZ06 use a Prandtl number of the order of 300$\Pr^{\odot}$, 
1434: which does not satisfy our criterion for the hierarchy of the expected 
1435: characteristic lengthscales. 
1436: 
1437: The magnetic field strength on the inner boundary at the pole is selected
1438: to be $B_0 = 7$T, so that the field strength in the tachocline would be 
1439: of the order of 1T in the absence of meridional flows. The radial
1440: flow velocity at the outer boundary is the one discussed in Section 
1441: \ref{subsubsec:urforcing} with $U_0 = -550$cm/s, and the latitudinal 
1442: flow velocity is everywhere zero.
1443: 
1444: \begin{figure}
1445: \epsfig{file=f6_color.epsf,width=8cm,height=18cm}
1446: %\epsfig{file=f6_bw.eps,width=8cm,height=18cm}
1447: \caption{Evolution of the poloidal field topology and of the angular 
1448: velocity profile in the fiducial model for, from top to bottom, 
1449: $\tilde{f} = 10^{13}$, $10^{12}$, $10^{11}$ and $10^{10}$. Note 
1450: that regions with $\Omega < 0.6 \Omega_{\rm eq}$ are drawn in black.}
1451: \label{fig:fvaryombp}
1452: \end{figure}
1453: \begin{figure}
1454: \epsfig{file=f7_color.epsf,width=8cm,height=16cm}
1455: %\epsfig{file=f7_bw.eps,width=8cm,height=16cm}
1456: \caption{Evolution of the topology of the streamlines and of the 
1457: temperature perturbations in the fiducial model for, from top to 
1458: bottom, $\tilde{f} = 10^{13}$, $10^{12}$, $10^{11}$ and $10^{10}$. 
1459: Note that the streamlines shown are representative streamlines, 
1460: the contours selected are different in each plot. Also note that 
1461: the temperature scale is different in each plot.}
1462: \label{fig:fvarypsit}
1463: \end{figure}
1464: 
1465: \subsubsection{From the diffusive regime toward the asymptotic regime}
1466: 
1467: The qualitative evolution of the poloidal field topology and of 
1468: the angular velocity profile as $\tilde{f}$ is reduced 
1469: can be seen in Figure \ref{fig:fvaryombp}, while the corresponding figure 
1470: for the streamlines and the temperature profile is Figure \ref{fig:fvarypsit}. 
1471: In all simulations, the viscous Ekman number is much smaller than one. 
1472: The magnetic field begins to affect the angular velocity profile as
1473: $\tilde{f}$ decreases from $10^{13}$ to $10^{12}$, notably in the deeper 
1474: interior\footnote{Recall that the magnetic diffusivity is smaller 
1475: in the deeper interior, see Appendix A.}. This transition occurs when 
1476: the magnetic Ekman number drops below one. The structure of the 
1477: temperature perturbation profile changes as $\tilde{f}$ is decreased 
1478: from $10^{12}$ down to $10^{11}$, which is attributed this time to 
1479: the ``thermal'' Ekman number approaching one. Finally, there is a clear 
1480: transition  between the ``unconfined field'' configuration 
1481: for $\tilde{f}=10^{11}$ and the ``confined field'' configuration for 
1482: $\tilde{f} = 10^{10}$, which corresponds to the magnetic Reynolds number 
1483: increasing to values above unity. 
1484: 
1485: Contrary to the impermeable-boundary case studied by BZ06, the radial 
1486: flow velocities do not decrease with the diffusivities, but are actually 
1487: seen to increase as $\tilde{f}$ is reduced. This is illustrated in Figure 
1488: \ref{fig:urfid}, which shows the radial flow velocity near the base of 
1489: the presumed tachocline (at $r=0.68 \rsun$, as in Figure \ref{fig:bzur}), 
1490: as a function of latitude, for three values of $\tilde{f}$. Provided 
1491: $E_\kappa < 1$, it appears that $\tilde{u}_r \propto 1/\tilde{f}$ 
1492: although this scaling cannot, for numerical reasons, be explored 
1493: over many orders of magnitude. As a result, the nonlinear interaction 
1494: of the flow and the field rapidly becomes stronger and the field appears 
1495: to be more and more confined, as can be seen in the last panel of Figure 
1496: \ref{fig:fvaryombp}. 
1497: 
1498: \begin{figure}
1499: \epsfig{file=f8.eps,width=8cm}
1500: \caption{Radial flow velocities at $r=0.68\rsun$ as a function of 
1501: latitude in the fiducial model, for three values of $\tilde{f}$: 
1502: $\tilde{f} = 10^{10}$ (plus symbol), $\tilde{f} = 3\times 10^{10}$ 
1503: (diamonds) and $\tilde{f} = 10^{11}$ (triangles). As in Figure 
1504: \ref{fig:bzur}, the solid lines denote upward flow ($\tilde{u}_r > 0$) 
1505: and the dotted lines denote downward flow ($ \tilde{u}_r < 0$). The 
1506: figure therefore shows a downwelling in mid-latitude, with upwelling 
1507: at the poles and the equator. Note how the amplitude of the radial 
1508: velocity appears to increase with decreasing $\tilde{f}$.}
1509: \label{fig:urfid}
1510: \end{figure}
1511: 
1512: \subsubsection{The ``lowest diffusivities'' simulation}
1513: 
1514: We now decrease $\tilde{f}$ as far as possible, until convergence becomes
1515: too difficult. The lowest value achieved in the fiducial model is 
1516: $\tilde{f} = 8\times 10^9$ and the 
1517: results are presented in Figure \ref{fig:bestsim}. The 
1518: number of radial meshpoints used is 3000, and the number of latitudinal modes 
1519: is 80 (note that since only even modes are selected to guarantee equatorial 
1520: symmetry, solving for 80 modes means that the highest Fourier mode considered 
1521: actually is $\cos(158 \theta)$, see Appendix C). 
1522: 
1523: \begin{figure}
1524: \epsfig{file=f9_color.epsf,width=8cm}
1525: %\epsfig{file=f9_bw.eps,width=8cm}
1526: \caption{Results of the simulations of the fiducial model for 
1527: $\tilde{f} = 8\times 10^9$. For each quantity plotted (from top 
1528: left to bottom right, respectively, representative streamlines, 
1529: temperature perturbations, field lines and rotation rate), we show 
1530: both a full quadrant as well as a zoomed-in strip of outer boundary 
1531: layer (the region between $r = 0.65 \rsun$ and $r=r_{\rm out}$). As 
1532: usual, the solid streamlines denote clock-wise flows, and dotted 
1533: streamlines denote counter-clockwise flows. Note that regions with 
1534: $\Omega < 0.6 \Omega_{\rm eq}$ are drawn in black.}
1535: \label{fig:bestsim}
1536: \end{figure}
1537: 
1538: The top-right quadrant of Figure \ref{fig:bestsim} 
1539: shows representative streamlines. The Ekman layer near the outer boundary, 
1540: of width $\sim 0.001 \rsun$, is just visible in the top-right strip. 
1541: The Ekman layer flow does indeed downwell at the poles and equator, 
1542: and upwells in mid-latitudes. 
1543: 
1544: More clearly visible is the ``thermo-viscous'' mode below the 
1545: Ekman layer, which 
1546: (rather unexpectedly, to be honest) has the opposite 
1547: structure: downwelling in mid-latitudes and upwelling near the poles and 
1548: equator. This mode appears to be confined roughly 
1549: within $r \in [0.67,0.7] \rsun$, an
1550: effect which can only be attributed to the Lorentz forces arising from 
1551: the magnetic field underneath. Indeed, the solutions of GB08
1552: in the absence of magnetic fields in the same parameter regime do not
1553: reveal any structure on this typical lengthscale. The associated 
1554: effect of the flow in confining the magnetic field is also obvious 
1555: in the bottom-left quadrant: the field lines are most 
1556: strongly distorted away from the strictly dipolar structure in the same 
1557: region ($r \in [0.67,0.7] \rsun$). Field confinement is discussed in more 
1558: detail in Section \ref{subsec:fieldconf}. 
1559: 
1560: The solution for the angular velocity profile is again more complex than 
1561: expected: the 
1562: combination of the two modes of propagation of the meridional flows creates
1563: radial structures both on the Ekman scale and on a larger scale 
1564: ($r \sim 0.1 \rsun$), as well as latitudinal structures which are clearly 
1565: seen to be correlated with the magnetic field lines. However, contrary to the 
1566: solution shown in Figure \ref{fig:bzsimu}, there is no clear evidence for a 
1567: {\it large-scale} latitudinal or radial gradient in $\Omega$ below 
1568: about 0.6$\rsun$, including in the polar regions. This is the first set of 
1569: self-consistent simulations to reveal this kind of solution. Since the 
1570: angular velocity profile inferred from helioseismic inversions can be thought
1571: of as a weighted spatial average of the true angular velocity profile 
1572: in the Sun, and since the spatial extent of the averaging kernels is 
1573: much larger than the features seen in the simulations, the angular 
1574: velocity profile shown in Figure \ref{fig:bestsim} would 
1575: be observationally interpreted to be uniform below $0.6\rsun$,
1576: whereas the one shown in Figure 
1577: \ref{fig:bzsimu} would not. The actual value of the rotation rate
1578: calculated in the simulations is discussed in Section \ref{subsec:omvsf}.
1579: 
1580: A closer look at the angular velocity profile in the region 
1581: $r \in [0.65,0.7] \rsun$ (see the bottom-right strip)
1582: reveals a change in the dominant angular momentum transport processes between
1583: the uppermost layers ($r > 0.67\rsun$) and the lower layers ($r < 0.67 \rsun$).
1584: When $r > 0.67\rsun$ the angular velocity contours are more-or-less
1585: aligned with the streamlines, whereas for $r < 0.67 \rsun$ the angular velocity
1586: contours are aligned with the magnetic field lines. This is at least
1587: qualitatively consistent with the idea proposed by GM98 of a multi-layered
1588: tachocline, in which the dynamics of the uppermost layer are controlled by 
1589: the tachocline flows, while the dynamics of the bulk of the radiative zone 
1590: are controlled
1591: by the primordial field. The interface between the two regions, which could
1592: be identified with the ``magnetic diffusion layer'' of the GM98 model, clearly 
1593: has quite a complex topology.
1594: 
1595: One can also observe a number of very strong localised 
1596: ``jets'' of more rapidly or more slowly rotating fluid. The jets appear to 
1597: be stronger for lower diffusivities (one can easily compare the  
1598: results of Figure \ref{fig:bestsim} with the last panel of Figure 
1599: \ref{fig:fvaryombp}). In fact, we tentatively attribute the numerical scheme's 
1600: failure to find solution for $\tilde{f} < 8 \times 10^9$ to an 
1601: intrinsic instability of the jets. 
1602: But whether such strong features would exist in the Sun is debatable: both 
1603: the imposed flow velocities and the magnetic field strength have been 
1604: artificially increased by many orders of 
1605: magnitude to ensure a magnetic Reynolds number greater than one in the 
1606: simulations, and a Hartmann number closer to that of the Sun (see Section 
1607: \ref{subsubsec:paramsreg}). As a result, the amplitude of these jets is also 
1608: artificially increased in the simulations compared with what one may expect 
1609: in the Sun. Unfortunately, given the nonlinear nature of the problem, 
1610: it is difficult 
1611: to predict what the actual strength of these jets would be should 
1612: diffusivities, magnetic field and imposed flows velocities be 
1613: simultaneously reduced to the expected solar values. 
1614: 
1615: \subsection{Magnetic field confinement}
1616: \label{subsec:fieldconf}
1617: 
1618: To study the nature of field confinement more quantitatively, we
1619: consider the ratio $\xi = |B_r / B_{r,{\rm dipolar}}|$ where
1620: $B_{r,{\rm dipolar}}$ is the hypothetical radial component of the
1621: solution for the magnetic field in the purely diffusive case:
1622: \begin{equation}
1623: B_{r,{\rm dipolar}}(r,\theta) =  B_0 \cos{\theta}
1624: \left(\frac{r}{r_{\rm in}}\right)^{-3} \mbox{  . }
1625: \end{equation} 
1626: Typically, we expect that $\xi < 1$ when magnetic field lines are
1627: either completely expelled from a region, or bent to the
1628: horizontal. On the other hand, we expect that $\xi > 1$ in regions
1629: where the magnetic field lines are pushed together by converging
1630: (latitudinal) flows. It is therefore a good diagnostic  of the
1631: processes we are interested in.
1632: 
1633: The quantity $\xi$ is shown in Figure \ref{fig:fielddiff} for various
1634: simulations. The top two quadrants are derived from the numerical
1635: solutions for $\tilde{f} = 10^{11}$ and $\tilde{f} = 8\times 10^9$
1636: respectively, and reveal a strong change in the behaviour of the
1637: solutions when ${\rm R}_m < 1$ and ${\rm R}_m > 1$.
1638: 
1639: In the case where $\tilde{f} = 10^{11}$, we observe a good match
1640: between the spatial variation of $\xi$ and the imposed forcing: $\xi <
1641: 1$ in downwelling regions and $\xi > 1$ in the upwelling regions (at
1642: mid-latitudes). However, the amplitude of $\xi$ remains close to one,
1643: revealing only a very weak effect of the flow on the field. This is
1644: not entirely surprising since the magnetic Reynolds number for
1645: $\tilde{f} = 10^{11}$ can be deduced from the flow amplitude shown in
1646: Figure \ref{fig:urfid} to be ${\rm R}_m \sim 0.03$ at
1647: $r=0.68\rsun$. In fact, in this very weakly nonlinear regime there is
1648: some evidence for a correlation (see Figure \ref{fig:fielddiff},
1649: bottom left quadrant) between $1-\xi$ and $\tilde{f}$: we find that
1650: $|1-\xi| \sim 10^{10} \tilde{f}^{-1}$.
1651: 
1652: However, Figure \ref{fig:fielddiff} also reveals that matters become
1653: more complex as ${\rm R}_m$ increases above unity. For $\tilde{f} =
1654: 8\times 10^9$ the magnetic Reynolds number is now of the order of
1655: ${\rm R}_m \simeq 1.5$, and the connection between the field and the
1656: meridional flow is correspondingly fully nonlinear. The correlation
1657: between $1-\xi$ and $\tilde{f}$ breaks down, as seen in the bottom
1658: right quadrant. This incidentally proves that the kind of boundary
1659: conditions advocated by Kitchatinov \& R\"udiger (2006) to mimic
1660: ``field confinement by meridional flows'' cannot accurately describe
1661: the nonlinear dynamics of the system. We also observe that the regions
1662: of enhancement and reduction of the radial field are no longer
1663: perfectly correlated with the {\it imposed} upwelling/downwelling
1664: regions. In fact, some enhancement in the radial field strength can be
1665: seen near the equator as well as near the polar regions. This can be
1666: attributed to the convergence points of the meridional flows {\it
1667: below} the Ekman layer, which, as seen in Figure \ref{fig:bestsim}
1668: have the opposite sign as that of the imposed flow.
1669: 
1670: All in all, although there is definite evidence for magnetic field
1671: confinement along the lines of the model proposed by GM98, it is
1672: probably fair to say  that the system behaves in a far more complex
1673: way than anticipated,  and that the simulations in the fiducial model
1674: are just beginning to  scratch the surface of the ``interesting''
1675: regime.
1676: 
1677: \begin{figure}
1678: \epsfig{file=f10_color.epsf,width=8cm}
1679: %\epsfig{file=f10_bw.eps,width=8cm}
1680: \caption{Measure of magnetic confinement, using the quantity $\xi$ as
1681: defined in the main text, Section \ref{subsec:fieldconf}. {\it Top
1682: left}: Variation of $\xi$ with radius and latitude, in a fiducial
1683: model for $\tilde{f} = 10^{11}$. {\it Top right}: Same as on the left
1684: but for $\tilde{f}=8\times 10^9$. Note the change in the scale of the
1685: perturbations between the two plots, and also note that here the
1686: contours are logarithmically spaced. {\it Bottom left}: $|1-\xi|$
1687: evaluated at the outer boundary, at three different latitudes
1688: respectively in the downwelling regions ($10^{\circ}$, plus symbols,
1689: and $60^{\circ}$, diamonds) and in the middle of the upwelling region
1690: ($30^{\circ}$, triangles). This log-log plot emphasises the power-law
1691: relationship between $|1-\xi|$ and $\tilde{f}$ for ${\rm R}_m < 1$,
1692: see main text. {\it Bottom right:} Similar to the left-side plot but
1693: showing $\xi$ with a linear scale, emphasising the breakdown of this
1694: relationship when ${\rm R}_m > 1$.}
1695: \label{fig:fielddiff}
1696: \end{figure}
1697: 
1698: \subsection{Interior angular velocity}
1699: \label{subsec:omvsf}
1700: 
1701: As in Paper I, we consider the angular velocity of the inner core
1702: $\Omega_{\rm in} = \overline{\Omega} + \tilde{\Omega}_{\rm in}$ (which
1703: is an  eigenvalue of the calculation performed) as a diagnostic of the
1704: most important angular momentum transport processes, and of the
1705: ``accuracy'' of the model  in reproducing the observations. This
1706: quantity is shown in Figure  \ref{fig:omvsf} as a function of
1707: $\tilde{f}$ for the fiducial model, and for simulations with faster 
1708: and slower imposed flows.
1709: 
1710: \begin{figure}
1711: \epsfig{file=f11.eps,width=8cm}
1712: \caption{Angular velocity of the inner core as a function of
1713: $\tilde{f}$, in the fiducial model (star symbols), and in a model
1714: where the amplitude of the imposed radial flow is ten times larger
1715: (triangular symbols) and ten times smaller (diamond symbols)
1716: respectively.  }
1717: \label{fig:omvsf}
1718: \end{figure} 
1719: 
1720: In very diffusive simulations, the predicted angular velocity
1721: $\Omega_{\rm in} \simeq 0.957 \Omega_{\rm eq}$  is consistent with
1722: purely viscous angular  momentum transport throughout the entire
1723: interior (see Gough, 1985).  As $\tilde{f}$ is progressively reduced,
1724: the system undergoes a rather  impressive bifurcation at $\tilde{f}
1725: \simeq 2.5 \times 10^{12}$, which  corresponds to the change of nature
1726: of the boundary layer near the inner  core from an Ekman layer to an
1727: Ekman-Hartmann layer. When this happens,  a new flow cell of opposite
1728: vorticity appears right against the inner core,  so that angular
1729: momentum transport changes direction. This leads to a rather dramatic
1730: jump in the inner core rotation rate, which is unlikely to be of any
1731: relevance to the Sun, but deserved an explanation.  Note that the
1732: critical value of $\tilde{f}$ for which the bifurcation occurs  is
1733: proportional to the imposed magnetic field strength $B_0$.
1734: 
1735: In the case of the fiducial model, the angular velocity of the core then 
1736: remains roughly constant ($\Omega_{\rm in} \simeq (0.875 \pm 0.005) 
1737: \Omega_{\rm eq}$) as 
1738: $\tilde{f}$ is further reduced by two orders of magnitude. This range 
1739: corresponds to the regime where ${\rm R}_m < 1$, where the magnetic 
1740: field lines 
1741: essentially remain dipolar. In fact, the same inner core velocity is also 
1742: found for lower imposed flow velocities (see Figure \ref{fig:omvsf}), as well 
1743: as for higher and lower field 
1744: strength (always with ${\rm R}_m < 1$). This suggests that the value  
1745: $ \Omega_{\rm in} \sim 0.875 \Omega_{\rm eq}$ is a rather universal 
1746: characteristic of angular momentum transport by dipolar fields 
1747: in the Hartmann regime, in 
1748: this particular geometry, and subject to this particular imposed angular 
1749: velocity profile at the outer boundary. 
1750: 
1751: However, it is clear from the analysis of Section \ref{subsec:fieldconf} 
1752: that the simulations in the fiducial model have not quite yet 
1753: reached the asymptotic 
1754: regime even for the lowest values of $\tilde{f}$ achieved. As a matter 
1755: of fact, in the last few points of the 
1756: fiducial model curve (in the region where ${\rm R}_m$ begins to exceed one), 
1757: $\Omega_{\rm in}$ is seen to change more rapidly, decreasing slightly 
1758: below 0.87 $\Omega_{\rm eq}$.
1759: 
1760: The role of the meridional flows as angular momentum transporters can 
1761: best be seen in a simulation with much higher imposed flow 
1762: velocities (see the curve with the triangular symbols, for $U_0 = -5500$cm/s). 
1763: In these simulations, the transition to ${\rm R}_m > 1$ 
1764: occurs for values of $\tilde{f}$ typically 10 times larger 
1765: than in the fiducial model. In that case, we observe that the predicted 
1766: core velocity is 
1767: notably different from the other two cases, and changes rather rapidly 
1768: with $\tilde{f}$. It is therefore more than likely that the predicted 
1769: angular velocity of the core in the fiducial model would also begin 
1770: to deviate more noticeably away from $\Omega_{\rm in} = 0.875 \rsun$ were
1771: we able to continue decreasing $\tilde{f}$. Again, we are only beginning to 
1772: approach the asymptotic regime so that the simulated 
1773: angular velocity of the core cannot yet and should not be compared 
1774: directly with that of the Sun. 
1775: 
1776: \section{Discussion and prospects.}
1777: \label{sec:disconcl}
1778: 
1779: \subsection{Summary of the results}
1780: 
1781: Despite the difficulties encountered when attempting to
1782: find numerical solutions for asymptotically low values of the 
1783: diffusivities, the nonlinear dynamics which emerge from our simulations 
1784: are closer to what one may expect from the GM98 model than any other 
1785: simulation performed to date. 
1786: 
1787: More precisely we do observe (see Section \ref{subsec:explor}) 
1788: the quenching of the large-scale differential rotation by the 
1789: primordial magnetic field, even in the polar regions.
1790: We also observe the partial confinement of the field by the meridional flows
1791: to the radiative interior (see Section \ref{subsec:fieldconf}), with 
1792: the reduction of the radial field strength (in some regions)  
1793: by more than 70\%. Finally, we observe the concurrent confinement of the 
1794: meridional flows to the upper layers of the radiative zone 
1795: ($r > 0.67\rsun$) by the magnetic field. We have not yet been 
1796: able to reduce the diffusivities down sufficiently far to observe
1797: a truly segregated structure where the bulk of the tachocline flows
1798: are completely magnetic-free. However, there is evidence in the 
1799: observed rotation profile (see Figure \ref{fig:bestsim}) for a
1800: transition between regions where angular-momentum transport is dominated
1801: by the meridional flows, and regions where it is dominated by the magnetic 
1802: field. The ``magnetic diffusion layer'' studied by GM98, 
1803: which is thought to control this transition,
1804: appears to have a rather complex geometrical structure which prevents a
1805: more detailed study of the GM98 scalings.
1806: 
1807: The calculated value of the interior angular velocity $\Omega_{\rm in}$ 
1808: (see Section \ref{subsec:omvsf}) does not match the observed value. 
1809: However, since even the lowest-diffusivity simulations presented here are 
1810: only just beginning to enter the 
1811: asymptotic parameter regime described in Section \ref{subsubsec:paramsreg} 
1812: we do not view the poor match with the observations as an intrinsic problem
1813: with the GM98 model (yet) but rather as evidence that more work should be 
1814: done to decrease the diffusivities even further -- a challenging task.
1815: 
1816: \subsection{The stability of the solutions?}
1817: 
1818: As described in Section \ref{subsec:explor}, the convergence of the solutions
1819: becomes rather difficult when the values of the diffusivities are decreased 
1820: below a certain threshold (in the fiducial model for 
1821: $\tilde{f} < 8 \times 10^9$). The typical reasons for this convergence failure 
1822: were listed and discussed in Section \ref{subsubsec:convsol}: 
1823: intrinsic linear/nonlinear instabilities or insufficient latitudinal 
1824: resolution. In this particular case, the various scenarios are
1825: equally plausible, and difficult to disentangle without 
1826: a supporting fully nonlinear 3D calculation. The field configuration 
1827: near the inner core could be subject to instabilities (as seen in the 
1828: simulations of BZ06). The Ekman(-Hartmann) layer near the outer boundary 
1829: could also be subject to instabilities, or could be developing latitudinal 
1830: structures which are too fine to resolve (see Figure \ref{fig:el}). 
1831: We have in fact tentatively identified 
1832: the emergence of strong jets in the angular velocity profile 
1833: (see Figure \ref{fig:bestsim} for example) as the most 
1834: likely source of instabilities (and convergence failure), but the reader 
1835: should be cautionned that this statement is only speculative.
1836: 
1837: As mentioned in Section \ref{subsec:explor}, whether the equilibrium governed 
1838: by the GM98 model is stable or unstable cannot directly be inferred from 
1839: the stability of the numerical solutions (since they operate in a different 
1840: parameter regime) but it is clearly a fundamental question. 
1841: Should our diagnostic concerning the stability of the jets 
1842: prove to be correct, then a possible path for 
1843: future investigation (using this relaxation method) may be to continue
1844: searching for solutions, starting where we left off, and slowly 
1845: lowering simultaneously 
1846: the amplitude of the imposed flow (and field) with the diffusivities to
1847: limit the strength of the jets generated. This could be seen as blindly
1848: navigating the stable regions of parameter space while avoiding instability 
1849: reefs. But what could ideally be derived from such an exercise, 
1850: should we be able to acquire sufficient data, are scaling relationships 
1851: between the jet strength, the diffusivities and the imposed flow strength 
1852: which may then be used to infer tentative information on the stability 
1853: of the GM98 solution itself. 
1854: 
1855: \subsection{The role of an Ekman layer?}
1856: \label{subsec:ekmanrole}
1857: 
1858: \begin{figure}
1859: \epsfig{file=f12.eps,width=8cm}
1860: \caption{Angular velocity profile on and just below the 
1861: radiative--convective interface (a) at 0.7$\rsun$ (b) at 
1862: 0.6999$\rsun$, (c) at 0.6995$\rsun$ and (d) at 0.699$\rsun$ 
1863: in the fiducial model for $\tilde{f}=8\times 10^9$. Note how 
1864: the angular velocity profile in (and therefore also below) the 
1865: Ekman layer is very different from the one imposed by the convection zone.}
1866: \label{fig:el}
1867: \end{figure}
1868: 
1869: Our simulations have also revealed a fundamental difference with 
1870: the original GM98 model: the presence and role of an 
1871: Ekman mode\footnote{Since the 
1872: bulk of the GM98 tachocline is presumed to be magnetic-free, 
1873: one should indeed consider an
1874: Ekman mode rather than an Ekman-Hartmann mode}. While it is clear 
1875: from Figure \ref{fig:bestsim} that the Ekman flows themselves play 
1876: no role in confining the field (contrary to the claims made by 
1877: Kitchatinov \& R\"udiger 2006) our simulations reveal that the role 
1878: of the Ekman layer is still far from trivial.
1879: 
1880: Indeed, the differential 
1881: rotation profile imposed at the radiative--convective interface 
1882: is quite different from the differential rotation 
1883: profile transmitted by the Ekman layer. This is illustrated
1884: in Figure \ref{fig:el}, which shows the evolution of the angular 
1885: velocity profile with depth across the Ekman layer. Thus while the GM98 model 
1886: correctly describes
1887: the non-viscous tachocline dynamics {\it below} the Ekman layer, 
1888: the Ekman mode could in fact influence the system by 
1889: modifying the angular velocity profile ``seen'' by the bulk of the 
1890: tachocline\footnote{adding another layer to the sandwich...}. 
1891: 
1892: The extent to which this effect influences the tachocline is unclear. Firstly,
1893: other processes which also transport angular momentum 
1894: (convective overshoot, gravity waves, small-scale and large-scale magnetic 
1895: stresses associated with the dynamo field) are known to take place in the close
1896: vicinity of the radiative--convective interface. These processes were not
1897: included in GM98's analysis and cannot be modelled with the 
1898: present numerical algorithm, but could equally affect the tachocline dynamics.
1899: Secondly, Ekman layers (laminar or turbulent) must {\it 
1900: by definition} be present in any rotating system which exhibits very 
1901: rapid changes in the imposed stresses. However, whether the Ekman 
1902: mode actually plays an important role in the tachocline dynamics depends
1903: on the relative importance of the bulk thermal-wind stresses and the 
1904: combination of all the other rapidly varying turbulent stresses 
1905: (see Figure \ref{fig:bls}). Indeed, if the base of the 
1906: convection zone is already essentially in thermal-wind balance (see 
1907: Miesch, 2005 for instance), then the amplitude of the Ekman mode and its effect
1908: on the angular velocity profile will be small (GB08). On the other hand, 
1909: if this is not the case then significant Ekman flows are expected 
1910: (as in the simulations shown in the present work for example), with the aforementionned consequences.
1911: 
1912: \begin{figure}
1913: \epsfig{file=f13.eps,width=8cm}
1914: \caption{\small A pictorial representation of the expected tachocline 
1915: flows in two extreme cases. In both pictures, flows are downwelling 
1916: from the convection zone into the radiative zone. As in the GM08 model, 
1917: the bulk of the tachocline is ventilated by the thermo-viscous mode, 
1918: which interacts with the magnetic field in a thin advection-diffusion 
1919: layer. On the left, and as in the simulations presented in this paper, 
1920: the convection zone is not dominated by thermal-wind balance, and a 
1921: significant portion of the flows downwelling from the convection zone 
1922: rapidly return within a thin Ekman layer. The angular velocity profile 
1923: seen by the tachocline differs from the one observed in the convection 
1924: zone. On the right, a hypothetical situation where the convection zone 
1925: is essentially in thermal-wind balance. In that case, the Ekman mode 
1926: is negligible and the GM98 model directly applies.}
1927: \label{fig:bls}
1928: \end{figure}
1929: 
1930: \subsection{Prospects}
1931: 
1932: The present work has emphasized the necessity of flows downwelling 
1933: from the solar convection zone as a means to confine the 
1934: internal primordial field and guarantee the uniform rotation of the radiative
1935: interior as originally proposed by GM98. The nature of the thermal 
1936: and dynamical balance 
1937: governing the convection zone itself therefore also controls 
1938: the dynamics of the tachocline through the spatial variation and
1939: amplitude of the flows crossing the radiative--convective interface. 
1940: Meanwhile, steady progress in 
1941: modelling the convection zone has emphasized its dependence on the thermal 
1942: stratification of the tachocline (Rempel, 2005; Miesch, Brun \& Toomre, 2006). 
1943: One can only conclude
1944: that the dynamics of the convective and radiative regions are 
1945: intrinsically and fundamentally coupled through this internal layer 
1946: called the tachocline, and that the only sensible way forward from this 
1947: point on is the construction of whole-Sun models including both regions. 
1948:  
1949: 
1950: 
1951: \section*{Acknowledgements}
1952: 
1953: This work is the outcome of nearly ten years of discussions with many 
1954: colleagues, including Nic Brummell, Sacha Brun, Gary Glatzmaier, Douglas Gough,
1955: Michael McIntyre, Tamara Rogers, Nigel Weiss, Jean-Paul Zahn. 
1956: P. Garaud thanks them all for their patience. Respective parts of this 
1957: work have been 
1958: supported, at various times, by New Hall (Cambridge), PPARC, and more recently
1959: by Calspace (for J.-D. Garaud) and by NSF-AST-0607495 (for P. Garaud). The
1960: numerical simulations were performed on the Pleiades cluster at UCSC, 
1961: purchased using an NSF-MRI grant.
1962: 
1963: 
1964: \begin{thebibliography}{99}
1965: 
1966: \bibitem{ah73}
1967:  Acheson, D. J. \& Hide, R., 1973, Rep. Prog. Phys., 36, 159
1968: %\bibitem{ba01}
1969: % Basu, S \& Antia, H. M., 2001, MNRAS, 324, 498 
1970: \bibitem{bal89}
1971:  Brown, T. M., Christensen-Dalsgaard, J. Dziembowsky, W. A., Goode, P., Gough, D. O. \& Morrow, C. A., 1989, ApJ, 343, 526
1972: %\bibitem{btz99}
1973: % Brun, A. S., Turck-Chi\`eze, S. \& Zahn, J.-P., 1999, ApJ, 525, 1032
1974: \bibitem{brunal04} 
1975:  Brun, A.S., Miesch, M.S. \& Toomre, J., 2004, ApJ, 614, 1073
1976: \bibitem{bz06}
1977:  Brun, A. S. \& Zahn, J.-P., 2006, A\&A, 457, 665
1978: %\bibitem{c01}
1979: % Cally, P. S., 2001, Solar Physics, 199, 231
1980: %\bibitem{cal99}
1981: % Charbonneau, P., Christensen-Dalsgaard, J., Henning, R., Larsen, R. M., Schou, J., Thompson, M. J., \& Tomczyk, S., 1999, ApJ, 527, 445
1982: \bibitem{jcdschou88} 
1983:  Christensen-Dalsgaard, J. \& Schou, J., 1988,  in {\it Seismology of the Sun and Sun-Like Stars}, ed. V. Domingo \& E.J. Rolfe (ESA-SP286), p. 149
1984:  \bibitem{jcdgoughthompson91}
1985:  Christensen-Dalsgaard, J., Gough D.O. \&  Thompson, M.J., 1991, ApJ, 378, 413
1986: \bibitem{cluneal99} 
1987:  Clune T. L. {\it et al.} 1999, Parallel Comp., 25, 361
1988: %\bibitem{dg99}
1989: % Dikpati, M. \& Gilman, P. A., 1999, ApJ, 512, 417
1990: %\bibitem{dcj98}
1991: % Dormy, E., Cardin, P. \& Jault, D., 1998, Earth \& Planetary Sci. Letters, 
1992: %160, 15
1993: %\bibitem{djs01}
1994: % Dormy, E., Jault, D. \& Soward, A. M., 2001, JFM, {\it in press}
1995: \bibitem{dal89}
1996:  Dziembowski, W. A., Goode, P. R. \& Libbrecht, K. G., 1989, ApJ, 337, L53
1997: \bibitem{e25}
1998:  Eddington, A. S., 1925, Observatory, 48,73
1999: \bibitem{eg98}
2000:  Elliott, J. R. \& Gough, D. O., 1999, ApJ, 516, 475
2001: \bibitem{f37}
2002:  Ferraro, V. C. A., 1937, MNRAS, 97, 458
2003: %\bibitem{g99}
2004: % Garaud, P., 1999, MNRAS, 304, 583
2005: \bibitem{g01}
2006:  Garaud, P., 2001, PhD Thesis, available from \\ http://www.ams.ucsc.edu/$\sim$pgaraud/Work.html
2007: \bibitem{g02}
2008:  Garaud, P., 2002, MNRAS, 329, 1
2009: \bibitem{g07}
2010:  Garaud, P., 2007, in {\it The Solar Tachocline}, pp. 147--181, eds. Hughes, D. W., Rosner, R. \& Weiss, CUP.
2011: \bibitem{gg07}
2012:  Garaud, P. \& Garaud, J.-D., 2007. AMS Technical Report ams2007-17, available from \\ http://www.ams.ucsc.edu/research.html
2013: \bibitem{gb08}
2014:  Garaud, P. \& Brummell, N. H., 2008, ApJ, 674, 498
2015: %\bibitem{gf97}
2016: % Gilman, P. A. \& Fox, P. A., 1997, ApJ, 522, 1167
2017: \bibitem{g89}
2018:  Gilman, P.A., Morrow, C.A. \& Deluca, E.E., 1989, 338, 528
2019: \bibitem{glatzmaier84} 
2020:  Glatzmaier, G.A., 1984, J. Comp. Phys., 55, 461
2021: \bibitem{go93}
2022:  Golub, G. H. \& Ortega, J.M., 1993, in {\it Scientific Computing: An Introduction with Parallel Computing}, pp. 302--321
2023: \bibitem{g85}
2024:  Gough, D. O., 1985, in {\it Proceedings of an ESA Workshop, Garmisch-Parkenkirschen, Germany}
2025: %\bibitem{gk95}
2026: % Gough, D. O. \& Kosovichev, A. G., 1995, {\it Proceedings of the 4th SOHO Workshop, Pacific Grove, California, 1995}, ESA SP-376
2027: \bibitem{gmi98}
2028:  Gough, D. O. \& McIntyre, M. E., 1998, Nature, 394, 755
2029: %\bibitem{g00}
2030: % Gough, D. O., 2000, Science, 287, 2434
2031: \bibitem{go07}
2032:  Gough, D. O., 2007. in {\it The Solar Tachocline}, pp. 3--30, eds. Hughes, D. W., Rosner, R. \& Weiss, CUP.
2033: %\bibitem{hal00b}
2034: % Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R. W., Larsen, R. M., Schou, J., Thompson, M. J. \& Toomre, J., 2000, Science, 287, 2456
2035: \bibitem{kr06}
2036: Kitchatinov, L. L. \& R\"udiger, G., 2006, A\&A, 453, 329.
2037: %\bibitem{kal97}
2038: % Kleeorin, N., Rogachevskii, I., Ruzmaikin, A., Soward, A. M. \& Starchenko, S., 1997, JFM, 344, 213 
2039: \bibitem{k88}
2040:  Kosovichev, A. G., 1988, Sov. Astron. Lett., 14, 145
2041: \bibitem{miesch00} 
2042:  Miesch, M.S. {\it et al.}, 2000, ApJ, 532, 596.
2043: \bibitem{miesch05}
2044:  Miesch, M.S., 2005, LRSP, 2, 1
2045: \bibitem{mbt06}
2046:  Miesch, M.S., Brun, A.S. \& Toomre, J., 2006, ApJ, 641, 618
2047: \bibitem{mc99}
2048:  MacGregor, K. B. \& Charbonneau, P., 1999, ApJ, 519, 911
2049: \bibitem{m53}
2050:  Mestel, L., 1953, MNRAS, 113, 716
2051: \bibitem{mw87}
2052:  Mestel, L. \& Weiss, N. O., 1987, MNRAS, 226, 123
2053: %\bibitem{ns69}
2054: % Nakagawa, Y. \& Swarztrauber, P., 1969, ApJ, 155, 295 
2055: \bibitem{pal96} 
2056:  Press, W.J. {\it et al.}, 1996, in {\it Numerical Recipes in Fortran 77}, second edition, pp. 753--763.
2057: %\bibitem{p56}
2058: % Proudman, I., 1956, JFM, 1, 505
2059: %\bibitem{p16}
2060: % Proudman, J., 1916, Proc. Roy. Soc. A, 92, 408
2061: \bibitem{r05}
2062:  Rempel, M., 2005, ApJ, 622, 1320
2063: \bibitem{rk97}
2064:  R\"{u}diger, G. \& Kitchatinov, L. L., 1997, Astr. Nachr. 318, 273
2065: \bibitem{sal98}
2066:  Schou, J., et al., 1998, ApJ, 505, 390
2067: \bibitem{sz92}
2068:  Spiegel, E. A. \& Zahn, J.-P., 1992, A\&A, 265, 106
2069: %\bibitem{s66}
2070: % Stewartson, K., 1966, JFM, 26, 131
2071: \bibitem{sal05}
2072: Sule, A., R\"udiger, G., \& Arlt, R., 2005, A\&A, 437, 1061
2073: \bibitem{s50}
2074:  Sweet, E., 1950, MNRAS, 110, 548
2075: %\bibitem{t21}
2076: % Taylor, G. I., 1921, Proc. Roy. Soc. A, 100, 114
2077: \bibitem{w92}
2078:  Wright, S. J., 1992, SIAM J. Sci. Statist. Comput., 13, 742
2079: \bibitem{z07}
2080:  Zahn, J.-P., 2007, in {\it The Solar Tachocline}, eds. Hughes, D. W., Rosner, R. \& Weiss, CUP.
2081: \end{thebibliography}
2082: 
2083: \appendix
2084: 
2085: \section[]{Selection of the background state}
2086: 
2087: %% This appendix could be in more or less final form.
2088: 
2089: The spherically symmetric background solar model selected for this
2090: paper is Model S of Christensen-Dalsgaard {\it et al.} (1991). Model S  
2091: provides calculated solar data for all of the relevant thermodynamical and
2092: compositional quantities on a fixed mesh. In particular, we use the provided
2093: fields $\overline{T}$, $\overline{p}$, $\overline{\rho}$, $\overline{g}$, 
2094: $\overline{N}^2$ (where $N$ is the buoyancy frequency), $\overline{c}_{\rm p}$ 
2095: and $\overline{\kappa}_{\rm R}$ (where $c_{\rm p}$ is the specific heat at 
2096: constant pressure, and $\kappa_{\rm R}$ is the Rosseland mean opacity). 
2097: In order to use this data for our purpose, we need to interpolate it
2098: upon our own selected numerical mesh, using a standard rational
2099: interpolation routine (cf. Numerical Recipes). The task is delicate
2100: since any ``roughness'' in the interpolated functions results in failure
2101: of convergence of the numerical algorithm, and the two numerical meshes are
2102: intrinsically non-uniform: Model S has meshpoints strongly 
2103: concentrated near 
2104: the solar surface, near $r=0$ and at
2105: the base of the convection zone $r_{\rm cz}  = 0.713 \rsun$, while
2106: our numerical mesh has meshpoints strongly concentrated near the
2107: domain boundaries $r_{\rm in}$ and $r_{\rm out}$. A new interpolating routine
2108: was created which automatically selects a certain number of points from 
2109: the original mesh, performs the rational interpolation upon the desired mesh, 
2110: and checks for the smoothness of the interpolated function and of its derivative. 
2111: Should the result be inadequate, the procedure is repeated with a different 
2112: set of target points. Selected background quantities are shown in Figure 
2113: \ref{fig:background}. 
2114: 
2115: The quantities $\overline{\nu}$, $\overline{\eta}$, and 
2116: $\overline{k}$ are not provided by Model S, and must instead be calculated
2117: using the formulae derived by Gough (2007):
2118: \begin{eqnarray}
2119: \overline{\nu}(r) &=& \nu_{\rm cz} \left[ 0.1 \left(\frac{\overline{T}}{T_{\rm cz}}\right)^4 \left(\frac{\overline{\rho}}{\rho_{\rm cz}}\right)^{-2} \left(\frac{\overline{\kappa}_{\rm R}}{\kappa_{\rm R,cz}}\right)^{-1} \right. \mbox{   ,   }\nonumber \\
2120:  &+& \left. 0.9  \left(\frac{\overline{T}}{T_{\rm cz}}\right)^{5/2} \left(\frac{\overline{\rho}}{\rho_{\rm cz}}\right)^{-1} \left(\frac{\ln \overline{\Lambda}}{\ln \Lambda_{\rm cz}}\right)^{-1} \right] \mbox{   ,   }\nonumber \\
2121: \overline{\eta}(r) &=& \eta_{\rm cz} \left(\frac{\overline{T}}{T_{\rm cz}}\right)^{-3/2} \left(\frac{\ln\overline{\Lambda}}{\ln \Lambda_{\rm cz}}\right) \mbox{   ,   }\nonumber \\
2122: \overline{k}(r) &=& \overline{\rho} \overline{c}_{\rm p} \overline{\kappa}(r) \mbox{   ,   }
2123: \end{eqnarray} 
2124: where the Coulomb logarithm $\ln\overline{\Lambda}$ and the thermal 
2125: diffusivity $\overline{\kappa}$ are given by 
2126: \begin{eqnarray}
2127: \overline{\Lambda}(r) &=& \Lambda_{\rm cz}\left(\frac{\overline{\rho}}{\rho_{\rm cz}}\right)^{-1/2} \left(\frac{\overline{T}}{T_{\rm cz}}\right)^{3/2} \mbox{   ,   }\nonumber \\
2128: \overline{\kappa}(r) &=& \kappa_{\rm cz} \left(\frac{\overline{T}}{T_{\rm cz}}\right)^{3}\left(\frac{\overline{\rho}}{\rho_{\rm cz}}\right)^{-2} \left(\frac{\overline{\kappa_{\rm R}}}{\kappa_{\rm R,cz}}\right)^{-1} \mbox{   .   }
2129: \end{eqnarray}
2130: The following quantities are those at the base of the convection zone: $T_{\rm cz} = 2.3 
2131: \times 10^6$K, $\rho_{\rm cz} = 0.21$g/cm$^3$, $\kappa_{\rm R,cz} = 19$cm$^2$/g, $\ln\Lambda_{\rm cz} = 2.5$, $\nu_{\rm cz} = 27$cm$^2$/s, $\eta_{\rm cz} = 410$cm$^2$/s and $\kappa_{\rm cz} = 1.4 \times 10^7$cm$^2$/s.
2132: 
2133: \begin{figure}
2134: \includegraphics[width=84mm]{fA1.eps}
2135: \caption{Background pressure, density and temperature (top left corner), 
2136: buoyancy frequency squared (top right corner), microscopic diffusivities 
2137: (bottom left corner) and ratio of diffusivities (bottom right corner) 
2138: from Model S of Christensen-Dalsgaard {\it et al.} (1991).}
2139: \label{fig:background}
2140: \end{figure}
2141: 
2142: 
2143: \section[]{Model equations in spherical coordinates}
2144: 
2145: The model equations described in the system (\ref{eq:global2}) are now 
2146: expanded in a spherical coordinate system $(r,\theta,\phi)$, with 
2147: $\tilde{\bu} = (\tilde{u}_r,\tilde{u}_\theta,\tilde{u}_\phi)$, $\bB = (B_r,B_{\theta},B_{\phi})$ and $\bj = (j_r,j_\theta,j_\phi)$.
2148: \noindent The $r-$component of the momentum equation:
2149: \begin{eqnarray}
2150: - \overline{\rho} \st \left( 2\overline{\Omega} +  \frac{\tilde{u}_\phi}{r\st}\right) \tilde{u}_{\phi} &=& -\frac{\partial \tilde{p}}{\partial r} - \tilde{\rho} \frac{\dd \overline{\Phi}}{\dd r} + j_\theta B_\phi - j_\phi B_\theta \nonumber \\ &+& f_\nu (\div\Pi)_r\mbox{   .   }
2151: \end{eqnarray}
2152: The $\theta-$component of the momentum equation:
2153: \begin{eqnarray}
2154: - \overline{\rho} \ct \left( 2 \overline{\Omega} +  \frac{\tilde{u}_\phi}{r\st} \right) \tilde{u}_{\phi} &=& j_\phi B_r- j_r B_\theta \nonumber \\ &+& f_\nu (\div\Pi)_\theta\mbox{   .   }
2155: \end{eqnarray}
2156: The $\phi-$component of the momentum equation:
2157: \begin{eqnarray}
2158: \overline{\rho} \left(2 \overline{\Omega} + \frac{\tilde{u}_\phi}{r\st}\right) ( \ct \tilde{u}_\theta + \st \tilde{u}_r )  & = & j_r B_\theta - j_\theta B_r \nonumber \\
2159:  &+& f_\nu (\div\Pi)_\phi\mbox{   .   }
2160: \end{eqnarray}
2161: where the role of $f_\nu$ was discussed in section \ref{subsubsec:diffs}.
2162: The divergence of the viscous stress tensor, for a stratified fluid, 
2163: can be derived from Batchelor (1994 edition, pp. 147 and 601).
2164: %\begin{eqnarray}
2165: %(\div \Pi)_r &=& 
2166: %\frac{4}{3} \overline{\rho}\overline{\nu} \frac{\ptl^2 \tilde{u}_r}{\ptl r^2} 
2167: %+ \frac{4}{3}\left( \frac{2}{r} \overline{\rho}\overline{\nu} +\frac{\dd}{\dd r} (\overline{\rho}\overline{\nu} )\right) \frac{\ptl \tilde{u}_r}{\ptl r} 
2168: %\nonumber \\ &-& 
2169: %\frac{4}{3r}\left( \frac{2}{r} \overline{\rho}\overline{\nu}  + \frac{\dd}{\dd r} (\overline{\rho}\overline{\nu} )\right) \tilde{u}_r \nonumber \\ 
2170: %&+& \frac{\overline{\rho}\overline{\nu}}{r^2 \st} \frac{\ptl}{\ptl \theta} \left( \st \frac{\ptl \tilde{u}_r}{\ptl \theta}\right) \nonumber \\ 
2171: %&-& \frac{1 }{3r\st} \left[ \frac{7\overline{\rho}\overline{\nu}}{r} + 2 \frac{\dd}{\dd r} \left(\overline{\rho}\overline{\nu} \right) \right]   \frac{\ptl }{\ptl \theta} \left(\st \tilde{u}_\theta  \right) \nonumber \\ 
2172: %&+& \frac{\overline{\rho}\overline{\nu} }{3r\st} \frac{\ptl^2 }{\ptl \theta \ptl r } \left( \st \tilde{u}_\theta \right)  \mbox{   ,   }
2173: %\end{eqnarray}
2174: %\begin{eqnarray}
2175: %(\div \Pi)_\theta &=& \overline{\rho}\overline{\nu} \frac{\ptl^2 \tilde{u}_\theta}{\ptl r^2} + \left( \frac{2}{r}  \overline{\rho}\overline{\nu} + \frac{\dd}{\dd r} ( \overline{\rho}\overline{\nu})\right) \frac{\ptl \tilde{u}_\theta}{\ptl r} - \frac{\dd}{\dd r} (\rho\nu)  \frac{\tilde{u}_\theta}{r} \nonumber \\ &+& \frac{ \overline{\rho}\overline{\nu}}{3r}  \frac{\ptl^2 \tilde{u}_r}{\ptl r \ptl \theta} + \left( \frac{8}{3r^2} \overline{\rho}\overline{\nu} + \frac{1}{r} \frac{\dd}{\dd r} ( \overline{\rho}\overline{\nu})\right) \frac{\ptl\tilde{u}_r}{\ptl \theta} \\ &+& \frac{4 \overline{\rho}\overline{\nu} }{3r^2\st} \frac{\ptl^2 }{\ptl \theta^2} \left(\st\tilde{u}_\theta \right) - \frac{4 \overline{\rho}\overline{\nu}}{3r^2}  \frac{\ct}{\s2t} \frac{\ptl}{\ptl \theta} \left(\st \tilde{u}_{\theta} \right) \mbox{   ,   } \nonumber
2176: %\end{eqnarray}
2177: %\begin{eqnarray}
2178: %(\div \Pi)_\phi &=& \overline{\rho} \overline{\nu}  \frac{\ptl^2 \tilde{u}_\phi}{\ptl r^2} + \left( \frac{2}{r}  \overline{\rho}\overline{\nu} + \frac{\dd}{\dd r} ( \overline{\rho}\overline{\nu})\right) \frac{\ptl \tilde{u}_\phi}{\ptl r} \nonumber \\
2179: %&-& \left( \frac{\overline{\rho}\overline{\nu}}{r^2 \s2t} + \frac{\dd}{\dd r}( \overline{\rho}\overline{\nu})\right) \frac{ \tilde{u}_\phi}{ r} \nonumber \\
2180: %&+& \frac{\overline{\rho}\overline{\nu} }{r^2} \frac{\ptl^2 \tilde{u}_\phi}{\ptl \theta^2} + \frac{\overline{\rho}\overline{\nu} }{r^2} \frac{\ct}{\st} \frac{\ptl \tilde{u}_\phi}{\ptl \theta}\mbox{   .   }
2181: %\end{eqnarray}
2182: \\
2183: The mass conservation equation:
2184: \begin{equation}
2185: \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \overline{\rho} \tilde{u}_r \right) + \frac{\overline{\rho}}{r \st}\frac{\partial }{\partial \theta} \left (\st \tilde{u}_{\theta} \right) = 0 \mbox{   .   }
2186: \end{equation}
2187: The thermal energy equation can be re-written as (see SZ92)
2188: \begin{equation}
2189: \frac{\overline{\rho} c_{\rm p} \overline{T} N^2 }{g} \tilde{u}_r = \frac{f_\kappa}{r^2} \frac{\partial }{\partial r} \left( r^2 \overline{k} \frac{\partial \tilde{T}}{\partial r} \right)  + \frac{f_\kappa \overline{k}}{r^2 \st } \frac{\partial }{\partial \theta} \left( \st \frac{\partial \tilde{T}}{\partial \theta} \right) \mbox{   ,   }
2190: \end{equation} 
2191: where $ g = \dd \Phi/\dd r$ and $N$ is the Brunt-Vaisala (buoyancy) frequency.
2192: 
2193: \noindent The equation for the conservation of magnetic flux can be integrated once, and axial symmetry implies that
2194: \begin{equation}
2195: \tilde{u}_r B_\theta - \tilde{u}_\theta B_r = f_\eta \overline{\eta} \left[ \frac{1}{r} \frac{\partial}{\partial r}(r B_\phi) - \frac{1}{r} \frac{\partial B_r}{\partial \theta} \right] \mbox{   .   }
2196: \end{equation}
2197: The $\phi-$component of the equation of conservation of magnetic flux:
2198: \begin{eqnarray}
2199: \frac{1}{r}\frac{\partial}{\partial r} \left( r \tilde{u}_\phi B_r - r \tilde{u}_r B_\phi \right) - \frac{1}{r } \frac{\partial }{\partial \theta} \left( \tilde{u}_\theta B_\phi - \tilde{u}_\phi B_\theta \right) \nonumber \\ = - \frac{f_\eta}{r} \frac{\partial }{\partial r} \left[ \overline{\eta} \frac{\partial }{\partial r} (r B_\phi) \right] - \frac{f_\eta \overline{\eta}}{r^2} \frac{\partial}{\partial \theta} \left[ \frac{1}{\st} \frac{\partial}{\partial \theta} ( \st B_\phi) \right]\mbox{   .   }
2200: \end{eqnarray}
2201: Finally, the solenoidal condition is
2202: \begin{equation}
2203: \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2  B_r \right) + \frac{1}{r \st}\frac{\partial }{\partial \theta} \left (\st B_{\theta} \right) = 0\mbox{   .   }
2204: \end{equation}
2205: 
2206: \section[]{Numerical implementation of the equations}
2207: 
2208: Numerical implementation of the equations is done by defining the non-dimensional independent variables $x = r/\rsun$ and $\mu = \ct$, and by working with the non-dimensional dependent variables 
2209: \begin{align}
2210: & \hat{u} = \frac{\tilde{u}_r}{\rsun \Omega_{\rm eq}} \mbox{   ,   } \hat{v} = \frac{\st \tilde{u}_\theta}{\rsun \Omega_{\rm eq}}  \mbox{   ,   } \hat{L} = \frac{r \st \tilde{u}_\phi}{ \rsun^2 \Omega_{\rm eq}} \\ 
2211: & \hat{B} = \frac{B_r}{B_0} \mbox{   ,   } \hat{b} = \frac{\st B_\theta}{B_0}  \mbox{   ,   } \hat{S} = \frac{r \st B_\phi}{\rsun B_0} \mbox{  ,   } \hat{J} =\frac{4\pi r \st j_\phi}{B_0}  \mbox{   .   }  \nonumber 
2212: \end{align}
2213: The perturbed thermodynamical quantities are also normalized with
2214: \begin{equation}
2215: \hat{\rho} = \frac{\tilde{\rho}}{\rho_0}  \mbox{   ,   } \hat{T} = \frac{\tilde{T}}{T_0} \mbox{   ,   } \hat{p} = \frac{\tilde{p}}{p_0}  \mbox{   ,   }
2216: \end{equation}
2217: where $\rho_0 = 1 {\rm g/cm}^3$, $T_0 = 1 {\rm K}$, and $p_0 = \rho_0 \rsun^2 \overline{\Omega}^2$.
2218: 
2219: The symmetries of the system in the latitudial coordinate $\theta$ suggest the expansion 
2220: of the dependent variables onto Fourier modes, which are equivalent to Chebishev polynomials 
2221: of the variable $\mu$ since the $n-$order Chebishev polynomial $T_n$ is defined as
2222: \begin{equation}
2223:  T_n(\mu) = \cos(n\theta) \mbox{   .  }
2224: \end{equation}
2225: We assume symmetry across the equator for the radial velocity, the angular velocity, the 
2226: thermodynamical quantities and the latitudinal component of the magnetic field. Antisymmetry 
2227: across the equator is then required for consistency for the latitudinal velocity as well as 
2228: the radial and toroidal components of the magnetic field. In addition, we require that the 
2229: total mass flux across a spherical surface be null. This leads to the suggested expansion:
2230: \begin{eqnarray}
2231: \hat{u}(x,\mu) &=& \sum_{n=1}^N \psi_n(x) \frac{\ptl}{\ptl \mu} \left((1-\mu^2) T_{2n-1}(\mu)\right) \nonumber \\
2232: \hat{v}(x,\mu) &=& (1-\mu^2) \sum_{n=1}^N v_n(x) T_{2n-1}(\mu) \mbox{   ,} \nonumber \\
2233: \hat{L}(x,\mu) &=& (1-\mu^2) \sum_{n =1}^N L_n(r) T_{2n-2}(\mu) \mbox{   , }\nonumber  \\
2234: \hat{p}(x,\mu) &=& \sum_{n=1}^N p_n(x) T_{2n-2}(\mu) \mbox{   ,   } \nonumber \\
2235: \hat{\rho}(x,\mu) &=& \sum_{n=1}^N \rho_n(x) T_{2n-2}(\mu) \mbox{   ,} \nonumber \\
2236: \hat{T}(x,\mu) &=& \sum_{n=1}^N \Theta_n(x) T_{2n-2}(\mu) \mbox{   ,}\nonumber 
2237: \end{eqnarray}
2238: \begin{eqnarray}
2239: \hat{B}(x,\mu) &=&  \sum_{n=1}^N B_n(x) T_{2n-1}(\mu) \mbox{   ,   } \nonumber \\
2240: \hat{b}(x,\mu) &=& (1-\mu^2)  \sum_{n=1}^N b_n(x) T_{2n-2}(\mu) \mbox{   ,}\nonumber \\
2241: \hat{S}(x,\mu)  &=& (1-\mu^2)  \sum_{n=1}^N S_n(x) T_{2n-1}(\mu) \mbox{   ,   }\nonumber \\
2242: \hat{J}(x,\mu)  &=& (1-\mu^2)  \sum_{n=1}^N J_n(x) T_{2n-1}(\mu) \mbox{   ,}
2243: \label{eq:defexp}
2244: \end{eqnarray}
2245: where each sum is truncated to retain the first $N$ modes only. 
2246: The equations are expanded using 
2247: the ansatz (\ref{eq:defexp}), then projected back onto the first $N$ 
2248: Chebishev modes (of relevant parity). The quadratic terms are simplified 
2249: analytically using the standard formulae 
2250: $2T_{n}(\mu) T_{m}(\mu) = (T_{n+m}(\mu) + T_{n-m}(\mu))$ 
2251: and $T_n(\mu) = T_{-n}(\mu)$, as well as a variety of others that have been 
2252: summarized by Garaud (2001, pp. 200-204). 
2253: This procedure yields a combination of $12N$ first-order ODEs 
2254: and $3N$ algebraic equations for a total of 
2255: 15$N$ dependent variables, namely $\{\psi_n\}$, $\{v_n\}$, 
2256: $\{L_n\}$, $\{T_n\}$ 
2257: and $\{S_n\}$ (and their respective radial first derivatives), as well 
2258: as $\{p_n\}$, $\{\rho_n\}$, $\{B_n\}$, $\{b_n\}$ and $\{J_n\}$ for $n=1..N$. 
2259: 
2260: These equations are solved using a versatile variant of the 
2261: standard Newton-Raphson-Kantorovich (NRK) relaxation solver (see for example 
2262: Press {\it et al.} 1996, chapter 17.3) developped by Garaud 
2263: (2001, pp. 108-111). The added feature compared with the standard 
2264: algorithm permits the solution of any set of equations that can be 
2265: written in the form
2266: \begin{equation}
2267: \sum_{j = 1}^{N_{\rm v}} M_{ij}({\bf y}, x) \frac{\partial y_j}{\partial x} = f_i({\bf y}, x) \mbox{   for  } i = 1..N_{\rm v} \mbox{  ,  }
2268: \label{eq:mynrk}
2269: \end{equation}
2270: where $x$ is the independent variable, ${\bf y}= \{y_i\}$ is the vector 
2271: containing the $N_{\rm v}$ dependent variables, and $M_{ij}({\bf y}, x)$ 
2272: and $f_i({\bf y}, x)$ can be any nonlinear function of $x$ and ${\bf y}$. 
2273: One of the many advantages of this form is that algebraic equations are 
2274: straightforwardly treated as any other equation by setting $M_{ij} =0$ 
2275: for all $j$ and for $i$ corresponding to the relevant equation(s).
2276: 
2277: As in the standard relaxation algorithm (see Press {\it et al.} 1996) 
2278: most of the computational cost (both in terms of time and memory)
2279: arises from reading and inverting a block-tridiagonal matrix 
2280: containing $N_x+1$ rows
2281: of 3 blocks\footnote{except for the boundary blocks, of course.} 
2282: (where $N_x$ is equal to the number of meshpoints used, 
2283: typically between 2000 and 3000), 
2284: and where each block has size\footnote{again, except 
2285: for the boundary blocks, which are smaller.}
2286:  $N_{\rm v} \times N_{\rm v}$  where 
2287: $N_{\rm v}$ is the total number of dependent variables ($N_{\rm v} = 15N$, 
2288: where $N$ is 60--80 for a typical run). 
2289: Calculation of the matrix components, followed 
2290: by its inversion using serial partial pivoting can require several hours 
2291: per iteration on a high-end desktop. In addition, memory becomes an issue 
2292: since the standard algorithm typically requires the storage of $N_x$ 
2293: double-precision arrays of size $N_{\rm v} \times N_{\rm v}/2$ before 
2294: back-substitution can proceed. For this reason, we have developped a 
2295: parallel version of the NRK solver.
2296: 
2297: \section[]{Parallel NRK algorithm}
2298: 
2299: A serial, simple and efficient way of inverting the typical block-tridiagonal
2300:  linear systems arising from two-point boundary value problems is described 
2301: by Press {\it et al.} (1996). Given the matrix structure
2302: \begin{equation}
2303: \begin{bmatrix}
2304: M & R &   &   &   &   &   &   & \\
2305: L & M & R &   &   &   &   &   & \\
2306:   & L & M & R &   &   &   &   & \\
2307:   &   & L & M & R &   &   &   & \\
2308:   &   &   & L & M & R &   &   & \\
2309:   &   &   &   & L & M & R &   & \\
2310:   &   &   &   &   & L & M & R & \\
2311:   &   &   &   &   &   & L & M & R  \\
2312:   &   &   &   &   &   &   & L & M 
2313: \end{bmatrix}
2314: \end{equation}
2315: the algorithm consists in diagonalizing the first top $M$ (middle) block using 
2316: Gauss-Jordan elimination with partial implicit pivoting, and storing the resulting 
2317: modified $R$ block and right-hand side for later back-substitution. Moving onto the 
2318: second block-row, we zero the $L$ block, diagonalize the $M$ block and store the 
2319: modified $R$ block (and modified right-hand-side vector), and so forth. 
2320: The resulting matrix structure is
2321: \begin{equation}
2322: \begin{bmatrix}
2323: I & S &   &   &   &   &   &   &\\ 0 & I & S &   &   &   &   &   &\\ &
2324: 0 & I & S &   &   &   &   &\\ &   & 0 & I & S &   &   &   &\\ &   &
2325: & 0 & I & S &   &   &\\ &   &   &   & 0 & I & S &   &\\ &   &   &   &
2326: & 0 & I & S & \\ &   &   &   &   &   & 0 & I & S\\ &   &   &   &   &
2327: &   & 0 & I
2328: \end{bmatrix}
2329: \end{equation}
2330: where $0$ denotes a ``zeroed'' block,  $I$ is the identity, and $S$ is
2331: a stored block. The last step involves sweeping the matrix back from
2332: bottom to top to back-substitute for the unknown variables. This
2333: method requires minimal storage: firstly, each block-row only needs to
2334: be read just before being processed and secondly, the backsubstitution
2335: step only requires knowledge of the blocks $S$ and modified
2336: right-hand-side vector.
2337: 
2338: Inverting the same linear system in parallel is not obvious since the
2339: previous algorithm is inherently serial. A standard parallel method
2340: used for  block tridiagonal matrices is cyclic reduction (see Golub \&
2341: Ortega, 1993).  In this case, the number of variables is successively
2342: halved by repeating the  following reduction algorithm: diagonalize
2343: all $M$-blocks in odd block-rows by  partial implicit pivoting, then
2344: zero the $R$ and $L$ matrices directly  above and below by Gaussian
2345: elimination. The resulting matrix after the first reduction is:
2346: \begin{equation}
2347: \begin{bmatrix}
2348: I & S &   &   &   &   &   &   & \\
2349: 0 & X & 0 & X &   &   &   &   & \\
2350:   & S & I & S &   &   &   &   & \\
2351:   & X & 0 & X & 0 & X &   &   & \\
2352:   &   &   & S & I & S &   &   & \\
2353:   &   &   & X & 0 & X & 0 & X  & \\
2354:   &   &   &   &   & S & I & S &  \\
2355:   &   &   &   &   & X & 0 & X & 0  \\
2356:   &   &   &   &   &   &   & S & I 
2357: \end{bmatrix}
2358: \end{equation}
2359: where $S$ denotes a block that has been modified and needs to be stored 
2360: for later back-substitution, and $X$ denotes blocks that have been modified 
2361: and will be used in the next reduction step. Note that all of the remaining 
2362: $X$ blocks (in the odd block-rows) form a new block-tridiagonal system 
2363: for roughly half the number of 
2364: variables (more precisely, if the initial number of block-rows is $2^k - 1$ 
2365: then the reduction step reduces it to $2^{k-1}-1$). The algorithm can 
2366: be repeated until only 1 block is left and finally inverted, 
2367: at which point backsubstitution can begin. The main advantage 
2368: of the method is its inherently parallelizable nature, since each process 
2369: can be instructed to reduce a number of blocks more-or-less independently 
2370: of the others; communication between processes is minimal until the final 
2371: steps where the number of remaining block-rows is equal to the number of 
2372: processes. Unfortunately, the stability of this cyclic reduction algorithm 
2373: is much weaker that the stability of the serial algorithm (since the matrix
2374: cannot be globally pivoted), and was found to be unreliable for our purpose. 
2375: 
2376: An alternative parallel algorithm was constructed, loosely 
2377: based on the work of 
2378: Wright (1992). The block-rows are first equally distributed between 
2379: the processes. A first sweeping reduction akin to the serial algorithm 
2380: is performed starting from the top block-row down to the second-to-last 
2381: block-row in each process\footnote{except in the last process, which is 
2382: swept all the way} to transform the matrix to 
2383: \begin{equation}
2384: \setcounter{MaxMatrixCols}{15}
2385: \begin{bmatrix}
2386: I & S &   &   &   &   &   &   &   &   &   &    \\
2387: 0 & I & S &   &   &   &   &   &   &   &   &     \\
2388:   & 0 & I & S &   &   &   &   &   &   &   &       \\
2389:   &   & 0 & X & X &   &   &   &   &   &   &       \\
2390: \hdotsfor{12} \\
2391:   &   &   & S & I & S &   &   &   &   &   &     \\
2392:   &   &   & S & 0 & I & S &   &   &   &   &   \\
2393:   &   &   & S &   & 0 & I & S &   &   &   &     \\
2394:   &   &   & X &   &   & 0 & X & X &   &   &     \\
2395: \hdotsfor{12} \\
2396:   &   &   &   &   &   &   & S & I & S &   &     \\
2397:   &   &   &   &   &   &   & S & 0 & I & S &     \\
2398:   &   &   &   &   &   &   & S &   & 0 & I & S \\
2399:   &   &   &   &   &   &   & X &   &   & 0 & I 
2400: \end{bmatrix}
2401: \end{equation}
2402: (assuming in the case shown here that 12 block-rows are equally 
2403: distributed between 3 processes). Again, by convention all $S$ blocks denote 
2404: blocks that need to be stored for backsubstitution, while $X$ blocks are 
2405: blocks that will be further modified. Until this point, work in each 
2406: process can be done without communication with other processes. 
2407: 
2408: At the end of the sweeping step, each process sends its very last block-row 
2409: to the following process. Thus the blocks are redistributed as
2410: \begin{equation}
2411: \setcounter{MaxMatrixCols}{15}
2412: \begin{bmatrix}
2413: I & S &   &   &   &   &   &   &   &   &   &    \\
2414: 0 & I & S &   &   &   &   &   &   &   &   &     \\
2415:   & 0 & I & S &   &   &   &   &   &   &   &       \\
2416: \hdotsfor{12} \\
2417:   &   & 0 & X & X &   &   &   &   &   &   &       \\
2418:   &   &   & S & I & S &   &   &   &   &   &     \\
2419:   &   &   & S & 0 & I & S &   &   &   &   &   \\
2420:   &   &   & S &   & 0 & I & S &   &   &   &     \\
2421: \hdotsfor{12} \\
2422:   &   &   & X &   &   & 0 & X & X &   &   &     \\
2423:   &   &   &   &   &   &   & S & I & S &   &     \\
2424:   &   &   &   &   &   &   & S & 0 & I & S &     \\
2425:   &   &   &   &   &   &   & S &   & 0 & I & S \\
2426:   &   &   &   &   &   &   & X &   &   & 0 & I 
2427: \end{bmatrix}
2428: \end{equation}
2429: Each process can then continue eliminating undesirable variables from its 
2430: top block-row by Gaussian elimination with the successive block-rows below,
2431: until the following form is achieved:
2432: \begin{equation}
2433: \setcounter{MaxMatrixCols}{15}
2434: \begin{bmatrix}
2435: I & S &   &   &   &   &   &   &   &   &   &    \\
2436: 0 & I & S &   &   &   &   &   &   &   &   &     \\
2437:   & 0 & I & S &   &   &   &   &   &   &   &       \\
2438: \hdotsfor{12} \\
2439:   &   & 0 & X & 0 & 0 & 0 & X &   &   &   &       \\
2440:   &   &   & S & I & S &   &   &   &   &   &     \\
2441:   &   &   & S & 0 & I & S &   &   &   &   &   \\
2442:   &   &   & S &   & 0 & I & S &   &   &   &     \\
2443: \hdotsfor{12} \\
2444:   &   &   & X &   &   & 0 & X & 0 & 0 & 0 & X    \\
2445:   &   &   &   &   &   &   & S & I & S &   &     \\
2446:   &   &   &   &   &   &   & S & 0 & I & S &     \\
2447:   &   &   &   &   &   &   & S &   & 0 & I & S \\
2448:   &   &   &   &   &   &   & X &   &   & 0 & I 
2449: \end{bmatrix}
2450: \end{equation}
2451: At this point, the top-rows in each process can be combined into 
2452: a much-reduced block tri-diagonal system. Solution from here on proceeds
2453: with a cyclic reduction of the remaining blocks across processes, 
2454: followed by a back-substitution step.
2455: 
2456: The operation count of this algorithm is the same as that of 
2457: the standard cyclic reduction. The advantages of this algorithm over
2458: the standard cyclic reduction are two-fold. Firstly, it is found to be much 
2459: more stable (Wright, 1992). Secondly, it is far more versatile, since 
2460: the standard algorithm only performs ideally with a number of block-rows 
2461: equal to $2^k + 1$ (for any integer $k$), which seriously constrains 
2462: the number of meshpoints one is allowed to select. On the other hand, 
2463: in this algorithm it is always possible to balance the load (for any 
2464: number of meshpoints) in such a way that any process contains at most one more
2465: block-row than the average. 
2466: 
2467: The scalability of the algorithm is naturally excellent until the number 
2468: of processes approaches a tenth of the number of meshpoints. 
2469: This is illustrated in Figure \ref{fig:procscal}.
2470: 
2471: \begin{figure}
2472: \centerline{\epsfig{file=fD1.eps,width=8.4cm,height=7cm}}
2473: \caption{Scaling properties of the parallel NRK algorithm RELAX 
2474: for the simulations 
2475: presented in this paper. The stars correspond to a simulation with 60 
2476: modes (i.e. 900 equations), and 1000 meshpoints. The diamonds correspond to 
2477: the same simulation but with 2000 meshpoints.}
2478: \label{fig:procscal}
2479: \end{figure}
2480: 
2481: Finally, we note that the user-interface of the complete associated 
2482: parallel version of the Newton-Raphson-Kantorovich algorithm (RELAX) 
2483: is entirely identical to that of the widely used serial NRK algorithm 
2484: written by Gough \& Moore. Transfer between the parallel and serial
2485: version should be transparent to the user. The RELAX algorithm is freely 
2486: available upon request from P. Garaud.
2487: 
2488: 
2489: 
2490: \bsp
2491: 
2492: \label{lastpage}
2493: 
2494: \end{document}
2495: 
2496: 
2497: