1: \documentclass{jpp}
2: \usepackage{subeqn}
3: \usepackage{upmath}
4: \usepackage{amssymb}
5: \usepackage{amsbsy}
6:
7: \usepackage{graphicx}% Include figure files
8: \usepackage{color} % Color
9:
10: % BIBLIOGRAPHY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: \usepackage{natbib}
12: \bibliographystyle{jpp}
13: \def\newblock{\hskip .11em plus .33em minus .07em}
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15:
16: % COLOR %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17: \def\note#1{{\textcolor{red}{\bf [#1]}}} % note
18: \def\opt#1{{\textcolor{green}{#1}}} % optional
19: \def\add#1{{\textcolor{blue}{#1}}} % addition
20: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
21:
22: \title[Energy transfer in Hall-MHD turbulence]
23: {Energy transfer in Hall-MHD turbulence: \\
24: cascades, backscatter, and dynamo action}
25: \author[P.\,D. Mininni, A. Alexakis and A. Pouquet]%
26: {P\ls A\ls B\ls L\ls O\ns D.\ns M\ls I\ls N\ls I\ls N\ls N\ls I%
27: \thanks{e-mail: mininni@ucar.edu},\\
28: A\ls L\ls E\ls X\ls A\ls N\ls D\ls R\ls O\ls S\ns
29: A\ls L\ls E\ls X\ls A\ls K\ls I\ls S%
30: \thanks{e-mail: alexakis@ucar.edu}\ns \and \ns
31: A\ls N\ls N\ls I\ls C\ls K\ns P\ls O\ls U\ls Q\ls U\ls E\ls T%
32: \thanks{e-mail: pouquet@ucar.edu}}
33: \affiliation{National Center for Atmospheric Research,
34: P.O. Box 3000, Boulder, Colorado 80307, USA}
35:
36: \pubyear{2005}
37: \volume{0}
38: \part{0}
39: \pagerange{1--25}
40: \date{\today}
41: \setcounter{page}{1}
42:
43: \begin{document}
44:
45: \maketitle
46:
47: \begin{abstract}
48: Scale interactions in Hall MHD are studied using both the mean field theory
49: derivation of transport coefficients, and direct numerical simulations in
50: three space dimensions. In the magnetically dominated regime, the eddy
51: resistivity is found to be negative definite, leading to large scale
52: instabilities. A direct cascade of the total energy is observed, although
53: as the amplitude of the Hall effect is increased, backscatter of magnetic
54: energy to large scales is found, a feature not present in MHD flows.
55: The coupling between the magnetic and velocity fields is different
56: than in the MHD case, and backscatter of energy from small scale magnetic
57: fields to large scale flows is also observed. For the magnetic helicity, a
58: strong quenching of its transfer is found. We also discuss non-helical
59: magnetically forced Hall-MHD simulations where growth of a large scale
60: magnetic field is observed.
61: \end{abstract}
62:
63: \section{ \label{Intro} Introduction }
64:
65: The relevance of two fluid effects has recently been pointed out in
66: several studies of astrophysical and laboratory plasmas
67: \citep{Balbus01,Sano02,Mirnov03,Ding04}. The effect of adding the Hall
68: current to the dynamics of the flow was studied in several scenarios,
69: particularly dynamo action
70: \citep{Helmis68,Galanti95,Mininni02,Mininni03b,Mininni05a} and
71: reconnection \citep{Birn01,Shay01,Wang01,Morales05}. Several of these
72: works showed that the Hall currents increase the reconnection rate
73: of magnetic field lines. However, most of the studies of magnetic
74: reconnection were done for particular configurations of current sheets.
75: It was shown in particular by \citet{Smith04} that when a turbulent
76: background is present the reconnection rate is dominated by the
77: amplitude of the turbulent fluctuations. The process of
78: magnetic reconnection is relevant in several astrophysical and
79: geophysical scenarios, such as the magnetopause, the magnetotail, the
80: solar atmosphere, or the interplanetary and interstellar medium.
81: Reconnection can also play a role in the generation of large
82: scale magnetic fields by dynamo action \cite{Zeldovich}.
83:
84: Some of the works in Hall-magnetohydrodynamics (Hall-MHD) present
85: conflicting results, indicating in some cases that the Hall effect can
86: help the growth of a large scale magnetic field \citep{Mininni05a} or a
87: large scale self-organization process \citep{Mahajan98,Numata04,Ohsaki05},
88: while in other cases the Hall currents were observed to generate small
89: scales and filamentation \citep{Laveder02a,Laveder02b,Rheinhardt02}.
90:
91: As a result, it becomes of interest to study the physical processes
92: leading to cascades and transfer of ideal invariants in three-dimensional
93: Hall-MHD turbulence. Phenomena observed in the laboratory and space
94: plasmas tend to show an intermittent or impulsive behavior
95: \citep{Bhattacharjee99} characteristic of turbulent flows. The relevance
96: of Hall-MHD turbulence in the solar wind was shown by \citet{Ghosh96}.
97: Also, Hall-MHD turbulence can play a crucial role in the transfer of
98: matter in the magnetopause as was pointed by \citet{Rezeau01}.
99:
100: In this work, we study both analytically and numerically three dimensional
101: Hall-MHD turbulence as the result of a dynamo process, and from a purely
102: electromotive forcing. Detailed studies of shell-to-shell energy transfer
103: from direct numerical simulations (DNS) have been done for hydrodynamic
104: \citep{Domaradzki90,Ohkitani92,Zhou93,Yeung95,Alexakis05b}
105: and magnetohydrodynamic flows \citep{Debliquy05,Alexakis05a,Mininni05b}.
106: To the best of our knowledge, the energy transfer in Hall-MHD
107: turbulence has not been studied before.
108:
109: We show evidence of non-locality of the transfer in Fourier space,
110: and that the Hall effect can increase both the transfer of magnetic
111: energy to smaller scales (locally), as well as give a novel non-local
112: backscatter of magnetic energy to large scales. These results become
113: clear when examining the modification to the turbulent magnetic
114: diffusivity due to the Hall term. Also, we observe that the Hall
115: currents impact on the coupling between the magnetic and velocity fields.
116: The transfer of energy between these two fields is different than in the
117: MHD case. The Hall-MHD equations also display a backscatter of energy
118: from small scale magnetic fluctuations to the large scale flows themselves.
119: The transfer of helicity is briefly discussed as well and observed to be
120: quenched by the Hall effect.
121:
122: The structure of the paper is as follows. In Sec. \ref{Equations} we
123: introduce the Hall-MHD equations and we define the various transfer
124: terms. In Sec. \ref{Transport} we derive turbulent transport
125: coefficients for the Hall-MHD induction equation. In Sec.
126: \ref{Simulations} we briefly discuss the code and details of the
127: mechanically forced numerical simulations for completeness.
128: Section \ref{Transfers} presents the
129: transfer terms in Hall-MHD as obtained from the numerical simulations.
130: Section \ref{Upscaling} shows backscatter of magnetic energy in
131: non-helical Hall-MHD magnetically forced simulations. Finally, Sec.
132: \ref{Discussion} summarizes the results and discusses implications of
133: our work for the understanding of turbulence, dynamo action, and
134: reconnection in Hall-MHD.
135:
136: \section{ \label{Equations} The Hall-MHD equations and transfer terms }
137:
138: In dimensionless Alfv\'enic units, the Hall-MHD equations are
139: \begin{equation}
140: \partial_t {\bf U} + {\bf U}\cdot \nabla {\bf U} = - \nabla {\cal P} +
141: {\bf B}\cdot \nabla {\bf B} + \nu \nabla^2 {\bf U} +{\bf f} ,
142: \label{eq:momentum}
143: \end{equation}
144: %
145: \begin{equation}
146: \partial_t {\bf B} = \nabla \times \left [ \left( {\bf U} - \epsilon
147: {\bf J} \right) \times {\bf B} \right] + \eta \nabla^2 {\bf B} ,
148: \label{eq:induction1}
149: \end{equation}
150: %
151: where ${\bf U}$ is the bulk velocity field, ${\bf B}$ is the magnetic
152: field, ${\bf J} = \nabla \times {\bf B}$ is the current density,
153: ${\cal P}$ is the pressure, $\nu$ is the kinematic viscosity, and
154: $\eta$ is the magnetic diffusivity. From the Maxwell equations and
155: incompressibility of the flow,
156: $\nabla \cdot {\bf U} = \nabla \cdot {\bf B} = 0$.
157:
158: The Hall term $\epsilon {\bf J} \times {\bf B}$ in Eq. (\ref{eq:induction1})
159: measures the velocity difference between species, where the electron
160: velocity is ${\bf U}^e = {\bf U} - \epsilon {\bf J}$. Here, $ \epsilon $
161: measures the relative strength of the Hall effect, with the Hall term being
162: dominant for wavenumbers larger than $k_{Hall} \sim 1/\epsilon$ if
163: equipartition between the fields is assumed. The measure of strength
164: of the Hall effect can be written as $\epsilon = L_{Hall}/L_0$ where
165: $L_0$ is a characteristic length (we will use $L_0=2\pi$, the size of
166: the box in our simulations). In terms of physical parameters, and for
167: a fully ionized plasma, the Hall length is $L_{Hall}=cU_A/(\omega_{pi}U_0)$,
168: where $U_A$ is the Alfv\'enic speed, $U_0$ is a characteristic speed,
169: $c$ is the speed of light, and $\omega_{pi}$ is the ion plasma
170: frequency (when $U_0=U_A$, $L_{Hall}$ reduces to the ion skin depth).
171: In a partially ionized plasma, expressions for $L_{Hall}$ can be found
172: in \citet{Sano02} and \citet{Mininni03b}.
173:
174: Of special interest is the ratio between the integral length $L$, the
175: Hall length $L_{Hall}$, and the Ohmic dissipation length $L_\eta$. For
176: $L_{Hall} \ll L_\eta$ ($\epsilon \to 0$), the Hall-MHD equations reduce
177: to the well known MHD case. In several astrophysical problems, such as
178: accretion disks, protoplanetary disks, or the magnetopause
179: \citep[see e.g.][]{Birn01,Balbus01,Sano02}, the Hall scale is larger than
180: Ohmic scales although smaller than the integral scale $L$ of the flow.
181: We will be interested in this regime in this work, although we remark
182: that the separation between these scales in astrophysical or geophysical
183: problems is far from what can be achieved in numerical simulations.
184:
185: The Hall-MHD equations have three ideal invariants \citep{Turner86}. In
186: this work we will focus on two invariants, the total energy
187: \begin{equation}
188: E = \frac{1}{2} \int (U^2 + B^2) \, d{\bf x}^3 ,
189: \end{equation}
190: and the magnetic helicity
191: \begin{equation}
192: H = \frac{1}{2} \int {\bf A} \cdot {\bf B} \, d{\bf x}^3 ,
193: \end{equation}
194: where {\bf A} is the vector potential, $\nabla \times {\bf A} = {\bf B}$.
195: These quantities are also ideal invariants of the MHD equations
196: ($\epsilon=0$). The third MHD invariant, the cross helicity, is replaced
197: in Hall-MHD by the hybrid helicity \citep{Turner86}
198: and is small in the simulations we will discuss.
199:
200: The expressions we will use for the shell-to-shell energy transfers have
201: been derived for the MHD case by \citet{Verma04,Debliquy05}; and
202: \citet{Alexakis05a}. Here we present the derivation of
203: the transfer terms for the Hall-MHD equations. Equation
204: (\ref{eq:induction1}) can be rewritten as
205: \begin{equation}
206: \partial_t {\bf B} + {\bf U}\cdot \nabla {\bf B} = {\bf B}\cdot \nabla
207: {\bf U} - \epsilon \nabla \times \left( {\bf J} \times {\bf B} \right)
208: + \eta \nabla^2 {\bf B} .
209: \label{eq:induction}
210: \end{equation}
211: We introduce a filter in shells in Fourier space, such as ${\bf F}_K$ which
212: denotes the components of the field with wavenumbers between $K$ and
213: $K+1$ [i.e.
214: $ {\bf F}_K ({\bf x})= \sum_{k=K}^{K+1} \hat{\bf F}({\bf k}) e^{i {\bf k}
215: \cdot {\bf x}} $], from Eqs. (\ref{eq:momentum}) and (\ref{eq:induction})
216: we can write detailed balance equations for the energy,
217: {\setlength\arraycolsep{2pt}
218: \begin{eqnarray}
219: \partial_t E_U(K) &=& \int \left\{ \sum_Q \left[ - {\bf U}_K
220: {\bf \cdot (U \cdot \nabla) \cdot U}_Q + {\bf U}_K
221: {\bf \cdot (B \cdot \nabla) \cdot B}_Q \right] - \right. {}
222: \nonumber \\
223: &&{} - \nu \nabla^2 {\bf U}_K + {\bf f \cdot U}_K \Bigg\} \; d{\bf x}^3 \, ,
224: \end{eqnarray}}
225: %
226: {\setlength\arraycolsep{2pt}
227: \begin{eqnarray}
228: \partial_t E_B(K) &=& \int \left\{ \sum_Q \left[ - {\bf B}_K
229: {\bf \cdot (U \cdot \nabla) \cdot B}_Q + {\bf B}_K
230: {\bf \cdot (B \cdot \nabla) \cdot U}_Q + {} \right. \right.
231: \nonumber \\
232: &&{} + \epsilon {\bf J}_K \cdot ({\bf B} \times {\bf J}_Q) \Big]
233: - \eta \nabla^2 {\bf B_K} \Bigg\} \; d{\bf x}^3 \, .
234: \end{eqnarray}}
235: %
236: Here, $E_U(K)$ and $E_B(K)$ denote respectively the kinetic and magnetic
237: energy in the shell $K$. The above equations can be written in the more
238: compact form:
239: \begin{equation}
240: \partial_t E_U(K) = \sum_Q [{\mathcal T}_{UU}(K,Q)+{\mathcal T}_{BU}(K,Q)]
241: - \nu {\mathcal D}_U(K) + {\mathcal F}(K) ,
242: \label{eq:Eu}
243: \end{equation}
244: %
245: \begin{equation}
246: \partial_t E_B(K) =
247: \sum_Q [{\mathcal T}_{UB}(K,Q)+{\mathcal T}_{BB}(K,Q)] -
248: \eta {\mathcal D}_B(K).
249: \label{eq:Eb}
250: \end{equation}
251: %
252: The functions ${\mathcal T}_{UU}(K,Q)$,
253: ${\mathcal T}_{UB}(K,Q)$, ${\mathcal T}_{BB}(K,Q)$, and
254: ${\mathcal T}_{BU}(K,Q)$
255: express the energy transfer between different fields and shells,
256: \begin{equation}
257: {\mathcal T}_{UU}(K,Q) \equiv
258: -\int {{\bf U}_K ({\bf U} \cdot \nabla) {\bf U}_Q} \; d{\bf x}^3 ,
259: \label{eq:TUU}
260: \end{equation}
261: %
262: \begin{equation}
263: {\mathcal T}_{UB}(K,Q) \equiv
264: \int {{\bf U}_K ({\bf B} \cdot \nabla) {\bf B}_Q} \; d{\bf x}^3 ,
265: \label{eq:TUB}
266: \end{equation}
267: %
268: \begin{equation}
269: {\mathcal T}_{BU}(K,Q) \equiv
270: \int {{\bf B}_K ({\bf B} \cdot \nabla) {\bf U}_Q} \; d{\bf x}^3 .
271: \label{eq:TBU}
272: \end{equation}
273: %
274: In general, for positive transfer, the first subindex denotes the
275: field that receives energy, the second subindex the field that gives
276: energy. The first wavenumber corresponds to the field receiving
277: energy, and the second wavenumber to the field giving energy. As
278: an example, positive ${\mathcal T}_{UU}(K,Q)$ represents energy
279: transfered from the velocity field at the shell $K$ to velocity
280: field at the shell $Q$. In the same way, positive $T_{UB}(K,Q)$
281: represents energy transfered from the magnetic field at wavenumbers
282: $K$ to the velocity field at wavenumbers $Q$.
283:
284: The transfer of magnetic to magnetic energy ${\mathcal T}_{BB}(K,Q)$
285: in Hall-MHD consists of two terms
286: \begin{equation}
287: {\mathcal T}_{BB}(K,Q) = {\mathcal T}_{BB}^\textrm{MHD}(K,Q) +
288: {\mathcal T}_{BB}^\textrm{Hall}(K,Q)
289: \end{equation}
290: where
291: \begin{equation}
292: {\mathcal T}_{BB}^\textrm{MHD}(K,Q) \equiv
293: -\int {{\bf B}_K ({\bf U} \cdot \nabla) {\bf B}_Q } \; d{\bf x}^3,
294: \label{eq:TBBMHD}
295: \end{equation}
296: is the usual MHD transfer of magnetic energy through advection by the
297: bulk velocity field, and
298: \begin{equation}
299: {\mathcal T}_{BB}^\textrm{Hall}(K,Q) \equiv
300: \epsilon \int {{\bf J}_K \cdot ({\bf B} \times {\bf J}_Q)} \; d{\bf x}^3,
301: \label{eq:TBBHall}
302: \end{equation}
303: is the transfer of magnetic energy due to the Hall current. Note that
304: the definition of the transfer terms corresponds to the MHD case in
305: \citet{Alexakis05a}, except for the new term
306: ${\mathcal T}_{BB}^\textrm{Hall}(K,Q)$. However, as will be shown later,
307: the behavior of the rest of the transfer terms in Hall-MHD will also be
308: indirectly modified by the presence of the Hall effect.
309:
310: All these transfer functions satisfy the identity
311: \begin{equation}
312: \label{tran_id}
313: {\mathcal T}_{vw}(K,Q)=-{\mathcal T}_{wv}(K,Q) ,
314: \end{equation}
315: where $v,w$ can be either $U$ or $B$. This detailed conservation is
316: what allows us to define the terms as transfers of energy between shells.
317: Note that other groupings of the nonlinear terms in the Hall-MHD equations
318: would not satisfy this symmetry condition.
319:
320: In Eqs. (\ref{eq:Eu}-\ref{eq:Eb}) we also have two dissipation functions
321: and the energy injection rate
322: \begin{equation}
323: \nu {\mathcal D}_{U}(K) \equiv \nu \int |{\bf \nabla U}_K|^2 d{\bf x}^3 ,
324: \end{equation}
325: \begin{equation}
326: \eta {\mathcal D}_{B}(K) \equiv \eta \int |{\bf \nabla B}_K|^2 d{\bf x}^3 ,
327: \end{equation}
328: \begin{equation}
329: {\mathcal F}(K) \equiv \int { \bf f} \cdot {\bf U}_K \, d{\bf x}^3 .
330: \end{equation}
331:
332: Finally, we can also define the transfer of magnetic helicity. From Eq.
333: (\ref{eq:induction1}) we have
334: \begin{equation}
335: \partial_t H(K) = \sum_Q {\mathcal T}_H (K,Q) - \eta {\mathcal D}_H (K),
336: \end{equation}
337: where the transfer of magnetic helicity from the wavenumber $K$ to the
338: wavenumber $Q$ is given by
339: \begin{equation}
340: {\mathcal T}_H (K,Q) \equiv \int{ {\bf B}_K \cdot \left[\left({\bf U}
341: - \epsilon {\bf J}\right) \times {\bf B}_Q \right] d{\bf x}^3}.
342: \label{eq:TH}
343: \end{equation}
344: This transfer function satisfies the relation
345: ${\mathcal T}_H (K,Q)=-{\mathcal T}_H (K,Q)$.
346: The first term in Eq. (\ref{eq:TH}) proportional to
347: ${\bf U}$ is the usual transfer of $H_M$ in MHD, while the second term
348: proportional to $\epsilon {\bf J}$ is the contribution due to the Hall
349: effect. Note that as a whole, magnetic helicity is transfered between
350: the shells $K$ and $Q$ interacting with the electron velocity field
351: ${\bf U} - \epsilon {\bf J}$. This is in agreement with the fact that
352: in the ideal limit the magnetic field in the Hall-MHD system is frozen
353: to the electron velocity field, instead of the bulk velocity field of
354: the plasma as in MHD.
355:
356: The dissipation rate of magnetic helicity at the wavenumber $K$ is given by
357: \begin{equation}
358: \eta {\mathcal D}_H (K) \equiv \eta \int{{\bf B}_K \cdot {\bf J}_K d{\bf x}^3}.
359: \end{equation}
360: It is also worth noting that, since the magnetic helicity is not a positive
361: defined quantity contrary to the energy, the interpretation of its transfer
362: is more difficult. We will not attempt here a separation of its different sign
363: components \citep[see e.g.][]{Waleffe91,Chen03a,Chen03b} for the case of
364: kinetic helicity in hydrodynamic turbulence).
365:
366: \section{ \label{Transport} Transport coefficients }
367:
368: In \citet{Mininni02}, the expression of the $\alpha$ dynamo coefficient
369: was derived for Hall-MHD. Although an expression of the Hall-MHD turbulent
370: diffusivity was derived by \citet{Mininni03a}, the closure was
371: only valid for specific solutions of the Hall-MHD equations. To
372: interpret the results from the energy transfer, it will be useful to
373: have expressions for all the turbulent transport coefficients in the
374: induction equation. To this end, and for the sake of simplicity, we
375: will use mean field theory (MFT) \citep{Steenbeck66,Krause} and the
376: reduced smooth approximation (RSA) \citep{Blackman99}. RSA was introduced
377: to solve some ambiguities present in MFT when the magnetic field is
378: strong enough to affect the velocity field trough the Lorentz force.
379: Although there are still assumptions in MFT not completely justified,
380: at least a qualitative agreement has been observed with simulations
381: in the MHD case \citep{Brandenburg01} and the Hall-MHD case
382: \citep{Mininni03b,Mininni05a}. The transport coefficients can also be
383: derived using more elaborate closures, such as the Lagrangian History
384: Direct Interaction Approximation (LHDIA) or the Eddy Damped Quasi Normal
385: Markovian (EDQNM) closures \citep[see e.g.][]{Lesieur}. It is worth noting
386: that the analysis that follows in Sections \ref{Simulations} and
387: \ref{Transfers} is of general validity and independent of the assumptions
388: we will use here to derive the turbulent transport coefficients.
389:
390: We split the fields into
391: \begin{eqnarray}
392: {\bf U} &=& \overline{\bf U} + {\bf u} + {\bf u}_0 , \\
393: {\bf B} &=& \overline{\bf B} + {\bf b} + {\bf b}_0 ,
394: \end{eqnarray}
395: where ${\bf u}_0$ and ${\bf b}_0$ are isotropic and homogeneous
396: solutions of Eqs. (\ref{eq:momentum}) and (\ref{eq:induction}) in
397: the absence of the mean fields $\overline{\bf U}$ and $\overline{\bf B}$.
398: The fields with overbars are large scale fields, and ${\bf u}$ and ${\bf b}$
399: are small scale corrections to the isotropic and homogeneous solutions
400: due to the presence of the large scale fields. The fluctuating fields
401: satisfy
402: $\left< {\bf u} \right> = \left< {\bf u}_0 \right> =
403: \left< {\bf b} \right> = \left< {\bf b}_0 \right> = 0 $,
404: where the brackets denote an average that satisfies Taylor's
405: hypothesis \citep{Krause}. Replacing in Eq. (\ref{eq:induction}), using
406: the equations for the ${\bf u}_0$ and ${\bf b}_0$ fields, dropping terms
407: quadratic in the fluctuating fields ${\bf u}$ and ${\bf b}$, and averaging leads to
408: \begin{equation}
409: \partial_t \overline{\bf B} = \nabla \times \left[ \left( \overline{\bf U}
410: - \epsilon \overline{\bf J} \right) \times \overline{\bf B}
411: + \varepsilon \right] + \eta \nabla^2 \overline{\bf B} ,
412: \label{eq:largeB}
413: \end{equation}
414: where $\varepsilon$ is the mean field electromotive force
415: \begin{equation}
416: \varepsilon = \left< {\bf u}_0^e \times {\bf b} + {\bf u}^e \times
417: {\bf b}_0 \right> .
418: \label{eq:emfepsilon}
419: \end{equation}
420: Our main aim in this section is to close Eq. (\ref{eq:largeB}) and
421: write $\varepsilon$ only as a function of averages of the fields
422: ${\bf u}_0$, ${\bf b}_0$, and spatial derivatives of $\overline{\bf B}$.
423: A simple argument of symmetry shows that in the approximately isotropic
424: case
425: \begin{equation}
426: \varepsilon = \alpha \overline{\bf B} - \beta \nabla \times \overline{\bf B}
427: + \gamma \nabla \times \nabla \times \overline{\bf B} .
428: \label{eq:emf}
429: \end{equation}
430:
431: From Eqs. (\ref{eq:momentum}) and (\ref{eq:induction}), and subtracting
432: the equations for the mean flows, we can also write equations for the
433: evolution of the turbulent fluctuations ${\bf u}$ and ${\bf b}$. We drop
434: terms quadratic in ${\bf u}$ and ${\bf b}$, and keep only terms to
435: zeroth and linear order in $\overline{\bf B}$,
436: \begin{equation}
437: \partial_t {\bf b} = \nabla \times \left( \overline{\bf U} \times {\bf b}_0
438: + {\bf u}_0^e \times \overline{\bf B} + \epsilon {\bf b}_0^e \times
439: \overline{\bf J} + {\bf u}^e \times {\bf b}_0 +
440: {\bf u}^e_0 \times {\bf b} - \varepsilon \right) +
441: \eta \nabla^2 {\bf b} .
442: \label{eq:smallb}
443: \end{equation}
444: In this equation, $\varepsilon$ involves averaged quantities, and
445: from the Taylor's hypothesis it gives no contribution to the mean
446: electromotive force. The fourth and fifth terms on the {\it r.h.s.} can be
447: dropped using RSA, namely that $|{\bf u}| , |{\bf b}| \ll |\overline{\bf B}|$
448: (note that this condition is less stringent than the usual assumptions
449: in MFT, since the amplitude of the fields ${\bf u}_0$, ${\bf b}_0$ can
450: be much larger than the amplitude of the mean magnetic field). We will
451: assume the viscosity and diffusivity are small, and as a result we will
452: also drop the last term on the right hand side of Eq. (\ref{eq:smallb}).
453: Terms proportional to $\overline{\bf U}$ can be removed in the proper
454: frame of reference. Finally we obtain
455: \begin{equation}
456: \partial_t {\bf b} \approx \nabla \times \left( {\bf u}_0^e \times
457: \overline{\bf B} + \epsilon {\bf b}_0^e \times \overline{\bf J} \right) .
458: \label{eq:RSAb}
459: \end{equation}
460: The second term on the right hand side of Eq. (\ref{eq:RSAb}) involves
461: only spatial derivatives of $\overline{\bf B}$ and gives no contribution
462: to the $\alpha$ coefficient, but is retained here since it will give
463: contributions to $\beta$ and $\gamma$.
464:
465: Following the same steps, we can also write an equation for the evolution
466: of ${\bf u}$,
467: \begin{equation}
468: \partial_t {\bf u} \approx \overline{\bf B} \cdot \nabla {\bf b}_0
469: + {\bf b}_0 \cdot \nabla \overline{\bf B} - \nabla p .
470: \label{eq:smallu}
471: \end{equation}
472:
473: To obtain the mean field electromotive force we replace time derivatives
474: in Eqs. (\ref{eq:smallb}) and (\ref{eq:smallu}) by the inverse of a
475: correlation time $\tau$. This step, common in MFT, assumes the existence
476: of a finite correlation time. At present there is no evidence of its
477: validity in general
478: \citep[see e.g.][]{Gruzinov95,Blackman02,Brandenburg05}. Since we are
479: introducing a correlation time to close these equations, the expressions
480: obtained for the turbulent transport coefficients will be considered as
481: symbolic expressions.
482:
483: Before replacing the expression for ${\bf u}$ in
484: Eq. (\ref{eq:emf}), Eq. (\ref{eq:smallu}) has to be solved for the small
485: scale pressure $p$. We will use a technique developed by \citet{Gruzinov95}
486: \citep[see also][]{Blackman02}. The $\alpha$ effect is linear in
487: $\overline{\bf B}$ and therefore the correct result can be obtained
488: assuming $\overline{\bf B}$ uniform. Then
489: $\partial_t {\bf u} \approx \overline{\bf B} \cdot \nabla {\bf b}_0 $.
490: Replacing the time derivative by $\tau^{-1}$ and replacing the expression
491: in Eq. (\ref{eq:emf}), we obtain in the weak isotropic case
492: \begin{equation}
493: \alpha = \frac{\tau}{3} \left< -{\bf u}_0^e \cdot \nabla \times
494: {\bf u}_0^e + {\bf b}_0 \cdot \nabla \times {\bf b}_0 -
495: \epsilon {\bf b}_0 \cdot \nabla \times \nabla \times
496: {\bf u}_0^e \right> .
497: \end{equation}
498:
499: To compute $\beta$ and $\gamma$ we have to keep spatial derivatives of
500: $\overline{\bf B}$ in Eq. (\ref{eq:smallu}), and therefore we have to
501: solve for the pressure. This was done by \citet{Gruzinov95} transforming
502: Eq. (\ref{eq:smallu}) to Fourier space, and doing a Taylor expansion of
503: the projector operator for incompressible ${\bf u}$ assuming a large
504: scale separation between the mean and fluctuating fields. In three spatial
505: dimensions, it was shown that the pressure and
506: ${\bf b}_0 \cdot \nabla \overline{\bf B}$ terms give no contribution to
507: $\varepsilon$ in Eq. (\ref{eq:emfepsilon}). As a result, we are
508: only left with the terms proportional to spatial derivatives of
509: $\overline{\bf B}$ when Eq. (\ref{eq:RSAb}) and the first term on the
510: r.h.s. of Eq. (\ref{eq:smallu}) are replaced on Eq. (\ref{eq:emfepsilon}).
511: Again, assuming weak isotropy, we obtain the expressions
512: for the remaining turbulent transport coefficients
513: {\setlength\arraycolsep{2pt}
514: \begin{eqnarray}
515: \beta &=& \frac{\tau}{3} \left< {{\bf u}_0^e}^2 + \epsilon \left( {\bf u}_0
516: \cdot \nabla \times {\bf b}_0^e + {\bf b}_0
517: \cdot \nabla \times {\bf u}_0^e \right) \right> , \\
518: \gamma &=& - \frac{\tau \epsilon}{3} \left< {\bf b}_0 \cdot {\bf u}_0^e
519: \right> .
520: \end{eqnarray}}
521: The two last terms in $\alpha$, and the third term in $\beta$ come from
522: the small scale momentum equation and are related with the backreaction
523: of the magnetic field into the velocity field. In the kinematic regime
524: of a dynamo, \
525: $\alpha = - \tau/3\left<{\bf u}_0^e\cdot\nabla\times {\bf u}_0^e\right>$,
526: which for $\epsilon=0$ reduces to the MHD case \citep{Krause}. The general
527: expression for $\epsilon=0$ reduces to the MHD expression first found
528: using the EDQNM closure by \citet{Pouquet76},
529: $\alpha = \tau/3\left<-{\bf u}_0^e\cdot\nabla\times {\bf u}_0^e
530: + {\bf b}_0\cdot\nabla\times {\bf b}_0 \right>$. Note also that in
531: the MHD case in three dimensions, the turbulent diffusivity
532: $\beta=-\tau \left< {{\bf u}_0}^2 \right>/3 $ is not changed during the
533: nonlinear saturation \citep{Gruzinov95}.
534:
535: The turbulent diffusivity $\beta$ in Hall-MHD is not positive definite,
536: contrary to the pure MHD case \citep[note that negative effective
537: diffusivities can be found in MHD if the assumption of homogeneity
538: is dropped, see e.g.][]{Lanotte99}. A negative value of $\beta$ represents
539: non-local transfer of energy from the small scale turbulent fields to the
540: large scale magnetic field. This result will be of interest in the
541: following sections.
542:
543: It is worth studying the values of $\beta$ for particular cases. If
544: $\epsilon$ is large enough and the system is magnetically dominated
545: ($E_B \gg E_U$), then
546: $\beta \approx - \tau \epsilon^2 \left< j_0^2 \right>/3$, where
547: ${\bf j}_0 = \nabla \times {\bf b}_0$ and we assumed the average is
548: a spatial average. In this case, $\beta$ is always negative
549: implying transfer of energy from the small scales to the large.
550:
551: The normal modes of the Hall-MHD equations are circular polarized
552: ($\nabla \times {\bf u}_0^e = \pm k {\bf u}_0^e$,
553: $\nabla \times {\bf b}_0 = \pm k {\bf b}_0$) and dispersive, and in the
554: limit $k\gg1$ they satisfy dispersion relations
555: $\omega \sim \epsilon k^2 \overline{B}$ (whistlers, right-handed polarized)
556: and $\omega \sim \overline{B}/ \epsilon$ (ion-cyclotron waves, left-handed
557: polarized). Also, for these waves the fields are related by
558: ${\bf b}_0 = -k \overline{B}/\omega {\bf u}_0^e$ \citep{Mininni05c}.
559: If we assume a background of waves,
560: $\beta \sim \tau/3 \left< {u_0^e}^2 \right> (1 \pm 2 \epsilon k^2
561: \overline{B} / \omega)$, which for $k$ large enough can give positive
562: or negative turbulent diffusivity according to the orientation of the wave.
563: Note that from the dispersion relations, at small scales whistlers give a
564: finite contribution to the turbulent diffusivity, while ion-cyclotron waves
565: give a much larger turbulent diffusivity that grows as $k^2$.
566:
567: \section{ \label{Simulations} Simulations }
568:
569: In this section we summarize the simulations that will be used to
570: compute the energy and helicity transfer functions defined in
571: Sec. \ref{Equations}. We performed three simulations in three
572: dimensions with periodic boundary conditions, using a pseudospectral
573: Hall-MHD code as described in \citet{Mininni03b,Mininni05a}. Runge-Kutta
574: of second order is used to evolve the system of Eqs. (\ref{eq:momentum})
575: and (\ref{eq:induction1}). To ensure the divergence-free condition for the
576: magnetic field, a curl is removed from Eq. (\ref{eq:induction1}) and the
577: equation for the vector potential is instead solved, with the Couloumb's
578: gauge $\nabla \cdot {\bf A} = 0$. The three simulations are done with a
579: spatial resolution of $N^3=256^3$ grid points. The $2/3$ dealiasing rule is
580: used, and as a result the maximum wavenumber resolved by the code is
581: $k_{max} = N/3 \approx 85$. The kinematic viscosity and magnetic
582: diffusivity are set to $\nu = \eta = 2 \times 10^{-3}$, and all the
583: simulations are well resolved, in the sense that the kinetic
584: [$k_\nu = (\left<\omega^2\right>/\nu^2)^{1/4}$] and magnetic
585: [$k_\eta = (\left<J^2\right>/\eta^2)^{1/4}$] dissipation wavenumbers
586: are smaller than $k_{max}$ at all times.
587:
588: \begin{figure}
589: \begin{center} \includegraphics[width=9cm]{JPPmininni01} \end{center}
590: \caption{\label{fig:ener} Kinetic (thick curves) and magnetic
591: energy (thin curves) as a function of time, for runs
592: with $\epsilon = 0$ (solid lines), $\epsilon = 0.05$ (dotted lines),
593: and $\epsilon = 0.1$ (dashed lines).}
594: \end{figure}
595:
596: In Hall-MHD, the Courant-Friedrichs-Levy (CFL) condition is more
597: stringent than for MHD for which, with equipartition of kinetic and
598: magnetic energy, the CFL condition for explicit time-stepping imposes
599: an upper boundary on the time step $\Delta t \lesssim \Delta x/U_A$
600: where $\Delta x$ is the spatial step. In Hall-MHD, the dispersive
601: nature of the whistlers impose
602: $\Delta t \lesssim {\Delta x}^2/(\epsilon U_A)$. As a result, smaller
603: time steps will be needed as $\epsilon$ is increased. Also, since the
604: time step decreases quadratically as the spatial resolution is linearly
605: increased, we cannot achieve now spatial resolutions higher than
606: $256^3$ because of these constraints.
607:
608: A helical forcing at $k_0=2$ given by an ABC flow
609: {\setlength\arraycolsep{2pt}
610: \begin{eqnarray}
611: {\bf f} &=& \left[C\sin(k_0z)+B\cos(k_0y)\right] \hat{x} +
612: \left[A\sin(k_0x)+C\cos(k_0z)\right] \hat{y} + {}
613: \nonumber \\
614: &&{} + \left[B\sin(k_0y)+A\cos(k_0x)\right] \hat{z} ,
615: \end{eqnarray}}
616: with $A=0.9$, $B=1$, and $C=1.1$ was applied in the momentum equation.
617: This election of the amplitude coefficients was done to ensure breaking
618: the symmetries of the ABC flow and ensuring a faster development of
619: turbulence \citep{Archontis03}. After a first hydrodynamic run made
620: to reach a turbulent steady state, a random and small magnetic field was
621: introduced at small scales. Initially the ratio of kinetic to magnetic
622: energy was $E_u/E_b \sim 10^{-3}$.
623:
624: \begin{figure}
625: \begin{center} \includegraphics[width=9cm]{JPPmininni02} \end{center}
626: \caption{\label{fig:spec} Spectra of kinetic energy (thick curves)
627: and magnetic energy (thin curves) as a function of time, for
628: $t=5$ (solid), 15 (dash-dotted), 30 (dotted), and 45 (dashed). (a)
629: $\epsilon = 0.05$, and (b) $\epsilon = 0.1$.}
630: \end{figure}
631:
632: The simulation was continued to see exponential growth of the magnetic
633: energy (in the following, we will refer to this stage as the kinematic
634: regime), and finally nonlinear saturation of the small scale magnetic
635: field (in the following, Hall-MHD turbulence). Three simulations were
636: done, with $\epsilon = 0$ (MHD), $\epsilon = 0.05$ (which corresponds to
637: $k_{Hall} \approx 20$), and $\epsilon = 0.1$ ($k_{Hall} \approx 10$).
638: Figure \ref{fig:ener} shows the time history of the kinetic and magnetic
639: energies for these three runs.
640:
641: After $t \approx 20$, the small scale magnetic fields have reached
642: saturation for all values of $\epsilon$, while the large scale magnetic
643: field keeps growing slowly. As $\epsilon$ is increased, the magnetic
644: energy reached by the system after the non-linear saturation of the
645: small scales increases. However, this behavior is not monotonical in
646: $\epsilon$ as shown by \citet{Mininni03b}.
647:
648: \begin{figure}
649: \begin{center} \includegraphics[width=9cm]{JPPmininni03} \end{center}
650: \caption{\label{fig:heli} Absolute value of the magnetic helicity as a
651: function of time. Labels are as in Fig. \ref{fig:ener}. For
652: $\epsilon=0.1$ the magnetic helicity changes sign from negative
653: to positive at $t\approx24$.}
654: \end{figure}
655:
656: The saturation of the large scale magnetic field takes place in a longer
657: time \citep{Brandenburg01,Mininni05a}. Note that one of the biggest
658: challenges for DNS is to attain scale separation between the
659: different dynamical ranges that must be resolved. Reynolds numbers in
660: simulations are much smaller than the values observed in astrophysics
661: and geophysics. Moreover, compared with hydrodynamics and MHD, the extra
662: characteristic length scale in Hall-MHD (the Hall scale) makes it even
663: harder to achieve a proper separation between all these scales. As a
664: result, we will focus in this work in the energy transfer at scales
665: smaller than $k_0$ (the energy injection band), and the late time
666: large-scale evolution of these runs will not be discussed here
667: \citep[more details can be found e.g. in][]{Mininni05a}.
668:
669: Figure \ref{fig:spec} shows the time evolution of the magnetic and
670: kinetic energy spectra, for the runs with $\epsilon = 0.05$ and
671: $\epsilon = 0.1$. As previously mentioned ($t \approx 40$) the
672: spectrum of energy at scales smaller than $k_0$ has saturated and
673: reached a steady state, while the magnetic energy at $k=1$ keeps
674: growing slowly. Note that the ratio of kinetic to magnetic energy at
675: small scales in the saturated state depends on the value of $\epsilon$.
676:
677: Another quantity that will be of interest in the next section is the
678: magnetic helicity. Figure \ref{fig:heli} shows the time history of
679: the absolute value of magnetic helicity for the three runs. Note that
680: while in the MHD run ($\epsilon = 0$) the magnetic helicity grows
681: monotonically with time, in the Hall-MHD runs the time evolution is
682: strongly modified. For $\epsilon = 0.05$ the magnetic helicity grows
683: slower than in the MHD case, and for $\epsilon = 0.1$ it changes sign
684: at $t \approx 24$. As was observed by \citet{Mininni03b}, the Hall
685: effect inhibits the generation of net magnetic helicity at large scales
686: by the helical dynamo process. This inhibition grows monotonically with
687: the amplitude of Hall term, and for values of $\epsilon$ large enough
688: the magnetic helicity fluctuates around zero. The reason for this
689: behavior will be discussed in the next section.
690:
691: \section{ \label{Transfers} Transfers }
692:
693: In this section we discuss the energy transfer terms defined in
694: Sec. \ref{Equations} as obtained from the three DNS discussed in
695: the previous section.
696:
697: \subsection{The run with $\epsilon = 0.1$}
698:
699: \begin{figure}
700: \begin{center} \includegraphics[width=9cm]{JPPmininni04} \end{center}
701: \caption{\label{fig:T1uu} Transfer of kinetic energy from $Q=20$ to $K$,
702: ${\mathcal T}_{UU}(K,Q=20)$, in the kinematic regime (solid line),
703: at $t=26$ (dotted line), and at $t=45$ (dashed line) in the run with
704: $\epsilon = 0.1$.}
705: \end{figure}
706:
707: We start discussing in detail the transfer in the Hall-MHD run with
708: $\epsilon =0.1$. At late times in this simulation, when the system
709: is close to equipartition ($U_A \approx U$), the Hall wavenumber is
710: $k_{Hall} \approx 10$. Since we consider transfer functions at different
711: times, for the sake of comparison and unless explicitly said, all
712: transfers in this subsection will be normalized using the {\it r.m.s.}
713: velocity and magnetic field according to their expressions
714: [Eqs. (\ref{eq:TUU}-\ref{eq:TBBHall})]. Note that since $\epsilon {\bf J}$
715: has units of velocity (and ${\bf U}-\epsilon {\bf J}$ is the electron
716: velocity), the transfer function $T_{BB}^\textrm{Hall}$ is normalized
717: using $\left<B^2 |{\bf U}|\right>$. This election also allows for a
718: direct comparison of this term against $T_{BB}^\textrm{MHD}$ [see Eq.
719: (\ref{eq:TBBMHD})].
720:
721: Figure \ref{fig:T1uu} shows the transfer of kinetic energy from
722: the shell $Q=20$ to kinetic energy in shells $K$ for three different
723: times. As previously mentioned, positive transfer denotes energy
724: given by the shell $Q$, while negative transfer corresponds to
725: energy received by this shell. In this case, kinetic energy
726: in the shell $Q=20$ is mostly received from the shell $K=18$
727: (negative peak), and given to $K=22$ (positive peak). This
728: function represents the local and direct transfer of kinetic energy
729: to small scales. There are no noticeable differences in this transfer
730: between the Hall-MHD runs ($\epsilon = 0.05$ and $0.1$), and the MHD
731: run ($\epsilon = 0$).
732:
733: The curve for early times (kinematic regime) corresponds to the
734: initial exponential growth of magnetic energy, and is a time average
735: properly normalized. As time evolves and the magnetic energy grows,
736: the amount of kinetic energy transfered to small scales diminishes,
737: since a larger amount of kinetic energy at large scales is turned
738: into magnetic energy. This effect was previously observed in MHD runs
739: \citep{Mininni05b}.
740:
741: \begin{figure}
742: \begin{center} \includegraphics[width=9cm]{JPPmininni05} \end{center}
743: \caption{\label{fig:T1bbM} ${\mathcal T}_{BB}^\textrm{MHD}(K,Q=20)$
744: in the kinematic regime (solid line), at $t=26$ (dotted line), and at
745: $t=45$ (dashed line) in the run with $\epsilon = 0.1$.}
746: \end{figure}
747:
748: Figure \ref{fig:T1bbM} shows
749: ${\mathcal T}_{BB}^\textrm{MHD}(K,Q=20)$, the transfer of magnetic
750: energy at the shell $Q=20$ to magnetic energy in shells $K$ due to the
751: advection by the bulk velocity field. As in the case of
752: ${\mathcal T}_{UU}$, the transfer is local and the shell $Q=20$
753: receives most of the energy from $K=18$ (negative peak) and gives
754: energy to the shell $K=22$ (positive peak). Again, no significant
755: differences are observed between the three runs with different values
756: of $\epsilon$, except that this transfer, in amplitude, gets
757: substantially stronger as $\epsilon$ (and ${\bf B}$) increases.
758:
759: \begin{figure}
760: \begin{center} \includegraphics[width=9cm]{JPPmininni06} \end{center}
761: \caption{\label{fig:T1bbh} ${\mathcal T}_{BB}^\textrm{Hall}(K,Q=20)$ in
762: the kinematic regime (solid line), at $t=26$ (dotted line), and at
763: $t=45$ (dashed line) in the run with $\epsilon = 0.1$.}
764: \end{figure}
765:
766: \begin{figure}
767: \begin{center} \includegraphics[width=9cm]{JPPmininni07} \end{center}
768: \caption{\label{fig:T1buk} ${\mathcal T}_{BU}(K=20,Q)$ in the kinematic
769: regime (solid line), at $t=26$ (dotted line), and at $t=45$ (dashed
770: line) in the run with $\epsilon = 0.1$.
771: The inset shows a blow up of the last transfer.}
772: \end{figure}
773:
774: The total shell-to-shell transfer of magnetic energy is given by
775: ${\mathcal T}_{BB}^\textrm{MHD}$ plus
776: ${\mathcal T}_{BB}^\textrm{Hall}$. Figure \ref{fig:T1bbh} shows the
777: ${\mathcal T}_{BB}^\textrm{Hall}(K,Q)$ transfer at $Q=20$. As in the
778: previous cases, positive transfer denotes energy is given from the
779: shell $Q=20$ to shells $K$, while negative transfer indicates the
780: shell $Q$ receives energy from $K$. The
781: ${\mathcal T}_{BB}^\textrm{Hall}$ transfer is small during the kinematic
782: regime, but grows as the small scales reach nonlinear saturation.
783: Although this transfer is noisier than the previous terms studied,
784: two regions can be identified at late times. Around $Q=K=20$, the
785: transfer is local and direct: positive and negative peaks can be
786: observed at $K \approx 25$ and $K \approx 15$, indicating
787: energy is received and given respectively by the shell $Q$ from
788: and to these wavenumbers. On the other hand, at large scales
789: (up to $K \approx 10$) a region with positive transfer can also be
790: identified. This region indicates a non-local and inverse transfer
791: of energy: the shells with $K$ between 1 and 10 receive magnetic
792: energy from the shell $Q=20$. This combination of a local direct
793: transfer of energy and a non-local inverse transfer is characteristic
794: of the Hall term, and is in qualitative agreement with the turbulent
795: dissipation derived in Sec. \ref{Transport} where it was shown that it
796: can take negative values.
797:
798: \begin{figure}
799: \begin{center} \includegraphics[width=9cm]{JPPmininni08} \end{center}
800: \caption{\label{fig:T1buq} ${\mathcal T}_{BU}(K,Q=20)$ at three
801: different times: the kinematic regime (solid line), $t=26$ (dotted
802: line), and $t=45$ (dashed line) for the run with $\epsilon = 0.1$.}
803: \end{figure}
804:
805: The remaining transfer term is ${\mathcal T}_{BU}(K,Q)$, which
806: when positive represents transfer of kinetic energy from the shell
807: $Q$ to magnetic energy in the shell $K$. Although the expression
808: of this transfer function is equal for MHD and Hall-MHD, the
809: transfer is modified by the Hall currents. The reason for this can
810: be explained in two ways. On the one hand, the expression of the
811: $\alpha$-effect in Sec. \ref{Transport} is modified by the Hall term,
812: and this term represents transfer of energy from the turbulent
813: velocity field to the mean magnetic field. On the other hand,
814: waves are expected to give non-local coupling between the velocity
815: and magnetic fields \citep[see e.g.][in MHD]{Iroshnikov63,Kraichnan65}.
816: In Hall-MHD, the non-dispersive Alfv\'en waves of MHD are replaced by
817: dispersive circularly polarized waves and as a result, the coupling
818: between the two fields should also be modified.
819:
820: Figure \ref{fig:T1buk} shows ${\mathcal T}_{BU}(K=20,Q)$, the
821: energy transfered to the magnetic field at $K=20$ from the velocity
822: field at shells $Q$. In the kinematic regime this transfer is non-local
823: and similar to the MHD transfer \citep{Alexakis05a,Mininni05b}: the
824: magnetic field at $K=20$ receives energy from the large scale flow at
825: $Q=3$ and from all turbulent scales up to $Q \approx 20$. However, at
826: late times the transfer is strongly modified. The magnetic field at
827: $K=20$ still receives energy from a broad range of wavenumbers $Q$
828: smaller than $K$ \citep[as was found in][]{Alexakis05a},
829: but it also receives energy from larger wavenumbers
830: ($Q \approx 22$), and gives energy to the velocity field at slightly
831: smaller wavenumbers ($Q \approx 18$). Note that this indicates that in
832: Hall-MHD a magnetic field at a given scale can give rise to velocity
833: fluctuations at larger scales, a process studied by \citet{Mahajan05b}
834: and referred there as the {\it reverse dynamo}.
835:
836: \begin{figure*}
837: \begin{center} \includegraphics[width=12cm]{JPPmininni09} \end{center}
838: \caption{\label{fig:THall} ${\mathcal T}_{BB}^\textrm{Hall}(K,Q)$
839: (left column), and ${\mathcal T}_{BU}(K,Q)$ (right column) at three
840: different times: the kinematic regime (top), $t=26$ (middle), and
841: $t=45$ (bottom) for the run with $\epsilon = 0.1$. In all figures,
842: $K$ is on the $x$ axis and $Q$ in the $y$ axis. Shading go from dark
843: (${\mathcal T}<0$) to light (${\mathcal T}>0$).}
844: \end{figure*}
845:
846: Figure \ref{fig:T1buq} shows ${\mathcal T}_{BU}(K,Q=20)$, the
847: energy received by the magnetic field at all wavenumbers $K$
848: from the velocity field in the shell $Q=20$. During the kinematic regime,
849: the velocity field in this shell gives energy to all magnetic shells,
850: although the transfer peaks at wavenumbers larger than $Q$. But in
851: the saturated regime, the transfer changes drastically again. The magnetic
852: field at wavenumbers $K$ smaller than $Q \approx 16$, and in shells
853: between 20 and 23 gives energy to the velocity field (negative transfer),
854: while the magnetic field in shells between $K \approx 16 $ to 20 and for
855: $K \gtrsim 23$ receives energy from the velocity field (positive transfer).
856: This is just the counterpart of ${\mathcal T}_{BU}(K,Q)$ for constant $K$,
857: and again shows that in Hall-MHD a small scale magnetic field can create
858: large scale flows.
859:
860: Figure \ref{fig:THall} shows shaded plots of
861: ${\mathcal T}_{BB}^\textrm{Hall}(K,Q)$ and ${\mathcal T}_{BU}(K,Q)$
862: at different times. These are the two transfers that are strongly modified
863: by the Hall currents, and the figures allow for a study of the terms for all
864: values of $K$ and $Q$.
865: Although noisy, a characteristic pattern can be recognized in
866: ${\mathcal T}_{BB}^\textrm{Hall}$. As time evolves and the magnetic
867: energy grows, the relative importance of this term grows. For
868: wavenumbers $K,Q \gtrsim k_{Hall}\sim 10$, the function is positive
869: (light) near and below the diagonal $K=Q$, and negative (dark) near and
870: above this diagonal. This region close to the diagonal represents local
871: and direct transfer of energy: a cut at constant $Q$ shows that close to
872: the diagonal the shell $Q$ receives energy from neighboring shells with
873: $K \lesssim Q$ (negative ${\mathcal T}_{BB}^\textrm{Hall}$) and gives
874: energy to neighbor shells with $K \gtrsim Q$ (positive
875: ${\mathcal T}_{BB}^\textrm{Hall}$). As we move far from this diagonal,
876: the sign of the regions above and below the diagonal changes. This
877: indicates a non-local and inverse transfer of magnetic energy, from
878: small to large scales, in agreement with the expression for the
879: turbulent magnetic diffusivity obtained in Sec. \ref{Transport}.
880:
881: The ${\mathcal T}_{BU}(K,Q)$ also shows an interesting behavior as a
882: function of time. During the kinematic regime, ${\mathcal T}_{BU}$ is
883: positive in a triangle defined by $K \gtrsim Q$. This indicates that the
884: velocity field in a given shell amplifies the magnetic field in that shell
885: and all the shells with larger wavenumber (smaller scales). Also a strong
886: band around $Q = 3$ is observed, indicating that the velocity field in the
887: energy injection band gives a lot of energy to the magnetic field. These
888: results are similar to the kinematic MHD dynamo \citep[see][]{Mininni05b}.
889: However, at late times an inverse process can be identified close to the
890: diagonal $K=Q$. Above it, ${\mathcal T}_{BU}$ is positive, while below
891: it, it is negative. This represents transfer of magnetic energy from a shell
892: $K$ to kinetic energy in slightly smaller wavenumbers $Q$.
893:
894: \begin{figure}
895: \begin{center} \includegraphics[width=9cm]{JPPmininni10} \end{center}
896: \caption{\label{fig:T1flu} Energy fluxes in the run with $\epsilon = 0.1$ at
897: $t=45$: $\Pi_{BB}^\textrm{Hall}(k)$ (solid line),
898: $\Pi_{BB}^\textrm{MHD}(k)$ (dash-dotted line), $\Pi_{UU}(k)$ (dashed),
899: and $\Pi_{BU}(k)$ (dotted). The thick line is the total flux
900: in the simulation.}
901: \end{figure}
902:
903: Since ${\mathcal T}_{BB}^\textrm{Hall}$ and ${\mathcal T}_{BU}$ give both
904: direct and inverse transfers of energy (locally or non-locally), it is of
905: interest to quantify which direction wins when all the contributions to the
906: transfer are added. To this end, we computed the contribution of each
907: transfer term to the energy flux. The total energy flux at a wavenumber
908: $k$ is given by
909: \begin{equation}
910: \Pi(k) = \sum_{K=0}^k \sum_Q {\mathcal T}(K,Q) ,
911: \end{equation}
912: where ${\mathcal T} = {\mathcal T}_{UU} + {\mathcal T}_{BB} +
913: {\mathcal T}_{UB} + {\mathcal T}_{BU}$ is the total energy transfer. We
914: can split this flux into the energy flux due solely to the transfer of kinetic
915: energy
916: \begin{equation}
917: \Pi_{UU}(k) = \sum_{K=0}^k \sum_Q {\mathcal T}_{UU}(K,Q) ,
918: \end{equation}
919: the flux due to the transfer of magnetic energy
920: $\Pi_{BB} = \Pi_{BB}^\textrm{MHD} + \Pi_{BB}^\textrm{Hall}$, where
921: \begin{equation}
922: \Pi_{BB}^\textrm{MHD}(k) = \sum_{K=0}^k \sum_Q
923: {\mathcal T}_{BB}^\textrm{MHD}(K,Q) ,
924: \end{equation}
925: \begin{equation}
926: \Pi_{BB}^\textrm{Hall}(k) = \sum_{K=0}^k \sum_Q
927: {\mathcal T}_{BB}^\textrm{Hall}(K,Q) ,
928: \end{equation}
929: and the hybrid flux due to interactions between the velocity and magnetic
930: fields
931: \begin{equation}
932: \Pi_{BU}(k) = \sum_{K=0}^k \sum_Q \left[ {\mathcal T}_{BU}(K,Q) +
933: {\mathcal T}_{UB}(K,Q) \right] .
934: \end{equation}
935: To compute the fluxes, the transfer functions are not normalized.
936:
937: Figure \ref{fig:T1flu} shows the partial energy fluxes at $t=45$. Since
938: all transfer functions were computed up to $K,Q=40$, the partial fluxes
939: go to zero artificially at this wavenumber, although in the simulation
940: the total energy flux goes to zero only at the maximum resolved wavenumber
941: $k_{max}$.
942:
943: The total flux is positive at wavenumbers larger than $k_0$ (the energy
944: injection band), indicating a direct cascade of the total energy. At
945: wavenumbers smaller than $k_0$ the total flux is negative, an evidence
946: of large scale dynamo action. A substantial portion of the total flux is due
947: to the transfer of energy from the kinetic to the magnetic reservoirs
948: ($\Pi_{BU}$), and this contribution to the flux is positive at all
949: wavenumbers (larger than the forced ones) indicating a net direct
950: transfer of the energy. We note that this flux is due to the non-local
951: $T_{UB}$ and $T_{BU}$ transfer terms. The flux due to the transfer of
952: kinetic energy $\Pi_{UU}$ is also positive at all wavenumbers. But the
953: flux due to the transfer of magnetic energy $\Pi_{BB}^{Hall}$ is only
954: positive at wavenumbers larger than $k_{Hall}$. For wavenumbers smaller
955: than $k \approx 10\sim k_{Hall}$, $\Pi_{BB}^\textrm{Hall}$ changes sign,
956: giving as a result a net inverse transfer of magnetic energy, from small
957: to large scales. This indicates that a magnetically dominated Hall-MHD
958: system could display backscatter of the magnetic energy. Magnetic
959: fluctuations at small scales could give rise to large scale magnetic
960: fields, as is also implied by the expression of $\beta$ found in Sec.
961: \ref{Transport}.
962:
963: Note that although in MHD the inverse cascade of magnetic helicity
964: can give a similar result, the backscatter predicted in
965: Hall-MHD by the turbulent diffusivity is novel, since it can take place even
966: in the absence of helicity in the fields. To illustrate this we show
967: results of non-helical magnetically dominated simulations in Sec.
968: \ref{Upscaling}. It is worth noting that at wavenumbers smaller than
969: $k_{Hall}$, the Hall term increases the flux of magnetic energy to
970: smaller scales, thus also in agreement with results showing the Hall
971: currents increase the amount of small scale perturbations
972: \citep{Birn01,Laveder02a,Laveder02b,Morales05}.
973:
974: \subsection{Dependence with $\epsilon$}
975:
976: Now we discuss in detail the dependence of the results as $\epsilon$ (or
977: the Hall scale) is varied. To this end, we consider the runs with
978: $\epsilon = 0.1$, $0.05$, and 0 (MHD). As previously mentioned, the
979: transfer terms ${\mathcal T}_{UU}(K,Q)$ and ${\mathcal T}_{BB}(K,Q)$ do not
980: show a dependence with the amplitude of the Hall effect. These terms give
981: direct and local transfer of energy to small scales, as in MHD
982: \citep{Alexakis05a}. As a result, we will discuss the change in the
983: remaining transfer terms as $\epsilon$ is varied.
984:
985: Since the Hall effect is more relevant when the magnetic field is stronger,
986: we will consider the transfer terms at $t=45$, when a large scale magnetic
987: field is present and the small scales have reached saturation. Since we
988: will examine runs with different values of $\epsilon$ at the same time,
989: in this subsection the transfers are not normalized using the energies.
990:
991: The transfer of magnetic energy due to the Hall term
992: ${\mathcal T}_{BB}^\textrm{Hall}(K,Q=20)$ is of course zero in the MHD
993: case ($\epsilon=0$). As $\epsilon$ is increased, except for an increase
994: in its amplitude (not shown), no significant differences are observed
995: and its behavior is similar to the one examined in Sec. \ref{Transfers}.
996:
997: \begin{figure}
998: \begin{center} \includegraphics[width=9cm]{JPPmininni11} \end{center}
999: \caption{\label{fig:Tebuq} ${\mathcal T}_{BU}(K=20,Q)$ at $t=45$ for
1000: $\epsilon = 0$ (solid line), $0.05$ (dotted line), and $0.1$
1001: (dashed line).}
1002: \end{figure}
1003:
1004: Figure \ref{fig:Tebuq} shows the behavior of ${\mathcal T}_{BU}(K=20,Q)$
1005: (the transfer of kinetic energy in the shells $Q$ to magnetic energy
1006: in the shell $K=20$) as $\epsilon$ is varied. The strong peak at $Q=3$ is
1007: associated with the injection band. This transfer is non-local in the three
1008: runs, as is evidenced by the positive plateau from $Q\approx3$
1009: to $Q\approx 16$. As a result, the velocity field in all these shells
1010: gives energy to the magnetic field at $K=20$. As $\epsilon$ is increased,
1011: a local transfer grows in the neighborhood of $Q=20$. The velocity field
1012: at wavenumbers $K$ slightly larger give energy to the magnetic field at
1013: $K=20$, while the magnetic field gives energy to the velocity field at
1014: wavenumbers slightly smaller ($Q \approx 18$).
1015:
1016: \begin{figure}
1017: \begin{center} \includegraphics[width=9cm]{JPPmininni12} \end{center}
1018: \caption{\label{fig:fluxbb} $\Pi_{BB}(k)$ at $t=45$ for $\epsilon = 0$
1019: (solid line), $0.05$ (dotted line), and $0.1$ (dashed line).}
1020: \end{figure}
1021:
1022: We can compute the energy flux as $\epsilon$ increases. Since the transfer
1023: of kinetic energy is not changed, we will focus on two contributions to
1024: the total flux: the flux of magnetic energy $\Pi_{BB}(k)$, and the
1025: hybrid flux $\Pi_{BU}(k)$ due to the terms turning kinetic into magnetic
1026: energy and vice-versa. The magnetic energy flux
1027: $\Pi_{BB} = \Pi_{BB}^\textrm{MHD} + \Pi_{BB}^\textrm{Hall}$ is shown in
1028: Fig. \ref{fig:fluxbb}. At scales larger than $k_{Hall}$, negative flux
1029: of magnetic energy is observed, giving backscatter of magnetic energy
1030: to large scales. As $\epsilon$ is increased, the amplitude of the
1031: backscatter grows, and the wavenumber where the flux changes sign moves
1032: to larger $k$.
1033:
1034: \begin{figure}
1035: \begin{center} \includegraphics[width=9cm]{JPPmininni13} \end{center}
1036: \caption{\label{fig:fluxbu} $\Pi_{BU}(k)$ at $t=45$ for $\epsilon = 0$
1037: (solid line), $0.05$ (dotted line), and $0.1$ (dashed line).}
1038: \end{figure}
1039:
1040: Figure \ref{fig:fluxbu} shows the flux $\Pi_{BU}(k)$. At wavenumbers
1041: smaller than the forcing wavenumber ($k=3$), the flux is negative. This
1042: is a signature of large scale dynamo action: the magnetic field at
1043: large scales is fed by the small scale velocity field. Remarkably, as
1044: $\epsilon$ is increased, the amplitude of the negative flux at large
1045: scales increases. This is in good agreement with dynamo simulations
1046: where the large scale magnetic field was observed to grow faster in the
1047: presence of Hall currents \citep{Mininni03b,Mininni05a}.
1048:
1049: \subsection{Transfer of magnetic helicity}
1050:
1051: We discuss briefly the transfer of magnetic helicity in the saturated
1052: case $t=45$. To study the transfer associated with the inverse cascade
1053: of magnetic helicity at scales larger than the forcing scale, a large
1054: separation between this scale and the largest scale in the box is needed.
1055: At a fixed spatial resolution, this reduces the Reynolds numbers, and
1056: as a result reduces also the separation between the Ohmic scale and the
1057: Hall scale. This study is beyond the aim of this work. But we want to
1058: point out a remarkable feature observed in the transfer of magnetic
1059: helicity at scales smaller than the forcing scale.
1060:
1061: \begin{figure}
1062: \begin{center} \includegraphics[width=9cm]{JPPmininni14} \end{center}
1063: \caption{\label{fig:Thelicity} ${\mathcal T}_H(K,Q=20)$ normalized by the
1064: magnetic helicity at the shell $Q=20$, at $t=45$ for $\epsilon = 0$
1065: (solid line), $0.05$ (dotted line), and $0.1$ (dashed line).}
1066: \end{figure}
1067:
1068: Figure \ref{fig:Thelicity} shows the transfer ${\mathcal T}_H(K,Q=20)$
1069: normalized by the magnetic helicity in the shell $Q=20$ at $t=45$.
1070: The transfer is mostly local in the three simulations, peaking at
1071: wavenumbers $K$ slightly smaller and larger than $Q$. However, as
1072: $\epsilon$ is increased the transfer rate of magnetic helicity
1073: is strongly quenched. This slow down in the transfer in Hall-MHD
1074: explains the behavior observed in Fig. \ref{fig:heli}. In MHD and
1075: Hall-MHD dynamos, the external mechanical forcing generates equal
1076: amounts of magnetic helicity of opposite sign at scales smaller
1077: and larger than the forcing band \citep{Seehafer96,Brandenburg01,Mininni03b}.
1078: Since the transfer of magnetic helicity between different shells in
1079: the Hall-MHD runs is almost stopped, it takes more time for the
1080: magnetic helicity at scales smaller than the forcing scale to reach
1081: the dissipative scale where it can be destroyed. As a result, both
1082: signs of magnetic helicity piles up close to the forcing band,
1083: decreasing the growth rate of net magnetic helicity at scales larger
1084: than the forcing scale, and allowing also for
1085: the possibility for a sign change of the net magnetic helicity.
1086:
1087: \section{ \label{Upscaling} Backscatter of magnetic energy in Hall-MHD}
1088:
1089: The mechanically forced runs discussed in the previous section show
1090: negative flux of magnetic energy at large scales due to the Hall effect,
1091: in agreement with negative values of the turbulent diffusivity. This
1092: indicates that in a magnetically dominated simulation, backscatter of
1093: magnetic energy could be observed if the Hall term is strong enough.
1094: Note that here we are using the word {\it backscatter} to refer to this
1095: transfer of magnetic energy from the small to the large scales. This is
1096: done in opposition to the usual terminology of inverse cascades, since
1097: we have been unable to identify any ideal invariant of the Hall-MHD
1098: equations cascading inversely with constant flux to the large scales.
1099:
1100: To study this scenario, we did three simulations with $\epsilon = 0$,
1101: $0.2$, and $0.5$. The kinematic viscosity and magnetic diffusivity
1102: were $\nu = \eta = 3.5 \times 10^{-2}$, and the spatial resolution was
1103: $N^3 = 128^3$. The initial condition was ${\bf U} = {\bf B} = 0$. The
1104: system was forced with a non-helical and random electromotive force given
1105: by a superposition of harmonic modes at wavenumbers $k=9$ and 10. The
1106: phases of the force were changed with a correlation time of
1107: $\tau = 1.25 \times 10^{-2}$, and the time step was set to
1108: $\Delta t = 1.5 \times 10^{-3}$. Note that in the absence of magnetic
1109: helicity, no inverse cascade is expected in the MHD case
1110: \citep[see however][for cases where a large scale shear is present]{Lanotte99}.
1111:
1112: \begin{figure}
1113: \begin{center} \includegraphics[width=9cm]{JPPmininni15} \end{center}
1114: \caption{\label{fig:invsp1} (a) Kinetic energy spectrum $E_U(k)$ and
1115: (b) magnetic energy spectrum $E_B(k)$ at $t=3$, for runs with
1116: $\epsilon = 0$ (solid line), $0.05$ (dotted line), and $0.1$
1117: (dashed line).}
1118: \end{figure}
1119:
1120: \begin{figure}
1121: \begin{center} \includegraphics[width=9cm]{JPPmininni16} \end{center}
1122: \caption{\label{fig:invsp2} (a) Kinetic energy spectrum and (b)
1123: magnetic energy spectrum at $t=75$ for runs with $\epsilon = 0$
1124: (solid line), $0.05$ (dotted line), and $0.1$ (dashed line).}
1125: \end{figure}
1126:
1127: The system was run until reaching a turbulent steady state. All the
1128: quadratic invariants (with the exception of the total energy) were
1129: verified to be small: the magnetic helicity fluctuates in the three
1130: runs around zero, both the global quantity as well as its spectral
1131: density at each individual Fourier shell. Figure \ref{fig:invsp1}
1132: shows the kinetic and magnetic energy spectrum at early times ($t=3$).
1133: The shell of wavenumbers associated with the external magnetic force
1134: is easily recognized in the peak in Fig. \ref{fig:invsp1}(b).
1135:
1136: As time evolves, an increase in the magnetic energy at wavenumbers
1137: smaller than the forcing wavenumber is observed. Figure \ref{fig:invsp2}
1138: shows the kinetic and magnetic energy spectrum at $t=75$, when the
1139: system has reached a steady state. The three runs are dominated by
1140: the magnetic energy (note that the peak in the magnetic energy spectrum
1141: around the forcing band gives the largest contribution to the energy).
1142: The spectrum of kinetic energy is similar for the three runs, and
1143: large scale perturbations are observed because of the injection of
1144: kinetic energy by the Lorentz force. However, the magnetic energy
1145: spectrum is strongly modified as $\epsilon$ is increased. While the
1146: spectra of the three simulations peak in the energy injection band,
1147: the magnetic energy at $K=1$ in the run with $\epsilon=0.5$ is
1148: three orders of magnitude larger than in the MHD run. The magnetic
1149: energy in all wavenumbers smaller than the forcing wavenumbers
1150: increases as $\epsilon$ is increased.
1151:
1152: The backscatter of magnetic energy is in good agreement with the
1153: negative flux $\Pi_{BB}$ observed in the previous section, and the
1154: negative turbulent transport coefficients derived for the magnetically
1155: dominated case. Note that an increase in the level of the small
1156: scale magnetic fluctuations (for wavenumbers smaller than the energy
1157: injection wavenumbers) is also observed in Fig. \ref{fig:invsp2}(b).
1158:
1159: \section{ \label{Discussion} Discussion }
1160:
1161: In this work we presented energy transfer in Hall-MHD turbulence as
1162: obtained from numerical simulations. The properties of the spectral
1163: transfer is one of the building blocks of turbulence theories, and to
1164: the best of our knowledge no attempt to study transfer and cascades of
1165: ideal invariants in this system of equations had been attempted before.
1166:
1167: Before proceeding with the discussion of our results, we have to
1168: warn the reader about a clear limitation of the numerical results
1169: presented. As previously mentioned, an astrophysics-like scale
1170: separation between the box size, the energy injection scale, the Hall scale,
1171: and the Ohmic dissipation scale is well beyond today computing resources.
1172: We tested the dependence of our results as $\epsilon$ was varied, but
1173: no attempt was made to change the Reynolds numbers in our simulations.
1174: This being said, we believe that even under this limitation, an
1175: understanding of the transfer of energy between different scales is of
1176: uttermost importance for the development of a theory of turbulence for
1177: Hall-MHD or other extensions of magnetohydrodynamic to take into account
1178: kinetic plasma effects.
1179:
1180: Direct evidence of nonlocality of the energy transfer was observed. While
1181: the total energy displays a direct cascade to small scales, in the
1182: individual transfer terms, both directions (toward small and large scales)
1183: were identified. Coupling between the magnetic and velocity fields is
1184: strongly modified by the Hall effect, and a local backscatter of energy
1185: from the magnetic field to the velocity field at slightly larger scales
1186: was observed. This behavior can be expected since the Hall term changes
1187: the nature of the nondispersive MHD Alfv\'en waves, into dispersive and
1188: circularly polarized waves. As a result, the nonlinear coupling between
1189: the two fields is also changed.
1190:
1191: Also a nonlocal backscatter of magnetic energy was observed at scales
1192: larger than the Hall scale. This backscatter was verified in non-helical
1193: magnetically forced simulations, where the amplitude of the magnetic
1194: field at scales larger than the forcing scale was observed to grow in
1195: the Hall-MHD simulations, but not in the MHD run. In some sense, the
1196: magnetic field in Hall-MHD being frozen in the ideal case to the
1197: electron velocity field, couples non-locally both small scales
1198: (the current) and large scales (the bulk velocity field).
1199:
1200: All these processes can be partially explained considering transport
1201: turbulent coefficients estimated from MFT.
1202: Unlike MHD, the turbulent diffusivity in Hall-MHD
1203: is not positive definite. In particular, its expression shows that
1204: ion-cyclotron waves are more likely to produce large values of negative
1205: (backscatter) or positive (reconnection) turbulent diffusivity than
1206: the whistler mode.
1207:
1208: The transfer of magnetic helicity at small scales was also observed
1209: to be quenched by the Hall effect. While the mechanisms generating
1210: magnetic helicity in the Hall-MHD dynamo are the same as in MHD
1211: \citep{Mininni03b}, the transport of helicity is expected to be changed
1212: by the Hall currents \citep{Ji99}. As a result of the slow down in the
1213: transfer rate of magnetic helicity by the Hall effect, the late time
1214: evolution of the system is not characterized by a maximally helical
1215: large scale magnetic field as in the MHD case
1216: \citep{Pouquet76,Meneguzzi81,Brandenburg01}.
1217:
1218: The Hall term gives a direct transfer of magnetic energy at scales
1219: smaller than the Hall scale, and an inverse transfer at scales larger
1220: than the Hall scale. This finding sheds light into the conflicting
1221: results reported in the literature, where the Hall effect was observed
1222: to increase the amount of small scales and magnetic dissipation in some
1223: cases, and to help large scale reorganization processes in other cases,
1224: as mentioned in the introduction.
1225:
1226: As a result of this dual direction of the Hall transfer, a change in
1227: the power law followed by the total energy spectrum can be expected close
1228: to the Hall wavenumber. Steepening of the energy spectrum for wavenumbers
1229: smaller than $k_{Hall}$ was observed in 2.5D simulations with strong
1230: magnetic fields imposed, when the cross-correlation between the velocity
1231: and magnetic fields was significant \citep{Ghosh96}. In three dimensional
1232: dynamo simulations where the cross correlation is in general small,
1233: no change was observed \citep{Mininni05a}, although a faster
1234: growth of the large scale magnetic field was found. Given the nonlocal
1235: nature of the transfer in Hall-MHD, and the scale separation needed
1236: to observe a clear change in the energy spectrum, probably a huge
1237: increase in the spatial resolution is needed to confirm it.
1238:
1239: \begin{acknowledgments}
1240: Computer time was provided by NCAR. The NSF grant CMG-0327888
1241: at NCAR supported this work in part and is gratefully acknowledged.
1242: \end{acknowledgments}
1243:
1244: \bibliography{ms}
1245:
1246: \end{document}
1247: