0805.2805/VKS.tex
1: %%STYLE PRL double colonne, ca coince avec les figures! le 16/04/08 mais ca tient en 4 pages!
2: \documentclass[aps,prl,twocolumn,showpacs,groupedaddress]{revtex4}  % for review and submission 
3: %\documentclass[aps,preprint,showpacs,groupedaddress]{revtex4}  % for double-spaced preprint
4: \usepackage{graphicx}  % needed for figures
5: \usepackage{dcolumn}   % needed for some tables
6: \usepackage{bm}        % for math
7: \usepackage{amssymb}   % for math
8: \usepackage{ulem}
9: %%STYLE REVTEX4
10: %\documentclass[12pt,showpacs,preprintnumbers,amsmath,amssymb,floatfix]{revtex4}
11: %\linespread{2} 
12: %%STYLE REVTEX4
13: %\DeclareGraphicsExtensions{.jpg,.pdf}
14: %\linespread{2} 
15: %\usepackage[pdftex]{color}
16: %\usepackage[pdftex,
17: %colorlinks=true,linkcolor=blue,citecolor=blue,urlcolor=blue]{hyperref}
18: \usepackage{epsfig} 
19: \usepackage{array} 
20: \usepackage{amsmath} 
21: \usepackage{lscape}
22: \usepackage{color}
23: %\usepackage{natbib}
24: \usepackage{graphicx}
25: %
26: %\newcommand{\balise}[1]{\textsc{[#1]}}
27: \newcommand{\bal}[1]{\textcolor{black}{#1}}
28: \newcommand{\ise}[1]{\textcolor{blue}{#1}}
29: %
30: \begin{document}
31: %\preprint{}
32: \title{Impact of  impellers on the axisymmetric magnetic mode in the VKS2 dynamo experiment}
33: \author{R. Laguerre,$^{1,2}$ C. Nore,$^2$
34:  A. Ribeiro,$^2$ J. L\'eorat,$^3$ J.-L. Guermond$^4$
35: and F. Plunian$^5$} 
36: 
37: \email{nore@limsi.fr}
38: 
39: \affiliation{$^1$ Universit\'e Libre de Bruxelles, CP.231, Boulevard
40:   du Triomphe, Brussels, 1050, Belgium; $^2$Laboratoire d'Informatique
41:   pour la M\'ecanique et les Sciences de l'Ing\'enieur, CNRS,
42:   Universit\'e Paris-Sud 11 BP 133, 91403 Orsay cedex, France;
43:   $^3$Luth, Observatoire de Paris-Meudon, place Janssen, 92195-Meudon,
44:   France; $^4$Department of Mathematics, Texas A\&M University,
45:   College Station, TX 77843, USA; $^5$ Universit\'e Joseph
46:   Fourier, CNRS, Laboratoire de G\'eophysique Interne et de
47:   Tectonophysique, 38041 Grenoble, France} \date{\today}
48: \begin{abstract}
49:   In the VKS2 (von K\'arm\'an Sodium 2) successful dynamo experiment of September 2006, the
50:   magnetic field that was observed showed a strong axisymmetric
51:   component, implying that non axisymmetric components of the flow field
52:   were acting.  By modeling the induction effect of the spiraling flow
53:   between the blades of the impellers in a kinematic dynamo code, we
54:   find that the axisymmetric magnetic mode is excited and becomes
55:   dominant in the vicinity of the dynamo threshold.
56: %
57:   The control parameters are the magnetic Reynolds number \bal{of the mean flow}, the
58:   coefficient measuring the induction effect, $\alpha$, and the type
59:   of boundary conditions (vacuum for steel impellers and normal field
60:   for soft iron impellers). We show that using realistic values of
61:   $\alpha$, the observed critical magnetic Reynolds number, $Rm^c
62:   \approx 32$, can be reached easily with ferromagnetic boundary
63:   conditions. We conjecture that the dynamo action achieved in this
64:   experiment \bal{may not be} related to the turbulence in the bulk of the flow,
65:   but 
66:   \bal{rather to} the alpha effect \bal{induced by} the impellers.
67: \end{abstract}
68: \pacs{47.65.-d, 52.65Kj, 91.25Cw}
69: \maketitle
70: 
71: 
72: %\section{Introduction}
73: %%%%%%%%%%%%%%%%%%%%%%
74: \bal{The interest of the scientific community for the dynamo action 
75: in liquid metal has been renewed since 2000 in the wake of successful
76: experiments~\cite{Ga2000,StMu01,Monchaux06}. We focus in this Letter on
77: the Cadarache VKS2 (von K\'arm\'an Sodium 2) experiment~\cite{Monchaux06} where the dynamo
78: effect occurred beyond the critical magnetic Reynolds number $Rm^c
79: \approx 32$.}
80: \bal{The growing magnetic field that was observed
81: above threshold was mainly axisymmetric.
82: Dynamo action was found with soft iron impellers but did not occur with
83: stainless steel impellers, using the same available power.}  
84: %
85: \bal{The purpose of the present Letter is to present arguments based on
86: modeling and numerical simulations to justify the observed axisymmetric mode.
87: The two key ingredients are the so-called
88: $\alpha$-effect and ferromagnetic boundary conditions.}
89: 
90: From Cowling's theorem
91: \cite{Cowling34} the axisymmetric part of the flow cannot be
92: responsible alone for the generation \bal{of an} axisymmetric field.
93: \bal{In the experiment, the} deviation from axisymmetry of the velocity field is thus expected
94: to participate to the dynamo process, generating an axisymmetric
95: electromotive force somewhere in the fluid and triggering the dynamo
96: growth.  Two other magnetohydrodynamic experiments using
97: counter-rotating propellers have also led to this conclusion
98: either on observational~\cite{Lathrop00,Forest06} or on numerical
99: grounds~\cite{Forest07}. It was argued in
100: \cite{Lathrop00,Forest06,Forest07} that turbulence effects were
101: responsible for the occurrence of the axisymmetric magnetic field. \bal{
102: In~\cite{Petrelis07} the turbulence effects were held responsible for the
103: dynamo threshold in VKS2. In addition an $\alpha$-effect close to the discs was
104: invoked to justify the axisymmetry of the generated field.} 
105:  
106: The VKS2 configuration is modeled as follows: The computational domain
107: is divided into three concentric cylinders of axis $z$, embedded in an
108: insulating sphere (figure~\ref{meanflow}).  The inner cylinder $r\le
109: R_0$ contains the moving sodium of conductivity $\sigma_0$.  The
110: cylindrical shell $R_0 \le r \le R_1$ contains stagnant sodium
111: ($\sigma_1=\sigma_0$). The cylindrical shell $R_1 \le r \le R_2$ is a
112: layer of copper ($\sigma_2=4.5 \sigma_0$).  We choose $R_0 $ as unit
113: length and the geometric parameters are $R_1/R_0=1.4$, $R_2/R_0=1.6$
114: and $R_{out}/R_0=10$, with the same height $H/R_0=1.8$ for the three
115: cylinders.
116: %
117: The velocity field of the liquid sodium $\textbf{U}$ ($r<R_0$) is
118: axisymmetric and is obtained from time averaged measurements of a
119: water experiment in the VKS2 setting~\cite{Ravelet05}.  The meridional
120: flow \bal{field and the isolines of the azimuthal component of the velocity}
121: are shown in figure~\ref{meanflow}.
122: \begin{figure}
123: \epsfig{file=figure1.eps,width=0.3\textwidth}
124: \caption{Geometry of the computational domain in a meridional plane.
125:   The axisymmetric meridional flow $(U_r,U_z)$ is represented on the
126:   left with arrows. On the right the isovalues of $U_{\theta}$ are
127:   plotted. The radii $R_i$ and height $H$ as well as the
128:   conductivities $\sigma_i$ are defined in the text.}
129: \label{meanflow}
130: \end{figure}
131: The flow field is maintained by two counter-rotating impellers 
132: located at both ends of the vessel.
133: They each occupy a cylindrical volume of radius $R_0 $ and thickness $H/20$.
134: 
135: The only non axisymmetric parts of the device are the blades composing
136: the impellers (eight each), and we represent the effect of these blades
137: by adding the electromotive force $\alpha
138: (\textbf{B}\cdot\textbf{e}_{\theta}) \textbf{e}_{\theta}$ to the
139: induction equation. This force is assumed to act only in the two fluid
140: cylinders occupied by the blades as they rotate.  This model is
141: deduced as follows: As the blades rotate, the fluid is ejected
142: radially outwards and radial vortices are created due to the rotation
143: rate gradient, see the white arrows in figure~\ref{blades}.
144: We
145: then assume that when a magnetic field is applied, the mean induced
146: current is dominated by its azimuthal component and that the shearing
147: action of the vortices is cumulative. The axial current is expected to
148: vanish in the vicinity of the discs \bal{and, as a result, we set the $\alpha_{zz}$
149: component of the $\alpha$ tensor to be zero}. This phenomenon is illustrated on
150: fig.~\ref{blades}. 
151: As the magnetic lines
152: $(\textbf{B}\cdot\textbf{e}_{\theta}) \textbf{e}_{\theta}$ (thin line
153: in figure~\ref{blades}) are distorted by the helical vortices, small
154: scale magnetic loops are created, which in turn create an azimuthal
155: electromotive force $\alpha
156: (\textbf{B}\cdot\textbf{e}_{\theta})\textbf{e}_{\theta}$, where the
157: parameter $\alpha$ is negative and depends on the impeller rotation
158: rate (see below for an estimate), the size, curvature and number of
159: blades. We thus attempt to represent the complexity of the induction
160: effects in a poorly known flow close to a boundary by a single scalar
161: parameter.  The handedness of the helical vortices generated between
162: the blades is directly linked to the rotation sign
163: of the impellers: orienting the rotation axis of the bottom impeller
164: towards the center of the container, a positive rotation
165: drives a set of eight radial positive
166: helices creating a negative $\alpha$ coefficient.
167: \bal{
168: The counter-rotation configuration with $\Omega_{\text{bot}}=
169: -\Omega_{\text{top}}$ is invariant by rotation of $\pi$ about any horizontal
170: axis~\cite{NTDX03}, therefore the two impellers produce the same
171: $\alpha$. Actually, whatever} \bal{the sign for the rotation of the bottom impeller, $\Omega$,
172: the product $\Omega \alpha$ is always negative. Henceforth we set $\Omega_{\text{bot}}=\Omega >0$. }
173: \begin{figure}
174: \includegraphics[width=0.4\textwidth]{figure2.eps} \\*[0cm]
175: \caption{Sketch of the top impeller with eight radial blades attached
176:   to the disc. Due to the rotation of the disc (gray arrow on the
177:   top), radial helical vortices are produced between the blades (thick
178:   white arrows). An axisymmetric azimuthal field
179:   $(\textbf{B}\cdot\textbf{e}_{\theta}) \textbf{e}_{\theta}$ (thin
180:   line) is distorted leading to eight magnetic loops.}
181: \label{blades}
182: \end{figure}
183: 
184: 
185: %\section { Numerical set up and parameters}
186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187: We follow the standard kinematic dynamo approach by solving the
188: induction equation
189: \begin{equation}
190:   \partial_t \textbf{B} = \nabla \times (\widetilde{\textbf{U}} \times \textbf{B} 
191:   + \alpha (\textbf{B}\cdot\textbf{e}_{\theta}) \textbf{e}_{\theta}) 
192:   - \nabla {\times} (\eta \nabla{\times}\textbf{B}).
193: \label{inducequation}
194: \end{equation}
195: where $\eta = 1/\sigma\mu_0$, $\mu_0$ being the vacuum magnetic
196: permeability, $\sigma$ the fluid conductivity and $\alpha$ the unknown
197: parameter which models the induction effect in the volume occupied in
198: average by the impellers.  The field $\widetilde{\textbf{U}}$ is equal
199: to $\textbf{U}$ in the cylinder $ r\le R_0$ and is zero in the two
200: cylindrical shells $R_0\le r\le R_1$, $R_1\le r\le R_2$. The
201: conductivity field is such that $\sigma=\sigma_0$ in the inner
202: cylinder $ r\le R_0$ and in the stagnant sodium shell, and
203: $\sigma=\sigma_2$ in the copper layer.  The conductivity jump at
204: $r=R_1$ is accounted for by enforcing the continuity of the tangent
205: component of the electric field.  Note that the flow behind the
206: impellers has not been simulated (see discussion below). To simplify
207: the parameterization of the alpha effect, $\alpha$ vanishes outside two
208: thin cylindrical layers of radius $R_0 $ and thickness $H/20$ situated at each end
209: of the container and its intensity varies smoothly within these
210: layers \bal{using a $\tanh$ regularization function}.  
211: The magnetic Reynolds number is $Rm \equiv \sigma_0 \mu_0
212: U_0R_0$ where $U_0$ is \bal{the} maximum speed of the flow $\textbf{U}$.
213: 
214: Two types of boundary conditions are compared to model the
215: experiment: they are referred to as insulating (I) and ferromagnetic
216: (F) boundary conditions.  We say that insulating boundary conditions
217: (I) are used when the continuity of the magnetic field is enforced at
218: the interface between the conducting regions and the vacuum (i.e., at
219: ${|z|=H/2, \, 0\le r \le R_2}$ and at ${0 \le |z| \le H/2, \,
220:   r=R_2}$). We say that ferromagnetic boundary conditions (F) are used
221: when we enforce $\textbf{B}\times \textbf{e}_z=0$ at ${|z|=H/2, \,
222:   0\le r \le R_0}$ and we require $\textbf{B}$ to be continuous across
223: the other boundaries ${0 \le |z| \le H/2, \, r=R_2}$ and ${|z|=H/2, \, R_0\le r \le R_2}$.  The condition
224: $\textbf{B}\times \textbf{e}_z=0$ is meant to mimic the effect of
225: ferromagnetic discs with infinite magnetic permeability located at the
226: top and bottom of the vessel\bal{~\cite{Durand68}}.  These boundary conditions model steel
227: impellers (I) and soft iron impellers (F), respectively.
228: 
229: Our numerical code~\cite{Laguerre06, Guermond07} uses a finite element
230: Galerkin method for the spatial discretization in the meridional plane
231: $(r,z)$ and a Fourier series decomposition along the azimuthal
232: direction.  Since the coefficients of the induction equation do not
233: depend on the azimuthal angle, the different azimuthal modes
234: ($m=0,1,2, \ldots$) are decoupled. The time is discretized with a
235: semi-implicit backward finite difference method of second order.
236: Growth rates are computed for the two types of magnetic boundary
237: conditions, (I) or (F), and for various values of $\alpha$ and $Rm$. The
238: zero growth rate condition defines a critical curve in the ($\alpha$,
239: $Rm$) plane. A compilation of results obtained for the (I) and (F)
240: cases and for Fourier modes $m=0$ and $m=1$ is shown in figure~\ref{marginal}.
241: 
242: 
243: %\section{Results and Discussion}
244: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
245: \begin{figure}
246:   \includegraphics[width=0.4\textwidth]{figure3.eps} \\*[0cm]
247:   % \includegraphics[width=0.4\textwidth]{Rmc_alpha_comp.eps} \\*[0cm]
248:   \caption{Dynamo threshold $Rm^c$ versus $-\alpha$, for ($\times$,
249:     $\triangle$) $m=0$, ($\square$,$\diamond$) $m=1$ and where (F) and
250:     (I) indicate ferromagnetic and insulating boundary conditions
251:     respectively.}
252: \label{marginal}
253: \end{figure}
254: 
255: \begin{figure}
256:   \begin{center}
257: \epsfig{file=figure4a.eps,width=0.4\textwidth}
258: \epsfig{file=figure4b.eps,width=0.45\textwidth}
259: %  % \includegraphics[width=0.4\textwidth]{marginal.eps} \\*[0cm]
260: %  % \includegraphics[width=0.4\textwidth]{marginal.eps} \\*[0cm]
261:   \caption{Top: Geometry of the $m=0$ (F) growing magnetic field for
262:     $Rm^c=32$, $|\alpha|=2.2 > |\alpha_c|=2.1$ . The three cylinders
263:     are also represented.  The lines correspond to the positive
264:     isovalues of $B_{\theta}$ and the arrows to the meridional field
265:     $(B_r,B_z)$. Bottom: isovalue of 25\% of the maximum energy
266:     plotted in the 3D domain, together with some field lines.}
267: \label{field}
268: \end{center}
269: \end{figure}
270: In order to make comparisons with experimental results published
271: in~\cite{Monchaux06}, we first focus our attention on the axisymmetric
272: magnetic mode ($m=0$), which was a priori eliminated in former studies
273: by Cowling's theorem.
274: \bal{The two critical curves corresponding to the boundary
275: conditions (I) and (F) and the Fourier mode $m=0$ reveal
276: that the dynamo action occurs only for $\alpha$ negative,
277: thus confirming the phenomenological
278: argument discussed above to justify the sign of $\alpha$.} 
279: Both curves clearly show that
280: increasing $|\alpha|$ lowers the critical magnetic Reynolds
281: number. Moreover, the (F) boundary condition yields significantly
282: smaller thresholds. This observation agrees with previous results
283: obtained in a different context~\cite{Avalos03, Avalos05}. Both curves
284: can be represented by the scaling law $Rm^c \sim |\alpha|^{-q}$, where
285: $q \approx 0.67$, suggesting that the dynamo process is similar for
286: both types of boundary conditions.
287: 
288: The value of $\alpha$ corresponding to the dynamo threshold in VKS2,
289: $Rm^c=32$, is found to be $|\alpha|=2.1$ for the (F) boundary condition and
290: $|\alpha|=3.3 $ for the (I) boundary condition.  Since the steel impellers have not
291: produced dynamo action, this suggests that, as a first guess, the
292: effective $|\alpha|$ for VKS2 is between 2.1 and 3.3.  A rough
293: estimate of $\alpha$ is given by $\alpha \sim u^2 h /\eta$ where $u$
294: is a typical flow intensity between the blades and $h=H/20$. From
295: \cite{Ravelet05} we estimate $u/U_0\sim 0.2$ at the impeller half
296: radius. Then, for $Rm=32$, $\eta$=0.1 m$^2$.s$^{-1}$ and $R_0=0.2$m,
297: we find $\alpha \sim 1.8$m.s$^{-1}$, making the value $|\alpha|$ =
298: 2.1m.s$^{-1}$ plausible.  If $|\alpha|$ is indeed close to 2.1, the critical
299: value of $Rm$ for the (I) boundary condition is close to 43, which
300: apparently contradicts the experiments since $Rm= 50$ has been reached
301: without dynamo action using steel impellers.  To explain this
302: discrepancy we recall that it has been shown in \cite{Stefani06} that
303: for the Fourier mode $m=1$ the flow of liquid sodium behind the
304: impellers acts against the dynamo action. We expect this anti-dynamo
305: effect to be active also for the Fourier mode $m=0$, but we did not
306: include this extra layer of fluid in the present study.
307: % This could be done in a near future, if urged by experimental
308: % results.
309: The negative consequences of this secondary flow are probably screened
310: in the case of soft iron impellers. The numerical evidence justifying this
311: assertion has to await the availability of a code describing
312: conducting domains with different magnetic permeabilities (work in progress).
313: %As mentioned earlier, another argument to answer the discrepancy is
314: %that the $\alpha$ coefficient itself may well depend on the boundary
315: %conditions and be less efficient for insulating than for ferromagnetic
316: %boundary conditions.
317: 
318: The stationary eigenvector corresponding to the (F) boundary condition
319: for $m=0$ and $|\alpha|=2.2$ is shown in figure~\ref{field}.  The
320: magnetic field is mainly concentrated along the $z$-axis in the fluid
321: region $r\le R_0$, and it is azimuthally dominated 
322: in the plane $z=0$ 
323: for $r\ge R_0$. The radial component is odd with
324: respect to $z$ whereas the azimuthal and vertical components are even
325: and of opposite sign. These features are compatible with the 
326: magnetic field measured at saturation in the experimental dynamo regime
327: obtained with soft iron impellers~\cite{Monchaux06}.
328: % The magnetic components recorded by the magnetometer located in the
329: % equatorial plane close to the copper inner wall fluctuate around a
330: % quasi-steady state dominated by the azimuthal component.
331: The eigenvector corresponding to the (I) boundary condition
332: for $m=0$ and $|\alpha|=2.2$ (not shown here)
333: is associated to an eigenvalue with a nonzero imaginary part. The resulting dynamo is periodic
334: with a reversing dipolar moment. This result, not detailed
335: here, is in strong contrast with the case $\alpha=0$ (I), for which
336: a non oscillatory transverse dipole mode ($m=1$) with
337: internal banana-like structures is the only 
338: unstable mode~\cite{Ravelet05}.
339: 
340: We have also investigated the effect of the $\alpha$ model on the
341: Fourier mode $m=1$. This mode has not been observed in the VKS2
342: experiment although it has been predicted to be the most unstable in
343: axisymmetric kinematic dynamo
344: simulations~\cite{Ravelet05,Marie06,Stefani06,Laguerre_thesis}.
345: % \textcolor{red}{The geometry of the growing $m=1$ mode is
346: %   represented in figure~\ref{field2}. The $\theta=0$ view displays
347: %   the horizontal dipole while the $\theta=\pi/2$ view shows one of
348: %   the two banana-like vertical structures aligned with the cylinder
349: %   axis. The structure is dominated by the dipole near $r=R_0$ as
350: %   in~\cite{Ravelet05,Stefani06}.}
351: %Our results are reported in  for both boundary
352: %conditions (I) and (F).  
353: \bal{In figure~\ref{marginal}, for the mode $m=1$, we see that} the critical magnetic Reynolds
354: number grows sharply with $|\alpha|$ for $\alpha <0$ and
355: decreases with $|\alpha|$ for $\alpha >0$. We are thus led to
356: conclude that the growth of the $m=1$ mode is hindered by the
357: very induction effect that generates the $m=0$ mode!
358: This surprising result may explain the absence
359: of the $m=1$ mode in the experiment, at least for $Rm < 50$.
360: 
361: %\section{discussion}
362: 
363: %We have not investigated a
364: %more complex spatial distribution of helicity in the bulk of the flow,
365: %which requires first an experimental investigation using a water
366: %model.
367: %%% Contrary to what is expressed in other contributions of this special
368: %%% CRAS issue [????] , we expect that turbulent fluctuations have a
369: %%% negative impact on dynamo action in the small $ Rm$ regime reached
370: %%% by the experiments by increasing the effective magnetic diffusivity.
371: %A requirement of the present mean induction model is to get a better
372: %knowledge of the impeller flow.  A more quantitative estimate would
373: %involve numerical simulations using a turbomachinery type of code,
374: %with an appropriate subgridscale modeling needed for large Reynolds
375: %number flows. This non axisymmetric flow could then be used in a
376: %kinematic code to derive the corresponding alpha tensor.  Specific
377: %laboratory experiments of induction within an assembly of vortices
378: %could also settle the value of the induced electromotive force.  Such
379: %an experiment could be related to the alpha-box used in~\cite{SKR66}
380: %to illustrate the $\alpha$ effect in a helical flow with a low level
381: %of turbulence.  Note that this configuration is a simple restriction
382: %to a plane slab of the periodic volume filling flow of G.O. Roberts
383: %which has been used in the Karlsruhe experiment~\cite{Muller01}. One
384: %could also act directly on the alpha effect by ad-hoc changes of the
385: %impeller geometry.
386: 
387: %{\bf{Franck, voici la partie a travailler}}
388: %Whatever the improvements which may be brought to the numerical model,
389: %a conclusion of wider scope can be drawn: The $\alpha$-effect is an efficient
390: %mechanism to generate a poloidal field from a toroidal component in
391: %confined domains. While the VKS design and its numerical optimization
392: %procedure focused on bulk flow properties such as the poloidal to
393: %toroidal ratio, our model points out the complementary role of
394: %disconnected domains. This property is indeed present in classical
395: %dynamo models (such as Herzenberg dynamo) but also in astrophysical
396: %dynamos. For example, in the Babcock-Leighton phenomenological model
397: %of the solar activity cycle, the poloidal field is generated from the
398: %toroidal field within a latitude band of a surface layer~\cite{BL}. In
399: %the ``interface dynamo'' of Parker~\cite{Parker93}, to avoid a
400: %``catastrophic quenching'', the $\alpha$-effect operates away from the
401: %differential rotation. One could also describe the geomagnetic dynamo
402: %as the result of the interplay of differential rotation in the bulk of
403: %the flow and small scale vortices in the shear layer tangent to the
404: %solid inner core.  To distinguish the above dynamos within the wide class
405: %of the homogeneous dynamos we propose to name them ``partitioned dynamos.''
406: %Whereas in industrial dynamos, spatial domains correspond to different
407: %circuits (electrical, magnetic or moving parts), in partitioned dynamos
408: %the spatial domains correspond to different functionalities
409: %such as poloidal and toroidal field mutual conversion.
410: %{\bf{Franck, fin de la partie a travailler}}
411: 
412: The present study shows that the positive results of the dynamo
413: experiment of September 2006 (VKS2) as well as the former negative
414: results (VKS1) may be explained by a simple mean induction effect
415: induced by the radial vortices trapped between the blades of the impellers.
416: \bal{We conjecture that turbulence may not play a role as essential as 
417: initially believed in the VKS experiments.}
418: 
419: \bal{This work also shows the importance of
420: using ferromagnetic materials, corresponding to the (F) boundary conditions in fig.~\ref{marginal}.
421: A research program aiming at exploring the saturation regime
422: using nonlinear computations and materials with
423: heterogeneous magnetic permeabilities is currently engaged.}
424: 
425: 
426: This work was supported by ANR project no. 06-BLAN-0363-01
427: ``HiSpeedPIV''.  We are pleased to acknowledge the Saclay VKS-team for
428: providing us with a mean velocity field measured in a water
429: experiment.  We acknowledge fruitful discussions with F.~Stefani.  We
430: warmly thank D. Carati and B. Knaepen, the organizers of the MHD
431: Summer Program, Bruxelles, July 2007. The computations were carried
432: out on the IBM Power 4 computer of IDRIS of CNRS (project \# 0254).
433: 
434: \begin{thebibliography}{}
435: %
436: % and use \bibitem to create references.
437: %
438: %\bibitem{Gilbert03}
439: %A.D. Gilbert,
440: %``Dynamo theory'',
441: %In: Handbook of Mathematical Fluid Dynamics, \textbf{2} (ed. S. Friedlander, D. Serre), 355-441, Elsevier, New-York (2003)
442: 
443: \bibitem{Ga2000}
444: A. Gailitis, O. Lielausis, S. Dement'ev, E. Platacis and A. Cifersons,
445: % Detection of a flow induced magnetic field eigenmode in the Riga dynamo facility,
446:  {\it Phys. Rev. Lett.}, {\bf 84}, 4365 (2000).
447: 
448: \bibitem{StMu01}
449: R. Stieglitz and U. M\"uller,
450: % Experimental demonstration of a homogeneous two-scale dynamo,
451:  {\it Phys. Fluids}, {\bf 13}, 561 (2001).
452:  
453: \bibitem{Monchaux06}
454: R. Monchaux et al., 
455: %M. Berhanu, M. Bourgoin, Ph. Odier, M. Moulin, J.-F. Pinton, R. Volk, S. Fauve, N. Mordant, F. P\'etr\'elis, A. Chiffaudel, F. Daviaud, B. Dubrulle, C. Gasquet, L. Mari\'e, and F. Ravelet,
456: %``Generation of magnetic field by a turbulent flow of liquid sodium'', 
457: Phys. Rev. Lett. \textbf{98}, 044502 (2007).
458: 
459: \bibitem{Cowling34}
460: T.G. Cowling,
461: %``The magnetic field of sunspots'', 
462: Mon. Not. Roy. Astr. Soc. \textbf{94}, 39-48 (1934).
463: 
464: \bibitem{Lathrop00}
465: N. Peffley, A. Cawthorne and D. Lathrop,
466: %``Toward a self-generating magnetic dynamo: The role of turbulence'',
467: Phys. Rev. E, 61, 5287 (2000).
468: 
469: \bibitem{Forest06}
470: E. J. Spence, C. B. Forest, M. D. Nornberg and R. D. Kendrick, 
471: %``Observation of a Turbulence-Generated Large Scale Magnetic Field'', 
472: Phys. Rev. Lett. 96, 055002 (2006). 
473: 
474: \bibitem{Forest07}
475: R. A.~Bayliss, C. B.~Forest, M. D.~Nornberg, E. J.~Spence and P. W.~Terry,
476: %``Numerical simulations of current generation and dynamo excitation in a mechanically forced turbulent flow'',
477: Phys. Rev. E \textbf{75}:2, 026303 (2007).
478: 
479: 
480: \bibitem{Petrelis07}
481: F. P\'etr\'elis,  N. Mordant and S. Fauve,
482: %``On the magnetic fields generated by experimental dynamos'',
483: Geophys. Astrophys. Fluid Dyn. \textbf{101}, 289 (2007).
484: 
485: %\bibitem{Avalos07}
486: %R. Avalos-Zun\~niga, F. Plunian and K.-H. R\"adler
487: %%``Rossby waves and alpha-effect'',
488: %submitted to Geophys. Astrophys. Fluid Dyn. (2008).
489: 
490: 
491: \bibitem{Ravelet05}
492: F. Ravelet, A. Chiffaudel, F. Daviaud and J. L\'eorat,
493: %``Towards an experimental von Karman dynamo: numerical studies for an optimized design'',
494: Phys. Fluids \textbf{17}, 117104  (2005).
495: 
496: \bibitem{NTDX03}
497: C. Nore, L. S. Tuckerman, O. Daube and S. Xin,
498: %``The 1:2 mode interaction in exactly counter-rotating von K\'arm\'an swirling flow'',
499: J. Fluid Mech., {\bf 477}, 51-88 (2003). 
500: 
501: \bibitem{Durand68}
502: E. Durand, {\it Magn\'etostatique}, Masson (1968).
503: 
504: \bibitem{Laguerre06}
505: R. Laguerre, C. Nore, J. L\'eorat and J. L. Guermond,
506: %``Effects of conductivity jumps in the envelope of a kinematic dynamo flow'',
507: CR M\'ecanique \textbf{334}, 593 (2006).
508: 
509: \bibitem{Guermond07}
510: J. L. Guermond, R. Laguerre, J. L\'eorat and C. Nore,
511: %``An Interior Penalty Galerkin Method for the MHD equations in heterogeneous domains'',
512: J. Comp. Physics \textbf{221}, 349 (2007).  
513: 
514: 
515: \bibitem{Avalos03}
516: R. Avalos-Zu\~{n}iga, F. Plunian and A. Gailitis,
517: %``Influence of electro-magnetic boundary conditions onto the
518: %onset of dynamo action in laboratory experiments'',
519: Phys. Rev. E \textbf{68}, 066307 (2003).
520: 
521: \bibitem{Avalos05}
522: R. Avalos-Zu\~{n}iga and F. Plunian,
523: %``Influence of inner and outer walls electromagnetic properties on the onset of a stationary
524: %dynamo'',
525: Eur. Phys. J. B \textbf{47}, 127 (2005).
526: 
527: \bibitem{Stefani06}
528: F. Stefani et al.,
529: %, Xu, M., Gerbeth, G., Ravelet, F., Chiffaudel, A., Daviaud, F., Leorat, J.
530: %``Ambivalent effects of added layers on steady kinematic dynamos in cylindrical geometry: application to the VKS experiment''
531: Eur. J. Mech. B/Fluids \textbf{25} 894 (2006).
532: 
533: %\bibitem{Laval06}
534: %J-P. Laval, P. Blaineau, N. Leprovost, B. Dubrulle, and F. Daviaud,
535: %``Influence of Turbulence on the Dynamo Threshold'',
536: %Phys. Rev. Lett. \textbf{96}, 204503 (2006). 
537: 
538: %\bibitem{Ponty07}
539: %Y. Ponty, P.D. Mininni, H. Politano, J.-F. Pinton and A. Pouquet,
540: %``Dynamo action at low magnetic Prandtl numbers:
541: %mean flow vs. fully turbulent motion'',
542: %New J. Physics, 9, 296  (2007).
543: 
544: %\bibitem{Peyrot07}
545: %M. Peyrot, F. Plunian and C. Normand,
546: %``Parametric instability of the helical dynamo'',
547: %Phys. Fluids \textbf{19}, 054109 (2007). 
548: 
549: % \bibitem{Berhanu07} M. Berhanu, R. Monchaux, S. Fauve, N. Mordant,
550: %   F. P\'etr\'elis, A. Chiffaudel, F. Daviaud, B. Dubrulle,
551: %   L. Mari\'e, F. Ravelet, M. Bourgoin, Ph. Odier, J.-F. Pinton, and
552: %   R. Volk ``Magnetic field reversals in an experimental turbulent
553: %   dynamo'', Eur. Phys. Lett. \textbf{77} 59001 (2007).
554: 
555: \bibitem{Marie06} L. Mari\'{e}, C. Normand and F. Daviaud,
556: %``Galerkin analysis of kinematic dynamos in the von K\'{a}rm\'{a}n geometry'',
557: Phys. Fluids \textbf{18}, 017102 (2006).
558: 
559: \bibitem{Laguerre_thesis} R.~Laguerre,
560:   % ``Approximation des \'equations de la MHD par une m\'ethode
561:   % hybride spectrale-\'el\'ements finis nodaux~: application \`a
562:   % l'effet dynamo'',
563: PhD thesis, Universit\'e Paris VII (2006).
564: 
565: %\bibitem{SKR66} M. Steenbeck et al.,
566: %  Monats. Dt. Akad. Wiss. \textbf{9}, 716 (1967).
567: 
568: %\bibitem{Muller01} R. Stieglitz and U. M\"uller, Phys. Fluids
569: %  \textbf{13}, 561 (2001).
570: 
571: 
572: %\bibitem{BL}
573:   % Dynamo Models of the Solar Cycle,
574: %  P. Charbonneau, Living Reviews in Solar Physics,\textbf{2.2} (2005),
575: %  http://www.livingreviews.org/lrsp-2005-2.
576:   
577: %\bibitem{Parker93} E.N. Parker,
578:   % A solar dynamo surface wave at the interface between convection
579:   % and nonuniform rotation,
580: %Astrophys. J. \textbf{408}, 707 (1993).
581: 
582: 
583: \end{thebibliography}
584: \end{document}
585: 
586: %%% Local Variables:
587: %%% mode: latex
588: %%% mode: flyspell
589: %%% TeX-master: t
590: %%% End: