1: % $Id: ms.tex 4009 2009-06-05 15:34:22Z graham $
2: \documentclass[aps,showpacs,preprint,groupedaddress,amsmath,amssymb,superscriptaddress,letterpaper]{revtex4}%letterpaper
3: \usepackage{graphicx}% Include figure files
4: \usepackage{dcolumn} % Align table columns on decimal point
5: \usepackage{bm} % bold math
6: \usepackage{times}
7: \usepackage{color}
8: \usepackage{aas_macros}
9:
10: \usepackage{setspace}
11:
12: \newcommand{\leraya}{{Leray$-\alpha$} }
13: \newcommand{\clarka}{{Clark$-\alpha$} }
14: \newcommand{\lansa}{{LANS} }
15: \newcommand{\leray}{Leray$-\alpha$}
16: \newcommand{\clark}{Clark$-\alpha$}
17: \newcommand{\lans}{LANS}
18: \newcommand{\lamhd}{LAMHD}
19: \newcommand{\lamhda}{{LAMHD} }
20: \newcommand{\Res}{Reynolds numbers}
21: \newcommand{\subgrid}{{subgrid} }
22: \newcommand{\dof}{{$N_{dof}$}}
23: \newcommand{\dofa}{{$N_{dof}$} }
24: \newcommand{\pdf}{{\sl pdf}}
25: \newcommand{\pdfa}{{\sl pdf} }
26: \renewcommand{\vec}{\mathbf}
27: \newcommand{\uvec}{\bar{\vec{v}}}
28:
29: % COLOR %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: %\def\note#1{{\textcolor{red}{\bf [#1]}}} % note
31: %\def\ADD#1{{\textcolor{red}{\bf [#1]}}} % remark or question
32: %\def\ADP#1{{\textcolor{green}{\bf #1}}} % add or modify
33: %\def\ADDP#1{{\textcolor{magenta}{\bf [#1]}}} % remark or question
34:
35: %\def\vier#1{{\textcolor{blue}{\bf #1}}} % response 4
36: \def\vier#1{{#1}} % responsefour
37: %\def\neu#1{{\textcolor{blue}{\bf #1}}} % response
38: \def\neu#1{{#1}} % add or modify
39: %\newcommand{\resp}[2]{ \marginpar{\fbox{\color{blue}\bf #1}}{\bf \color{blue}\bf #2} }
40: \newcommand{\resp}[2]{{#1}{#2}}
41: %\def\resp#1{#1} % response
42: %\newcommand{\neu}[2]{{#1}{#2}}
43: %\newcommand{\neu}[2]{ \marginpar{\fbox{\color{blue}\bf #1}}{\bf \color{blue}\bf #2} }
44:
45: %\newcommand{\add}[2]{ \marginpar{\fbox{\color{blue}\bf #1}}{\bf \color{blue}\bf #2} }
46: %\def\AD#1{{\textcolor{blue}{\bf [#1]}}} % add or modify
47: \newcommand{\add}[2]{{#1}{#2}}
48: \def\AD#1{{#1}} % add or modify
49:
50: %\def\thesis#1{} % in thesis only
51: %\def\paper#1{#1} % in paper only
52: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53:
54: %\topmargin -3pt
55: \begin{document}
56:
57: \title{
58: The Lagrangian-averaged model for
59: magnetohydrodynamics turbulence and the absence of bottleneck}
60:
61: \author{Jonathan \surname{Pietarila Graham}}
62: \affiliation{Max-Planck-Institut f\"ur
63: Sonnensystemforschung, 37191 Katlenburg-Lindau, Germany}
64: \author{Pablo D. Mininni}
65: \affiliation{National
66: Center for Atmospheric Research,\footnote{The National Center for
67: Atmospheric Research is sponsored by the National Science
68: Foundation} P.O. Box 3000, Boulder, Colorado 80307, USA}
69: \affiliation{Departamento
70: de F\'\i sica, Facultad de Ciencias Exactas y Naturales, Universidad de
71: Buenos Aires, Ciudad Universitaria, 1428 Buenos Aires, Argentina}
72: \author{Annick Pouquet}
73: \affiliation{National
74: Center for Atmospheric Research,\footnote{The National Center for
75: Atmospheric Research is sponsored by the National Science
76: Foundation} P.O. Box 3000, Boulder, Colorado 80307, USA}
77:
78: \date{\today}
79:
80: \begin{abstract}
81: We demonstrate that, for the case of quasi-equipartition between the
82: velocity and the magnetic field, the Lagrangian-averaged
83: magnetohydrodynamics $\alpha-$model (LAMHD) reproduces well both the
84: large-scale and small-scale properties of turbulent flows; in
85: particular, it displays no increased (super-filter) bottleneck effect
86: with its ensuing enhanced energy spectrum at the onset of the
87: sub-filter-scales. This is in contrast to the case of the neutral
88: fluid in which the Lagrangian-averaged Navier-Stokes $\alpha-$model is
89: somewhat limited in its applications because of the formation of
90: spatial regions with no internal degrees of freedom and subsequent
91: contamination of super-filter-scale spectral properties. \resp{}{We
92: argue that, as the Lorentz force breaks the conservation of
93: circulation and enables spectrally non-local energy transfer
94: (associated to Alfv\'en waves), it is responsible for the absence of
95: a viscous bottleneck in MHD, as compared to the fluid case. As
96: LAMHD preserves Alfv\'en waves and the circulation
97: properties of MHD, there is also no (super-filter)
98: bottleneck} found in LAMHD, making this method capable of large
99: reductions in required numerical degrees of freedom; specifically, we
100: find a reduction factor of $\approx 200$ when compared to a direct
101: numerical simulation on a large grid of $1536^3$ points at the same
102: Reynolds number.
103:
104: \end{abstract}
105:
106: \pacs{47.27.ep; 52.30.Cv; 95.30.Qd; 47.27.E-}
107: \maketitle
108:
109:
110: \section{Introduction}
111:
112: When large-scale numerical simulations of astrophysical or geophysical
113: magnetohydrodynamics (MHD) are desired, all dynamical scales of the
114: physical system are rarely, if ever, resolved. For this reason,
115: sub-grid-scale (SGS) modeling of MHD dynamics in the context of computations in the geophysical and astrophysical context is required.
116: This modeling can be achieved implicitly, in the simplest example by employing a dissipative
117: numerical scheme, or it can be done explicitly by creating a Large Eddy Simulation
118: (LES--see \cite{MK00} for a recent review). Explicit methods for MHD are not as pervasive as they are in engineering, or for geophysical and atmospheric flows. In fact,
119: modeling for MHD is a
120: relatively new field (see \cite{PFL76,Y87}). One problem with
121: extending the LES methodology for hydrodynamic turbulence to MHD is that most LES are
122: based upon eddy-viscosity concepts \cite{MK00}, which can be related
123: to a known power law of the energy spectrum \cite{ChLe1981} (although generalizations can be devised, see e.g. \cite{BaPoPo+2008}), or upon
124: self-similarity. For MHD, the underlying assumption of
125: locality of interactions in Fourier
126: space is not necessarily valid \cite{AMP05a,MAP05} (a contradiction of
127: self-similarity) and spectral eddy-viscosity concepts \cite{ZSG02}
128: cannot be applied in a straightforward manner as neither kinetic nor magnetic energy is a
129: conserved quantity and the general expression of the energy spectrum
130: is not known at this time \cite{I64,K65,GoSr1995,GaNaNe+2000,MiPo2007a,PoMiMo+2008,MaCaBo2008}.
131: %\note{BaPoPo+2008 is: J. Baerenzung, H. Politano, Y. Ponty and A. Pouquet, ``Spectral Modeling of Turbulent Flows and the Role of Helicity,'' {\it Phys. Rev. E} {\bf 77}, 046303 (2008). JPG: Sorry, but there is also "{Spectral Modeling of Magnetohydrodynamic Turbulent Flows}" by the same authors--one should be BaPoPo+2008a and the other BaPoPo+2008b. This page at least seems to be the MHD paper...}
132: Purely dissipative models
133: \cite{TFS94,AMK+01} are inadequate as they ignore the exchange of
134: energy at sub-filter scales between the velocity and magnetic fields
135: and such models have been shown to suppress small-scale dynamo action
136: \cite{HB06} and any inverse cascade from the sub-filter scales
137: \cite{MC02}. \add{}{A satisfactory LES for MHD has been proposed} for the case starting
138: with some degree of alignment between the velocity and magnetic fields
139: \cite{LS91,MC02}. Other restricted-case MHD-LES are applicable to low
140: magnetic Reynolds number \cite{PPP04,KM04,PMM+05}.
141: Extensions of spectral models to MHD based on two-point closure formulations of the dynamical equations proposed recently look promising in the analysis of turbulent flows and of the dynamo mechanism \cite{BaPoPo+2008}.
142: Finally, though technically
143: not an LES, there are also hyper-resistive models for MHD which
144: require rescaling of the length (wavenumber) scales to a known direct
145: numerical simulation (DNS) \cite{HB06}.
146:
147: One model which can be written as an LES is the Lagrangian-averaged
148: MHD (\lamhd) equations \cite{H02a, H02b, MP02}. It has been shown to
149: reproduce a number of features of DNS. \resp{}{In two dimensions (2D)
150: for Taylor Reynolds numbers ($R_\lambda$) up to $\approx5000$ it has
151: been shown to reproduce selective decay, the inverse cascade of
152: mean-square vector potential, and dynamic alignment between
153: the velocity and magnetic fields \cite{MMP05a} as well as the statistics \vier{of} small-scale cancellation \cite{PGMP05} and intermittency
154: \cite{PGHM+06}. In three dimensions (3D) at Reynolds numbers ($Re$)
155: of $\approx500$, \lamhda reproduced the inverse cascade of magnetic
156: helicity (associated with the development of force-free magnetic
157: field) and the helical dynamo effect \cite{MMP05b}.} \neu{It has
158: also been tested (up to kinetic $Re\approx3000$, magnetic
159: $Re\approx300$) for its ability to predict the critical magnetic
160: Reynolds number for a non-helical dynamo at low magnetic Prandtl
161: number \cite{Mi2006a}. \lamhda performed well in all these
162: tests. Its equivalent hydrodynamic model, the Lagrangian-averaged
163: Navier-Stokes (LANS) equations, also performed well in tests at
164: $R_\lambda\lessapprox300$ (see \cite{CHO+05} and references in
165: \cite{PGHM+07a}). However, above $Re\approx3000$
166: ($R_\lambda\approx800$),} it was shown that placing the filter width
167: in the inertial range leads to contamination of the super-filter-scale
168: properties (such as the spectra) \vier{for \lans.} We refer here to this effect as the
169: super-filter-scale bottleneck, which \resp{}{may be} different in
170: nature from the viscous bottleneck observed in some DNS of the
171: Navier-Stokes equations. The contamination may be linked to the
172: formation of spatial regions in the flow with no internal degrees of
173: freedom (so-called ``rigid bodies'') \cite{PGHM+07a}, which also
174: correspond to the development of a secondary inertial range of the
175: LANS equations at sub-filter scales. This \resp{}{super-filter-scale
176: contamination} provides an effective constraint on the filter size
177: and, hence, on the available reduction of the total number of the
178: (numerical) degrees of freedom (\dof) needed to reproduce the
179: large-scale dynamics of the flow at a given Reynolds number; a factor
180: of $\approx 10$ \resp{}{can be achieved. This limitation is not
181: apparent} in low and moderate \resp{}{Reynolds number} (resolution)
182: simulations (e.g., $64^3$ \lansa compared with $256^3$ DNS) as the
183: scale separation is not enough for the above-mentioned phenomenon of
184: contamination of small-scale spectra because of rigid body regions in
185: the flow to appear. \resp{}{The bottleneck (and super-filter-scale
186: contamination) was not studied as such but neither was it observed
187: in 2D \lamhda for high Reynolds number \cite{MMP05a,PGMP05,PGHM+06}.
188: 3D \lamhda has only been tested} at \add{}{more} moderate Reynolds
189: number \cite{MMP05b} (see also \cite{MiPoSu2008} for a recent review).
190: The aim of the present work is, thus, to \resp{}{determine if} \lamhda
191: in three space dimensions, \add{}{for higher Reynolds number}
192: \resp{}{develops problems similar to that of \lans. Specifically, we
193: test for the existence of} spatial regions with no available
194: internal degrees of freedom. We show in the following that \lamhda
195: behaves better in this respect than \lans, and, thus, continues to
196: appear as a promising model for MHD flows.
197:
198: \section{The equations of motion}
199: \label{SEC:DETAILS}
200:
201: We consider the incompressible MHD equations for a fluid with
202: constant density,
203: \begin{eqnarray} \partial_t\vec{v} + \boldsymbol{\omega} \times \vec{v}
204: = \vec{j} \times \vec{b} - \boldsymbol{\nabla} p + \nu \nabla^2 \vec{v} \nonumber \\
205: \partial_t \vec{b} = \boldsymbol{\nabla} \times \left( \vec{v} \times \vec{b} \right) + \eta \nabla^2 \vec{b}
206: \nonumber \\
207: \boldsymbol{\nabla} \cdot \vec{v} = \boldsymbol{\nabla} \cdot \vec{b} = 0,
208: \label{eq:mhd}
209: \end{eqnarray}
210: where $\vec{v}$ and $\vec{b}$ denote respectively the velocity and magnetic fields, $p$
211: the pressure divided by the density, $\nu$ the kinematic viscosity,
212: and $\eta$ the magnetic diffusivity. %The time, $t$, is expressed in
213: %units of the eddy-turnover time.
214: \resp{}{As is well known, in incompressible MHD, Alfv\'en waves will travel
215: along a uniform background field, $\vec{b}_0$. From linear
216: perturbation analysis the dispersion relation between wavenumber,
217: $k$, and frequency, $\omega$, is
218: \begin{equation}
219: \left(\omega+i\eta k^2\right)\left(\omega+i\nu k^2\right) = k^2b_0^2\,.
220: \end{equation}
221: The wave speed is $|\vec{b}_0|$ and, assuming $\eta=\nu$, the amplification
222: factor is given by $\exp(-\eta k^2t)$.}
223: The ideal ($\eta = \nu = 0$) quadratic invariants for MHD are in the
224: $L^2$ norm. For example, the total energy is given by:
225: \begin{equation}
226: E_T = \frac{1}{2} \left( ||v||_2 + ||b||_2 \right)\equiv
227: \frac{1}{2} \frac{1}{D} \int_D \left( |\vec{v}|^2 + |\vec{b}|^2 \right) d^3x.
228: \end{equation}
229:
230: The \lamhda equations \cite{H02a}
231: are given by
232: \begin{eqnarray} \partial_t\vec{v} + \boldsymbol{\omega} \times \vec{u}
233: = \vec{j} \times \bar{\vec{b}} - \boldsymbol{\nabla} \Pi + \nu \nabla^2 \vec{v} \nonumber \\
234: \partial_t \bar{\vec{b}} = \boldsymbol{\nabla} \times \left( \vec{u} \times \bar{\vec{b}} \right) + \eta \nabla^2 \vec{b}
235: \nonumber \\
236: \boldsymbol{\nabla} \cdot \vec{v} = \boldsymbol{\nabla} \cdot \vec{u} = \boldsymbol{\nabla} \cdot \vec{b} = \boldsymbol{\nabla} \cdot \bar{\vec{b}} = 0,
237: \label{eq:lamhd}
238: \end{eqnarray}
239: where $\vec{u}$ ($\bar{\vec{b}}$) denotes the filtered component of
240: the velocity (magnetic) field and $\Pi$ the modified pressure.
241: Filtering is accomplished by the application of a normalized
242: convolution filter $L: f \mapsto \bar{f}$ where $f$ is any scalar or
243: vector field. By convention, we define $\vec{u} \equiv
244: \bar{\vec{v}}$.
245: %The parameter $\alpha$ is a filtering length that can be shown to be related to the ratio of the dissipation scales of the \lamhda flow and the modeled
246: %MHD flow.
247: \add{}{\lamhda in the form given in Eqs. (\ref{eq:lamhd}) is both computationally efficient and makes clear that Alfv\'en's theorem is preserved by the model: the smoothed magnetic field is advected by the smoothed velocity.}
248: In the remainder of this paper, we take $\eta=\nu$ (unit magnetic Prandtl number) and, thus, it is sufficient to introduce \add{}{the same filtering for the velocity and magnetic fields in this case. This allows us to write \lamhda in
249: LES form,
250: \begin{eqnarray} \partial_t\vec{u} + \bar{\boldsymbol{\omega}} \times \vec{u}
251: = \bar{\vec{j}} \times \bar{\vec{b}} - \boldsymbol{\nabla} \bar{\Pi} + \nu \nabla^2 \bar{\vec{v}} - \vec{\nabla}\cdot\bar{\tau} \nonumber \\
252: \partial_t \bar{\vec{b}} = \boldsymbol{\nabla} \times \left( \vec{u} \times \bar{\vec{b}} \right) + \eta \nabla^2 \bar{\vec{b}} - \vec{\nabla}\cdot\bar{\tau}^b \nonumber \\
253: \boldsymbol{\nabla} \cdot \vec{v} = \boldsymbol{\nabla} \cdot \vec{u} = \boldsymbol{\nabla} \cdot \vec{b} = \boldsymbol{\nabla} \cdot \bar{\vec{b}} = 0.
254: \label{eq:lamhdLES}
255: \end{eqnarray}
256: } We choose as our filter the inverse of a Helmholtz
257: operator, $L = \mathcal H^{-1} = (1 - \alpha^2\nabla^2)^{-1}$.
258: Therefore, $\vec{u} = g_\alpha \otimes \vec{v}$ where $g_\alpha$ is
259: the Green's function for the Helmholtz operator, $g_\alpha(r) = \exp
260: (-r/\alpha)/(4\pi\alpha^2r)$ (i.e., the Yukawa potential), or in
261: Fourier space, $\hat{\vec{u}}(k) = \hat{\vec{v}}(k)/(1+\alpha^2k^2)$.
262: The effective filter width is, thus, approximately $\alpha$.
263: \add{}{With this choice,
264: the Reynolds (turbulent) SGS stress tensor is given by
265: \begin{equation}
266: \bar{\tau} = \alpha^2 \left( \vec{\nabla}\vec{u}\cdot\vec{\nabla}\vec{u}^T +
267: \vec{\nabla}\vec{u}\cdot\vec{\nabla}\vec{u} -
268: \vec{\nabla}\vec{u}^T\cdot\vec{\nabla}\vec{u} -
269: \vec{\nabla}\bar{\vec{b}}\cdot\vec{\nabla}\bar{\vec{b}}^T -
270: \vec{\nabla}\bar{\vec{b}}\cdot\vec{\nabla}\bar{\vec{b}} +
271: \vec{\nabla}\bar{\vec{b}}^T\cdot\vec{\nabla}\bar{\vec{b}}
272: \right)
273: \end{equation}
274: and the divergence of the electromotive-force (emf) SGS stress tensor by
275: \begin{equation}
276: \vec{\nabla}\cdot\bar{\tau}^b = \eta\alpha^2 \nabla^4 \bar{\vec{b}}.
277: \label{eq:emf}
278: \end{equation}
279: In this form, the expression of the SGS tensors make explicit the fact that
280: $\vec{u} = \pm\bar{\vec{b}}$ Alfv\'en waves are preserved even in the subgrid scales.
281: \resp{}{These $\vec{u} = \pm\bar{\vec{b}}$ waves} \resp{}{travel along $\bar{\vec{b}}_0$ (the smoothed
282: and unsmoothed fields are identical for uniform ${\vec{b}}_0$) and the dispersion relation is
283: \begin{equation}
284: \left(\omega+i\nu k^2\right)\left(\omega+i\eta k^2 (1+\alpha_M^2k^2)\right) = k^2\bar{b}_0^2 \frac{1+\alpha_M^2k^2}{1+\alpha_K^2k^2}\,,
285: \end{equation}
286: where $\alpha_K$ and $\alpha_M$ are the filter widths for the
287: smoothing of the velocity and magnetic fields, respectively. For
288: $\alpha\equiv\alpha_K=\alpha_M$ \vier{and $\eta=\nu$} (the case we study), the wave speed is
289: given by $\bar{\vec{b}}_0 \vier{\left(1-(\eta k\alpha^2k^2/\bar{b}_0)^2/8 + \mathcal{O}((\eta k\alpha^2k^2/\bar{b}_0)^6)\right)}$, the strength of the smoothed background
290: magnetic field \vier{reduced by an order $\alpha^4k^4$ term.} The amplification factor is given by
291: \vier{$\exp\left(-\eta k^2t (1+\alpha^2k^2/2)\right)$ for both } $u=-\bar{b}$ waves
292: traveling in the direction of $\bar{\vec{b}}_0$ and
293: $u=\bar{b}$ waves traveling anti-parallel to
294: $\bar{\vec{b}}_0$.} Finally,} the ideal quadratic invariants for
295: \lamhda are in the $H^1_\alpha(\bar{f})$ norm. For example, the total
296: energy is given by a mixture of the smooth and rough fields, namely:
297:
298: \begin{eqnarray}
299: E_T^\alpha = \frac{1}{2} \left( ||u||_2^\alpha + ||\bar{b}||_2^\alpha \right) \equiv
300: \frac{1}{2} \frac{1}{D} \int_D \left( \vec{u} - \alpha^2\nabla^2 \vec{u}\right)
301: \cdot \vec{u} + \left( \bar{\vec{b}} - \alpha^2\nabla^2 \bar{\vec{b}}\right)
302: \cdot \bar{\vec{b}}~d^3x \nonumber \\
303: = \frac{1}{2} \frac{1}{D} \int_D \vec{v}\cdot\vec{u} + \vec{b}\cdot\bar{\vec{b}}~d^3x.
304: \end{eqnarray}
305:
306:
307: We solve both sets of equations, Eqs. (\ref{eq:mhd}) and
308: (\ref{eq:lamhd}), for one specific instance of a decaying MHD flow,
309: using a parallel pseudospectral code \cite{GMD05,GMD05b} in a
310: three-dimensional (3D) cube with periodic boundary conditions. The
311: initial conditions for the velocity and magnetic fields are
312: constructed from a superposition of three Beltrami (helical) ABC flows
313: to which smaller-scale random fluctuations are added with initial
314: kinetic and magnetic energy $E_K = E_M = 0.5$, magnetic helicity $H_M
315: = <\vec{a}\cdot\vec{b}> \approx 0.45$ ($\vec{b}
316: =\vec{\nabla}\times\vec{a}$ where $\vec{a}$ is the vector potential
317: and the brackets denote volume average), and the initial co-alignment
318: of the fields, $\left<\vec{v}\cdot \vec{b}\right>
319: \left<|\vec{v}||\vec{b}|\right>^{-1} \approx 10^{-4}$ (see
320: \cite{MiPoMo2006,MiPo2007a} for details). A MHD-DNS with a resolution
321: $N^3 = 1536^3$ \add{}{(i.e., 1536 grid points in real space in each
322: direction)} and $\eta = \nu = 2 \cdot 10^{-4}$ is used as our high
323: Reynolds number test case for the \lamhda model. The DNS computation
324: is stopped when the growth of the total dissipation begins to enter
325: the saturation phase ($t=3.7$), at which time the Reynolds number
326: based on the mechanical integral scale is $Re \approx 9200$ and the
327: Taylor Reynolds number $\approx 1100$. The MHD flow resulting from
328: the initial conditions employed has previously been analyzed for its
329: spectral properties and for the spatial structures it develops
330: \cite{MiPoMo2006,MiPo2007a,MaPoMi+2008}. In this paper, we perform a
331: simulation with similar initial conditions and parameters but now
332: using \lamhda at a resolution of $512^3$ grid points; we also perform
333: for comparison purposes a Navier-Stokes \lansa run with the same
334: initial velocity field but with $\vec{b} \equiv 0$, on a grid of
335: $512^3$ points. In both cases, the filter width is $\alpha=2\pi/18$
336: ($k_{\alpha}=18$) and is, thus, large enough to preclude any artifact
337: of numerical resolution altering the results. Based on previous
338: analyses \cite{FHT01,PGHM+07a}, we estimate ${k_{max}}/{k_\eta^\alpha}
339: \approx 2.4$ (where $k_{max}$ is the maximum wavenumber resolved in
340: the simulation, and $k_\eta^\alpha$ is the \lamhda dissipation scale)
341: using computations conducted for $\eta=\nu=6\cdot10^{-4}$ with a
342: Reynolds number of $Re \approx 2200$. \resp{}{However, the main point
343: of using such a large filter is to test if \lamhda fails in the same
344: way as \lans.} We finally perform a LES simulation in a $256^3$
345: grid using the \lamhda equations with the same viscosity and
346: diffusivity as the $1536^3$ DNS used for the comparison. \resp{}{In
347: this way, we extend the $Re\approx9200$ computation in time by a
348: factor of 3.}
349:
350: \section{Results}
351: \subsection{Spectral contamination in \lansa \ for an ABC flow and its absence in the MHD case}
352: \label{SEC:LAMHDGOOD}
353:
354: One of the main findings of our preceding work with \lansa on the Navier-Stokes equations is that a $k^{+1}$ scaling develops in the (kinetic) energy spectrum at sub-filter scales; this leads to a contamination of super-filter scales because of detailed energy conservation (per triadic interactions).
355: This \lansa $k^{+1}$ spectrum (together with super-filter-scale spectral
356: contamination) has only recently been recognized, in the case of one specific forcing function
357: at large Reynolds number \cite{PGHM+07a}, but such a spectral contamination
358: has not yet been generally demonstrated (although theoretical arguments for the $k^{+1}$ spectrum have been given in \cite{PGHM+07a}). Thus, we first confirm
359: its presence in a \lansa simulation with the same viscosity and the
360: (nearly) same initial conditions for the velocity field as for the
361: MHD-DNS (and \lamhda runs) examined in this paper, and based on large-scale ABC flows with superimposed random noise at small scale.
362: Due to the presence of random noise and considering the differences in resolution and the presence of a filter in the \lamhda runs, the initial conditions were not exactly reproduced, although the same procedure was used to generate them.
363: In the present Navier-Stokes case, we find again what can be called an enhanced (super-filter-scale) bottleneck:
364: the positive power-law spectral
365: contamination of the kinetic energy spectrum $E_K(k)$ in the \lansa run is observed for times after the peak of dissipation (see dotted line,
366: Fig. \ref{FIG:COMP1}a). The fitted spectrum is $k^{+0.5}$ (note that
367: $k^{+1}$ requires the entire LANS spectrum to be resolved, and therefore has only been observed for much larger values of
368: ${k_{max}}/{k_\eta^\alpha}$).
369:
370: \begin{figure}[htbp]\begin{center}\leavevmode \centerline{%
371: \begin{tabular}{c@{\hspace{.15in}}c}
372: \includegraphics[width=8.95cm]{fig_comp1a} &
373: \includegraphics[width=8.95cm]{fig_comp1b} \end{tabular}}
374: \caption{
375: {\bf(a)} Spectra of kinetic energy \AD{(normalized to DNS $E_K(14)$, see text)} for $1536^3$ MHD DNS (solid line), $512^3$ \lamhda (dashed), and $512^3$ \lansa (dotted), in the latter case with ${\bf b}\equiv 0$ at all times but otherwise identical conditions. For intermediate scales, $k \in [5,40]$, \lamhda reproduces the
376: scaling of the DNS, the larger scales being affected by slight differences
377: in initial conditions, see text. For $k$ close to the filter scale ($k \in [k_\alpha/2,k_\alpha]$), a
378: positive power law, $k^{0.5}$ (gray line), is found for \lans.
379: {\bf(b)} Spectra of magnetic energy \AD{(normalized to DNS $E_M(14)$)} for the same runs: \lamhda reproduces the
380: scaling of the DNS even beyond the filter wavenumber, $k_\alpha = 18$ as indicated by the vertical dashed line. \lamhda exhibits neither the positive
381: power-law nor the super-filter-scale spectral contamination
382: associated with high Reynolds number \AD{\lansa} modeling seen in (a). }
383: \label{FIG:COMP1} \end{center} \end{figure}
384:
385: \resp{}{For the given parameters and initial conditions, we find the
386: super-filter-scale bottleneck for \lans.}
387: However, when integrating the MHD equations with the Lagrangian model (dashed line, Fig. \ref{FIG:COMP1}a) \resp{}{with these same parameters,} no such contamination is present.
388: Note that the spectra for the DNS-MHD
389: are shown at the time of peak dissipation, while the spectra for the
390: Lagrangian-averaged models are for a slightly later time in order to
391: allow for the possible formation of rigid bodies, which are known to be at the source of the spectral contamination close to the filter wavenumber in the Navier-Stokes case.
392: \add{}{For this reason, and due to the slight differences in initial conditions,
393: we have chosen to plot spectra normalized to that of the DNS at $k=14$ to emphasize the scaling.}
394: For most of
395: the inertial range \add{}{(also in an approximate sense below the filter
396: width $\alpha$) the scaling of $E_K(k)$ is reproduced by the \lamhda
397: simulation. The sub-filter scaling for \lamhda is not as steep as MHD, but is not a positive scaling law.} The agreement for $E_M(k)$ is remarkable. More
398: importantly, neither positive-power-law spectra nor contamination of
399: the super-filter-scale spectra are evidenced at all.
400:
401: \subsection{The lack of rigid bodies in \lamhda \ in the large$-\alpha$ limit for unforced flows}
402:
403: \resp{}{Evidence for the development of rigid bodies in \lansa (which led
404: to its limited use as a LES) has only been shown for $l\ll\alpha$ \cite{PGHM+07a}.}
405: \add{}{Since investigation of the large$-\alpha$ limit
406: is not as computationally demanding \resp{}{ as the small$-l$ limit,} it is interesting to look at this
407: limit as a rough indication of what occurs for small $\alpha$ and
408: \resp{}{smaller $l$.} This approach has been employed both for}
409: the \lansa Navier-Stokes case in two dimensions \cite{LKT+07} and \add{}{in three dimensions} \cite{PGHM+07a}.
410: In such a case, the purpose is to examine the properties of the model itself, as opposed to trying to reproduce large-scale properties, the large-scale behavior being reduced to a very small span of wavenumbers.
411: With this practice, the properties of the sub-filter-scales can be studied, to better understand the origin (or lack) of super-filter-scale contamination. We now use this
412: limit to further explore the differences between \lamhda and \lans.
413: We employ simulations for the two models with the same initial conditions as
414: before, with $\eta = \nu = 5 \cdot 10^{-5}$ ($Re \approx 26,000$ at peak
415: of dissipation for \lamhd), and a resolution of $256^3$ grid points.
416: Note that these dissipative coefficients are four times smaller than what was considered in the previous section since, for a fixed resolution, the achievable Reynolds number goes as $\alpha^{2/3}$.
417: \add{}{This follows for \lansa from the predicted (and verified) degrees of
418: freedom, $N^\alpha_{dof} \propto \alpha^{-1} Re^{3/2}$ \cite{FHT01,PGHM+07a}. The scaling of \lamhda may differ, but the same value of the viscosity is employed
419: for the two models, regardless.}
420:
421: For \lans, we observe
422: the expected $k^{+1}$ zero flux inertial range (see
423: Fig. \ref{FIG:MASK}) which is followed by a viscous (sub-filter-scale) bottleneck feature,
424: $k^{+1.5\pm.2}$, before the dissipative range proper. We conducted a second
425: simulation with $\nu = 10^{-4}$ and found a $k^{1.4\pm.3}$ spectrum.
426: This is analogous to results for DNS of the Navier-Stokes equations where
427: only the viscous bottleneck is observed at moderate Reynolds number and is preceded by
428: an inertial range only for higher Reynolds. These viscous bottlenecks \resp{}{may be} different in nature from the (super-filter-scale) bottlenecks discussed before, which are not associated to the onset of the dissipative range but to the development of a secondary inertial range in LANS below the filtering length, and result in contamination of the large (resolved) scales when the LANS equations are used as an LES.
429: Having confirmed that our analysis from the forced \lansa case extends
430: to the decaying \lansa simulation, we now apply it to \lamhd. The
431: large$-\alpha$ \lamhda spectra are given in
432: Fig. \ref{FIG:LARGEASPEC}. Notably, there is no positive-power-law
433: spectrum.
434:
435: \begin{figure}[htbp]\begin{center}\leavevmode \centerline{%
436: \includegraphics[width=8.95cm]{fig_mask_e} }
437: \caption{Spectrum of kinetic energy for a $256^3$ grid with $k_\alpha=3$ ($\nu =
438: 5\cdot10^{-5}$) \lans, ${\bf b}\equiv 0$ (Navier-Stokes case). The fitted grey
439: line, $k^{+1.1\pm.4}$, agrees with the rigid-body hypothesis for
440: the inertial range \cite{PGHM+07a}. This slope is followed by a
441: steeper slope attributed to a bottleneck, with $k^{+1.5\pm.2}$.}
442: \label{FIG:MASK} \end{center} \end{figure}
443:
444: Predictions of energy spectra in the inertial range follow from the global scaling laws for third-order structure functions
445: for isotropic, homogeneous turbulence. Exact results for these structure functions have been \resp{}{found} for incompressible MHD \cite{PP98b}
446: and for \lamhda \cite{PGHM+06}. The latter are, in terms of both the smooth fields \vier{${\bar {\bf z}}^{\pm} \equiv \vec{u}\pm\bar{\vec{b}}$} and the rough fields $\vec{z}^\pm \equiv \vec{v}\pm\vec{b}$ (where the z-fields are called the Els\"asser
447: variables):
448: \begin{equation} \left< \delta\bar{z}_\|^\mp(\vec{l}) \delta\bar{z}_i^\pm(\vec{l})
449: \delta{z}_i^\pm(\vec{l}) \right> \sim \varepsilon_\pm^\alpha l \ , \label{eq:khlamhd} \end{equation}
450: where $\left< . \right>$ denotes volume averaging, $\delta
451: f(\vec{l}) \equiv f(\vec{x}+\vec{l}) - f(\vec{x})$, and $\delta
452: f_\|(\vec{l}) \equiv \left[\vec{f}(\vec{x}+\vec{l}) -
453: \vec{f}(\vec{x})\right] \cdot \vec{l}$. For sub-filter scales ($l
454: \ll \alpha$), $\bar{z}^\pm \sim l^2\alpha^{-2}{z}^\pm$ and the scaling law becomes dimensionally
455: $\bar{z}z\bar{z} \sim \varepsilon l$. This implies a sub-filter scale
456: spectrum
457: corresponding to the invariants $E_{\pm}^\alpha \equiv ||\bar{\vec{z}}^\pm||^2_\alpha/2$ for the ideal non-dissipative case. We then have
458: $E^\alpha_\pm(l)k \sim z^\pm\bar{z}^\pm \sim
459: (\varepsilon_\pm^\alpha)^{2/3}\alpha^{2/3}$ or, equivalently,
460: \begin{equation}
461: E^\alpha_\pm(k) \sim (\varepsilon_\pm^\alpha)^{2/3}\alpha^{2/3} k^{-1}
462: \label{eq:speclamhd}
463: \end{equation}
464: as for \lansa \cite{FHT01}.
465: Recall that in the flux relation,
466: Eq. (\ref{eq:khlamhd}), $\varepsilon_\pm^\alpha$ stands for the energy
467: transfer and dissipation rate of $E_\pm^\alpha$.
468: Hence, the prediction,
469: Eq. (\ref{eq:speclamhd}), for the spectra, $E^\alpha_\pm(k)$, is,
470: equivalently for $E_T^\alpha \equiv (||u||^2_\alpha+||b||^2_\alpha)/2$ and for $H_C^\alpha
471: \equiv \frac{1}{2} \frac{1}{D} \int_D \vec{v} \cdot \bar{\vec{b}}~d^3x$. The spectra shown in
472: Fig. \ref{FIG:LARGEASPEC} for large$-\alpha$ \lamhda do not exclude,
473: due to the large uncertainties of the fitted power laws, the predicted
474: $k^{-1}$ spectra. %Another possibility is that the observed power laws
475: %are viscous (sub-filter-scale) bottleneck features of an, as yet, unresolved inertial range.
476: %Only simulations at higher resolution can answer this possibility.
477:
478: \begin{figure}[htbp]\begin{center}\leavevmode \centerline{%
479: \includegraphics[width=8.95cm]{fig_mask_mhd}
480: }
481: \caption{Spectra for a $256^3$ grid with $k_\alpha=3$ ($\eta=\nu=5\cdot10^{-5}$)
482: \lamhd, $Re \approx 26,000$: Total energy, $E_T(k)$, (solid line)
483: and cross helicity, $H_C(k)$, (dashed). The fitted slopes,
484: $E_T(k)\sim k^{-0.7\pm.3}$ and $H_C (k)\sim k^{-0.5\pm.4}$ could agree
485: with either Kolmogorov or IK predictions for \lamhda (see text) at this
486: level of uncertainty.}
487: \label{FIG:LARGEASPEC} \end{center} \end{figure}
488:
489: A spectral prediction for \lamhda can also be arrived at by
490: dimensional analysis of the spectrum which follows the scaling ideas
491: originally due to Kraichnan \cite{K67} and which is developed
492: for \lansa in Ref. \cite{CHT05}. Here, the energy dissipation
493: rate, \add{}{$\varepsilon^\alpha_\pm = dE^\alpha_\pm/dt$,} is related to the spectral energy
494: density by
495: \begin{equation}
496: \varepsilon^\alpha_\pm \sim (\mathfrak{t}_k)^{-1} \int E^\alpha_\pm(k)
497: \label{eq:scaling}
498: \end{equation}
499: where $\mathfrak{t}_k$ is the turnover time for an eddy of size
500: $\sim k^{-1}$. This turnover time is related to a ``velocity,''
501: $\bar{Z}_k^\pm$, (i.e., $\mathfrak{t}_k \sim 1 / (k\bar{Z}_k^\pm)$), where
502: $(\bar{Z}_k^\pm)^2 \sim \bar{Z}_k^\pm{Z}_k^\pm/(1+\alpha^2k^2) \sim k
503: E^\alpha_\pm(k)/(1+\alpha^2k^2)$. Substitution into
504: Eq. (\ref{eq:scaling}) yields,
505: \begin{equation}
506: E^\alpha_\pm(k)
507: \sim (\varepsilon^\alpha_\pm)^{2/3} k^{-5/3}(1+\alpha^2k^2)^{1/3}
508: \label{eq:oldK41}
509: \end{equation}
510: or, for $\alpha k \gg 1$,
511: \add{}{\begin{equation}
512: E^\alpha_\pm(k)
513: \sim (\varepsilon^\alpha_\pm\alpha)^{2/3} k^{-1}.
514: \end{equation}}
515: In the Iroshnikov-Kraichnan \cite{I64,K65} (hereafter, IK) phenomenology, Alfv\'en waves
516: (corresponding to either $z^\mp = 0$)
517: can only
518: interact nonlinearly when they collide along field lines (along which they travel
519: in opposite directions). The characteristic time for an Alfv\'en wave
520: is $\mathfrak{t}_A \sim (kB_0)^{-1}$. If this is less than
521: $\mathfrak{t}_k$, the effective transfer time $\mathfrak{t}_T$ is
522: increased, $\mathfrak{t}_T \sim \mathfrak{t}_k^2 /
523: \mathfrak{t}_A$. Substitution of this new transfer time into
524: Eq. (\ref{eq:scaling}) yields, \add{}{instead of Eq. (\ref{eq:oldK41})
525: \begin{equation}
526: E^\alpha_\pm(k)
527: \sim (\varepsilon^\alpha_\pm B_0)^{1/2} k^{-3/2}(1+\alpha^2k^2)^{1/2}
528: \end{equation}
529: or, for $\alpha k \gg 1$,
530: \begin{equation}
531: E^\alpha_\pm(k)
532: \sim (\varepsilon^\alpha_\pm B_0)^{1/2} \alpha k^{-1/2}.
533: \label{eq:IK}
534: \end{equation}
535: } The spectra shown in Fig. \ref{FIG:LARGEASPEC} for large$-\alpha$
536: \lamhda also agree with the IK predicted spectra, Eq. (\ref{eq:IK}).
537: In fact, the spectra more closely correspond to this prediction;
538: this is consistent with the fact that, for this flow, an IK spectrum $E(k)\sim k^{-3/2}$ is observed at large scale (followed by a weak turbulence anisotropic spectrum $E(k_{\perp})\sim k_{\perp}^{-2}$ at small scale) \cite{MiPo2007a}.
539: Again, simulations at higher resolution are needed for a definite answer and
540: the result may not be universal
541: as shown for example in the context of reduced MHD dynamics due to the presence of a strong uniform magnetic field
542: ${\bf B}_0$ \cite{DmGoMa2003} or for MHD with a strong ${\bf B}_0$ \cite{MaCaBo2008}.
543:
544: \begin{figure}[htbp]\begin{center}\leavevmode
545: \centerline{%
546: \includegraphics[width=8.95cm]{fig_pdfs}
547: }
548: \caption{PDFs of cubed increments. \AD{The cubed increments when
549: averaged are equal to flux times length, $\varepsilon^\alpha
550: \cdot l$.} Here $l = 0.88\alpha$ ($\alpha=2\pi/3$). The dotted
551: line is \AD{$\delta u_\|(l) \delta u_i(l) \delta v_i(l)$} for
552: \lans, solid for \lamhda \AD{$\delta\bar{z}_\|^-(\vec{l})
553: \delta\bar{z}_i^+(\vec{l}) \delta{z}_i^+(\vec{l})$}, and dashed
554: for \lamhda \AD{$\delta\bar{z}_\|^+(\vec{l})
555: \delta\bar{z}_i^-(\vec{l}) \delta{z}_i^-(\vec{l})$.} More of
556: the volume gives no contribution to the flux for \lansa than for
557: \lamhd, indicating no rigid bodies in \lamhd. }
558: \label{FIG:PDFS}
559: \end{center}
560: \end{figure}
561:
562:
563: Another indication of the zero-flux regions in \lansa is \resp{}{found} by
564: examining the spatial variation of the cubed increments associated
565: with the scaling laws
566: $\delta u_\|(l) \delta u_i(l) \delta v_i(l)$ for \lansa and
567: $\delta\bar{z}_\|^\mp(\vec{l}) \delta\bar{z}_i^\pm(\vec{l})
568: \delta{z}_i^\pm(\vec{l})$ for \lamhda
569: (note that one can transform this relation into the $u,\ v,\ b,\ {\bar b}$ variables).
570: \add{}{For a given length $l$, these cubed increments when averaged are related
571: with the energy fluxes by Eq. (\ref{eq:khlamhd}) (the \lansa relation
572: and the hydrodynamic and MHD relations are contained in this expression in the corresponding limits). As a result of this correspondence, for brevity we will
573: indicate cubed increments in the figures as the corresponding energy flux times
574: the length used to compute the increments. This also allows us to identify regions with zero cubed increments as rigid bodies (a rigid rotation has zero longitudinal increments).}
575: Probability distribution
576: functions (PDFs), see Fig. \ref{FIG:PDFS}, indicate that \lamhda has a
577: much smaller proportion of its volume, which could potentially be rigid
578: bodies (i.e., frozen regions with no internal degrees of freedom \add{}{(zero velocity increment)}, which therefore do not contribute to the energy flux). That is, more of the volume is contributing to the turbulent
579: cascade. Snapshots for constructing the PDFs are taken from both $\alpha=2\pi/3$
580: Lagrangian-averaged models for times shortly after the peak of
581: dissipation and when the \lansa total dissipation is nearly equal to
582: that of \lamhd. The strengths of the central peaks of the PDFs for
583: large$-\alpha$ are another indication that \lamhda inherits none of
584: the rigid-body or zero-flux-region problems of \lans.
585:
586: \subsection{Why are spectral properties of \lamhd\ better than in the fluid case?}
587:
588: Why does \lamhda not exhibit the same spectral contamination as \lans?
589: %The first difference between the two models is seen by casting them in LES
590: %form. \lamhda leads to a
591: %for the divergence of the turbulent
592: %electro-motive force (emf) stress tensor, %, $\alpha^2\nabla^4\vec{\bar{b}}$,
593: \resp{}{One possible cause is the
594: hyperdiffusivity term
595: seen in the LES form for \lamhd,
596: Eq. (\ref{eq:emf}),
597: whereas} there is no hyperviscosity-like term in \lans. To test if this hyperdiffusion
598: is responsible for the lack of spectral contamination in \lamhd, we
599: removed the
600: hyperdiffusion by \add{}{setting $\bar{\tau}^b=0$ in Eqs. (\ref{eq:lamhdLES}) or,
601: equivalently, by} substituting $\eta \nabla^2\bar{\vec{b}}$ for $\eta
602: \nabla^2\vec{b}$ in Eqs. (\ref{eq:lamhd}).
603: We then start the run from the same initial conditions but now with these new equations \resp{}{employing} $\alpha = 2\pi/33$ \add{}{and} $\nu=\eta=2\cdot10^{-4}$ at a
604: resolution of $384^3$ (with hyperdiffusion, a smaller resolution of
605: $256^3$ is possible, see Section \ref{SEC:LES}).
606: Note that
607: such a modified \lamhda model is not expected to, nor found to, perform well as a SGS model;
608: this numerical experiment is performed here only in order to assess the effect of the hyper-diffusive term introduced by the $\alpha$ modeling.
609: We find that hyperdiffusion
610: is {\sl not} responsible for the lack of a $k^{+1}$ spectral
611: contamination in \lamhda (see Fig. \ref{FIG:HYPER}).
612:
613: \begin{figure}[htbp]\begin{center}\leavevmode
614: \centerline{%
615: \includegraphics[width=8.95cm]{fig_hyper}
616: }
617: \caption{Spectra for a $384^3$ grid with $k_\alpha=33$ obtained from the modified-\lamhd \ (see text) shortly after the maximum of dissipation: kinetic
618: energy (solid) and magnetic energy (dashed); the \lamhda
619: equations have been modified by removing the hyperdiffusive
620: turbulent emf. Even without hyperdiffusivity, no positive
621: power-law is found. Instead, fits (grey lines) for kinetic and
622: magnetic energy spectra near the filtering length are $k^{-1.7\pm.1}$
623: and $k^{-1.9\pm.1}$, respectively.}
624: \label{FIG:HYPER}
625: \end{center}
626: \end{figure}
627:
628: \resp{}{Other possible causes for \lamhda not exhibiting the
629: super-filter-scale bottleneck as does \lansa are the} actual
630: physical differences between the two fluids \resp{}{that are modeled,}
631: Navier-Stokes and MHD. First, unlike incompressible Navier-Stokes,
632: MHD supports oscillatory solutions (Alfv\'en waves) which are linked
633: to enhanced spectral nonlocality of energy transfer \cite{AMP05a,
634: Al2007a} leading to dynamic interactions between widely separated
635: scales. \resp{}{ For Navier-Stokes, the depletion of energy transfer
636: due to local interactions at some cutoff in wavenumber is believed
637: to bring about the bottleneck effect
638: \cite{HSL+82,LMG95,MCD+97,MAP06}. However, related to the
639: spectrally nonlocal energy transfer via Alfv\'en waves,} MHD does
640: not seem to exhibit a bottleneck in its spectra between the inertial
641: and dissipative ranges \cite{MiPo2007a}. \resp{}{As \lamhda supports
642: Alfv\'en waves at all scales (and alters their dissipation \vier{and wave speed}
643: appreciably only for sub-filter scales), the same physics could be
644: behind the lack of a super-filter-scale bottleneck in \lamhd.}
645:
646: Another difference between the fluid and MHD cases is the geometry of
647: the dissipative structures: one finds vortex filaments for
648: Navier-Stokes at high value of the vorticity, and current and
649: vorticity sheets for MHD; sheets which are found to roll-up at high
650: Reynolds number \cite{MiPoMo2006}. \resp{}{It has been claimed that
651: the development of helical filaments in the fluid case can lead to
652: the depletion of nonlinearity and the quenching of local
653: interactions \cite{MT92,T01} and, hence, to the viscous bottleneck.
654: A similar energy transfer depletion may occur in \lans.} In
655: \cite{PGHM+07a} \resp{}{evidence is presented} that Taylor's frozen-in turbulence
656: hypothesis applied to Lagrangian averages leads to the formation of
657: ``rigid bodies'' in the flow wherein there are no internal degrees of
658: freedom and no transfer of energy to smaller scales (i.e. regions with
659: $\varepsilon \sim \delta u_\|^3/l = 0$ as well as
660: $\boldsymbol{\omega}\times\vec{v} = 0$). These regions are likely
661: related to the shorter, thicker vortex filaments formed and the
662: suppression of vortex stretching dynamics as $\alpha$ is increased
663: \cite{CHM+99}. As MHD has spectrally non-local transfer (e.g.,
664: velocity at large scales does stretching of magnetic field lines at
665: small scales) \resp{}{this leads to the break up of these rigid bodies
666: in the \lamhda case and the breakup of the viscous bottleneck
667: in the MHD case. The magnetic field} interaction with the large
668: scale \resp{}{velocity can} re-enable transfer of energy to smaller
669: scales \resp{}{of the velocity field.} Indeed, defining the kinetic
670: spectral transfer due to the Lorentz force as
671: \begin{equation}
672: T^\alpha_L(k) \equiv \int \hat{\vec{u}}_k \cdot \left( \widehat{\vec{j} \times \bar{\vec{b}}} \right)^*_k d\Omega_k
673: \label{eq:transfer_alpha}
674: \end{equation}
675: \resp{}{for \lamhda, and as
676: \begin{equation}
677: T_L(k) \equiv \int \hat{\vec{v}}_k \cdot \left( \widehat{\vec{j} \times {\vec{b}}} \right)^*_k d\Omega_k
678: \label{eq:transfer}
679: \end{equation}
680: for MHD,} we see in Fig. \ref{FIG:TRANSFER} that the Lorentz force is
681: removing large-scale kinetic energy and supplying small-scale kinetic
682: energy; this \resp{}{effectively bypasses} the formation of rigid bodies
683: \resp{}{for \lamhda and the viscous bottleneck for MHD (note that
684: Eqs. (\ref{eq:transfer_alpha}) and (\ref{eq:transfer}) do not detail
685: the scales at which magnetic energy is created or destroyed).}
686:
687: \begin{figure}[htbp]\begin{center}\leavevmode \centerline{%
688: \includegraphics[width=8.95cm]{fig_transfer}}
689: \caption{(Color online) Spectral transfer due to the Lorentz force, \neu{$T_L$ (for $1546^3$ DNS)} and $T_L^\alpha$ (for $512^3$ $k_\alpha=18$
690: \lamhd) \neu{at a time just prior to the peak of dissipation.} \neu{Positive $T_L$ is shown as dash-dotted lines and
691: negative $T_L$ as dashed lines.} Positive $T_L^\alpha$ is shown as solid (green online) lines and
692: negative $T_L^\alpha$ as dotted (green online) lines. \neu{\lamhda qualitatively reproduces the transfer of kinetic energy in MHD.}}
693: \label{FIG:TRANSFER} \end{center} \end{figure}
694:
695: This \add{}{argument can also be recast} in terms of Kelvin's circulation theorem. \resp{}{For Navier-Stokes,} the circulation $\Gamma$ of the velocity \resp{}{$\vec{v}$} is conserved in the ideal case for barotropic flows. In ideal \resp{}{MHD,} this conservation is broken by the Lorentz force,
696: \resp{}{\begin{equation}
697: \frac{d \Gamma}{dt} = \frac{d}{dt} \oint_{\cal C} \vec{v}\cdot d\vec{r}
698: = \oint_{\cal C} \vec{j} \times {\vec{b}} \cdot d\vec{r},
699: \label{eq:kelvin}
700: \end{equation}
701: } where ${\cal C}$ is any material curve. As a result, while in ideal \resp{}{Navier-Stokes} a material curve ${\cal C}$ defines the boundary of a vorticity tube with fixed strength, in \resp{}{MHD} these structures are deformed and \resp{}{their} vorticity content changed by the Lorentz force.
702: \resp{}{A similar result follows for \lamhda and \lans,}
703: \begin{equation}
704: \frac{d \Gamma}{dt} = \frac{d}{dt} \oint_{\cal C} \vec{u}\cdot d\vec{r}
705: = \oint_{\cal C} \vec{j} \times \bar{\vec{b}} \cdot d\vec{r}\,.
706: \end{equation}
707: \resp{}{Breaking the conservation of circulation in this way can
708: prevent the formation of a bottleneck. For example, for the fluid
709: case in the Clark$-\alpha$ model (which differs from \lansa only in
710: the conservation of $\Gamma$), it was also found that no
711: super-filter-scale bottleneck was present \cite{PiGrHoMi+2008}.}
712:
713:
714: \subsection{LES Application}
715: \label{SEC:LES}
716:
717:
718: \begin{figure}[htbp]\begin{center}\leavevmode
719: \centerline{%
720: \begin{tabular}{c@{\hspace{.15in}}c}
721: \includegraphics[width=8.95cm]{figEvsT} &
722: \includegraphics[width=8.95cm]{figEvsTb}
723: \end{tabular}}
724: \caption{(Color online) Temporal evolution, $\tau_{eddy} \approx 4.5$, for $1536^3$ DNS (solid,
725: black), $256^3$ $k_\alpha = 33$ \lamhda (dashed, green online),
726: $256^3$ under-resolved ``DNS'' (dotted, red online),
727: \neu{and $384^3$ $k_\alpha = 33$ nonhyperdiffusive-\lamhda (dash-dotted, blue online).}
728: {\bf (a)} Time evolution of the energies: kinetic (lower curves),
729: magnetic (middle curves) and total (upper curves). {\bf (b)} Time
730: evolution of total enstrophy, $\left<j^2+\omega^2\right>$
731: ($\left<j^2+\boldsymbol{\omega}\cdot\bar{\boldsymbol{\omega}}\right>$ for \lamhda \neu{and $\left<\vec{j}\cdot\bar{\vec{j}}+\boldsymbol{\omega}\cdot\bar{\boldsymbol{\omega}}\right>$ for the nonhyperdiffusive case).} Note that \lamhda
732: gives a better agreement to the total dissipation rate up to the maximum time that the high resolution DNS is performed.
733: Also note that the DNS equivalent to the \lamhda run presented here is not feasible on present-day computers at a reasonable cost.}
734: \label{FIG:EVST} \end{center} \end{figure}
735:
736: Having now shown that \lamhda does not suffer the same drawbacks with
737: regards to energy spectra as \lans, we may turn our attention to \resp{}{a
738: practical application.} The purpose of a SGS model or LES is to make predictions
739: about large Reynolds number flows at a reduced computational expense.
740: From the scaling arguments in Refs. \cite{FHT01,PGHM+07a}, using
741: simulations conducted at $Re \approx 2200$, and assuming a
742: $k^{-1}$ scaling, we can estimate $\alpha=1/33$ for a $256^3$
743: \lamhd-LES ``prediction'' of our $1536^3$ MHD-DNS. Time evolution of
744: the energies and the total enstrophy are shown in Fig. \ref{FIG:EVST}
745: for much later times than reasonably attainable with the MHD DNS with present-day computers. Also shown are
746: results for solving the MHD equations, Eqs. (\ref{eq:mhd})
747: with $\nu =2 \cdot 10^{-4}$ and a resolution of $256^3$: a so-called ``unresolved
748: DNS'' \resp{}{and the non-hyperdiffusive modified-\lamhda from the previous section.} Before the peak of dissipation, $t\approx4$, the unresolved
749: DNS gives a poorer prediction of the total dissipation and total
750: energy which is then followed by a
751: significantly larger and
752: somewhat later peak of
753: dissipation, at $t\approx5$ than the resolved DNS and the \lamhda LES.
754: \resp{}{The non-hyperdiffusive \lamhda is not expected to perform well as a
755: SGS model and it is seen to be clearly under-dissipative. The ratio of magnetic
756: to kinetic dissipation is $\approx1.5$ for the DNS, $\approx2.9$ for \lamhd, $\approx1.1$ for the under-resolved DNS, and $1.4$ for the non-hyperdiffusive model.} \resp{}{Together with Fig. \ref{FIG:EVST} (b) these ratios show that
757: \lamhda achieves accurate total dissipation by an excess of magnetic dissipation and a reduction of kinetic dissipation (both at the small scales). This
758: feature has already been depicted in Fig. 15 of Ref. \cite{MMP05a}.}
759: Compensated energy spectra for \resp{}{the peak of dissipation} ($t\in[2.7,3.7]$) are shown in
760: Fig. \ref{FIG:COMP2}. For the under-resolved DNS, \resp{}{we observe} the appearance of a tail at large wavenumbers with a $k^2$ spectrum as predicted using statistical mechanics arguments for truncated systems in the ideal ($\nu=0$, $\eta=0$) case \cite{FrPoLe+1975}.
761: The under-resolved spectra are not significantly different from the resolved DNS, but note that a reliable and convincing determination of spectral indices, beyond visual inspection, does require high resolutions.
762: Comparing now the resolved DNS and the \lamhda run, the quality of the spectra are similar for scales
763: larger than $\alpha$. Recall that differences at the largest scales, stem from the
764: differences in initial conditions as stated in
765: Section \ref{SEC:LAMHDGOOD}, and from time evolution of the flow.
766: Finally, noting that the computer saving here is $6^3$ in memory and $6^4$ in running time, we conclude that the \lamhda continues to behave satisfactorily, as already shown both in two space dimensions
767: \cite{MMP05a,PGMP05,PGHM+06}
768: and in 3D \cite{MMP05b},
769: in particular in the context of the dynamo problem of generation of magnetic fields by velocity gradients; thus, \lamhda may prove to be a useful tool in many astrophysical contexts where magnetic fields are dynamically important, such as in the solar and terrestrial environments, or in the interstellar and intergalactic media.
770:
771: %Given this demonstration of the spectral properties of \lamhda as a LES with a decrease in linear resolution of a factor of $6$ and \note{ considering previous tests in Pablo's and our papers, dynamo run, etc.}, we can confidently recommend continued use of this model.
772:
773: \begin{figure}[htbp]\begin{center}\leavevmode
774: \centerline{%
775: \begin{tabular}{c@{\hspace{.15in}}c}
776: \includegraphics[width=8.95cm]{fig_comp2a} &
777: \includegraphics[width=8.95cm]{fig_comp2b}
778: \end{tabular}}
779: \caption{
780: Spectra compensated by $k^{3/2}$ for the kinetic {\bf (a)} and magnetic {\bf (b)} energies \neu{averaged over $t \in [2.7,3.7]$;} labels are as in Fig. \ref{FIG:EVST} and the dashed vertical line indicates
781: $k_\alpha=33$. \neu{Note}
782: the $k^2$ tail at high wavenumber that is known to develop for under-resolved runs, a prediction stemming from statistical mechanics.
783: }
784: \label{FIG:COMP2} \end{center} \end{figure}
785:
786: \vier{We also computed a $512^3$ \lamhd-LES ($\alpha=1/85$) which retains
787: more of the small scales than the $256^3$ \lamhd-LES while still
788: yielding significant computational savings over the $1536^3$ DNS. We
789: compare this with the result for $\alpha=1/18$ (chosen not as a LES but
790: to stress the model) in Fig. \ref{FIG:SHEETS}. The structure of
791: sheets observed in MHD dissipative structures is preserved in the
792: \lamhda simulations, although current and vortex sheets become thicker
793: in \lamhda as a result of the filter as $\alpha$ is increased. This
794: is necessary to achieve reduced resolution computations. Note that
795: these sheets are different in nature from the fat 'rigid bodies'
796: observed in \lans, as the turbulent energy transfer to small scales is
797: not quenched and there is no super-filter-scale bottleneck.}
798:
799: \begin{figure}[htbp]\begin{center}\leavevmode
800: \centerline{%
801: \begin{tabular}{c@{\hspace{.15in}}c}
802: \includegraphics[width=8.95cm]{fig_current85} &
803: \includegraphics[width=8.95cm]{fig_current18} \\
804: \includegraphics[width=8.95cm]{fig_current33} &
805: \includegraphics[width=8.95cm]{fig_currentNS} \\
806: \end{tabular}}
807: \caption{\vier{2D cross sections of square current, $j^2$, for
808: $512^3$ \lamhd-LES ($\alpha=1/85$) {\bf (Upper Left)} and
809: model-stress-case ($\alpha=1/18$) {\bf (Upper Right)}. MHD
810: dissipative structures, sheets, are retained which become
811: thicker as $\alpha$ is increased. {\bf (Lower Left)} $256^3$
812: \lamhd-LES ($\alpha=1/33$) and {\bf (Lower Right)} $256^3$
813: unresolved DNS. For the unresolved run, current sheets are
814: somewhat smeared out by numerical noise.}}
815: \label{FIG:SHEETS} \end{center} \end{figure}
816:
817: \section{Discussion}
818:
819:
820: In this paper, we have tested the \lamhda model against high Reynolds
821: number direct numerical simulations (up to \resp{}{Reynolds numbers of
822: $\approx 9200$)} and in particular we have focused our attention on
823: the dynamics of small scales near the $\alpha$ cut-off. We find that
824: the small-scale spectrum presents no particular defect; specifically,
825: we find that, unlike in the hydrodynamical case, the
826: Lagrangian-averaged modeling for MHD exhibits, even at large Reynolds
827: numbers, neither a positive-power-law spectrum nor any contamination
828: of the super-filter-scale spectral properties. \resp{}{This
829: difference between \lansa and \lamhda} is not due to the inclusion
830: of a hyper-diffusive term in \lamhda that stems from the derivation of
831: the model; rather, it stems from fundamental differences between
832: hydrodynamics and MHD. Indeed, neither the (non-consistent) removal of
833: hyperdiffusion from \lamhda nor the examination of scales much smaller
834: than $\alpha$ gave any indication of problems similar to those caused
835: by the zero-flux regions found in computations using \lans. These
836: regions limited the computational gains of using \lansa as a LES in
837: hydrodynamics to a factor of only $10$ in computational degrees of
838: freedom or $30$ in computation time. \lamhda is not subject to the
839: same limitations and, as we demonstrated, a gain of a factor of $200$
840: in the number of degrees of freedom, or a factor of $1300$ in computation time, obtains when comparing to
841: the highest Reynolds number in turbulent MHD available today in a DNS.
842:
843: There are two obvious candidates to explain the lack of a
844: (super-filter-scale) bottleneck effect in \lamhd: the enhanced
845: (hyper-)diffusion in \lamhda compared with \lans, and \resp{}{physical
846: differences between fluids and magneto-fluids, specifically,
847: spectrally nonlocal transfer via Alfv\'en waves and its associated} breaking of
848: the circulation conservation. The first candidate would eliminate the
849: super-filter-scale bottleneck by removing energy from the system and
850: precluding the formation of a secondary range below the filtering
851: scale $\alpha$ (note that this term becomes of the same order as the
852: ordinary diffusion when $l\sim\alpha$). Simulations of \lamhda
853: performed without the hyper-diffusion term \resp{}{ruled out} this
854: scenario, as no super-filter bottleneck was found.
855:
856:
857: The second candidate is the \resp{}{presence of the Lorentz force in
858: MHD (and \lamhd) which} breaks down the circulation conservation
859: \resp{}{and provides the restoring force for Alfv\'en waves. Both
860: properties were shown to be preserved by \lamhd. In Navier-Stokes,
861: the development of helical filaments could quench local interactions
862: \cite{MT92,T01} depleting the energy transfer and leading to the
863: viscous bottleneck. However, in MHD, the conservation of the
864: circulation ($d\Gamma/dt=0$ in the absence of dissipation) is broken
865: by the Lorentz force, which modifies Kelvin's theorem (see
866: Eq. (\ref{eq:kelvin})). The forcing term is associated with the
867: Alfv\'en waves, and represents the removal of circulation (and of
868: kinetic energy) that is transfered to the magnetic field. Note that
869: in Fourier space, the term scales as $k E_M(k)$ and is dominant compared to
870: the dissipation in the inertial range. This term precludes the
871: formation of rigid bodies, giving as a result a larger net flux
872: towards smaller scales and a resulting larger dissipation in
873: MHD/\lamhd. This is illustrated in Fig. \ref{FIG:PDFS}. This sink of
874: circulation may also be the cause of the lack of a viscous-scale
875: bottleneck in MHD. In LANS it was shown
876: \cite{PGHM+07a,PiGrHoMi+2008} that conservation of the circulation
877: (except for viscosity) leads to the formation of rigid bodies that
878: fill a substantial volume of the fluid, and that in turn
879: substantially decrease the energy flux to small scales, reduce
880: dissipation, and create the super-filter scale bottleneck.} In
881: \lamhd, the destruction of sub-filter-scale rigid bodies by large
882: scale magnetic field and shear \resp{}{results} as the presence of a
883: magnetic field permits the development of long-range interactions in
884: spectral space \cite{MAP05,AMP05a,Al2007a}. This can also explain why
885: $\alpha-$models for other non-local equations, or for problems that do
886: not preserve the circulation provide good SGS models. As an example,
887: the use of \lansa in primitive equations ocean modeling gives
888: satisfactory results, e.g. in its reproducing the Antarctic
889: circumpolar current baroclinic instability that can be seen only at
890: substantially higher resolutions when using direct numerical
891: simulations \cite{HeHoPe+2008}.
892:
893: \resp{}{Energy is dissipated in MHD flows through two different
894: processes. Viscosity is responsible for the dissipation of
895: mechanical energy, while Ohmic losses are responsible for
896: dissipation of magnetic energy. Mechanical and magnetic energy are
897: not conserved separately, but rather coupled as illustrated by the
898: existence of Alfv\'en waves, which correspond to oscillations of the
899: magnetofluid with the velocity field parallel or anti-parallel to
900: the magnetic field, and associated to the interchange of magnetic
901: and kinetic energy. In MHD, it is believed that most of the total
902: energy in the flow is finally dissipated (mediated by this
903: interchange) through Ohmic losses, in a process that involves
904: reconnection of magnetic field lines. This is supported by several
905: simulations of MHD turbulence \cite{HaBrDo2003,Mi2007a} and is consistent
906: with phenomenology. While in hydrodynamics small scales are
907: permeated by a myriad of vortex filaments, in MHD the dominant
908: dissipative structures are current sheets, where strong gradients of
909: the magnetic field and their associated strong currents lead to
910: rapid Ohmic dissipation. Sub-grid models attempt to replace the
911: physical processes of small-scale dissipation by processes that
912: mimic the non-linear transfer of energy to smaller scales (where
913: energy is in reality dissipated, but now in scales that are not
914: resolved by the model). In traditional LES, this is done with
915: enhanced turbulent viscosities. Note that the eddy viscosity is not
916: obtained from the linear dissipative term (the term that describes
917: the actual physical process responsible for the dissipation) but
918: from the non-linear terms in the equations (the terms that describe
919: the coupling between fields at different scales). The final goal is
920: not to capture the dissipation processes, but to be able to preserve
921: (with computational gains) the large scale dynamics.}
922:
923: \resp{}{Lagrangian averaged models take a different (although related,
924: see e.g., \cite{PGHM+06}) approach. Besides adding (in some cases,
925: as in the case of MHD) an enhanced viscosity, the non-linear terms
926: are modified at small scales. This modification changes the
927: time-scale of the energy cascade, and as a result changes the
928: scaling law of the energy spectrum $E(k)$ at sub-filter scales. This
929: change leads to changes in the dissipation, as the dissipation is in
930: the original equations proportional to $k^2E(k)$. The end result (an
931: enhanced dissipation that is intended to mimic the transfer of
932: energy to smaller scales in the unresolved scales) should be the
933: same as in a traditional LES: gains in computing costs preserving as
934: much information of the large scale flow as possible. As in the
935: case of LES, the actual dissipation process is not as important as
936: the fact that large-scale dynamics should be reproduced with minimal
937: contamination \vier{by} the sub-grid model. We believe the results
938: presented here (and in earlier work
939: \cite{MMP05a,MMP05b,PGMP05,PGHM+06,Mi2006a}) show this is the case,
940: and allow the use of the LAMHD equations as a subgrid model of MHD
941: turbulence. However, considering the differences observed between
942: LANS and LAMHD, we discuss the dissipation processes in LAMHD. Two
943: mechanisms for dissipation can be identified in LAMHD: dissipation
944: of mechanical energy through the viscosity, and dissipation of
945: magnetic energy through (enhanced) Ohmic losses. From the equations,
946: the total variation of energy goes as \cite{MMP05a}: $dE/dt =
947: -\nu\left<\boldsymbol{\omega}\cdot\bar{\boldsymbol{\omega}}\right>
948: -\eta\left<j^2\right>$ and as a result the mechanical energy
949: dissipation scales as $k^2E_V(k)$ while the magnetic energy
950: dissipation scales as $(1+ \alpha^2k^2)k^2E_M(k)$. The extra $k^2$
951: factor in the latter gives more dissipation than in the LANS
952: case. This excess of magnetic dissipation in LAMHD mimics, as
953: previously mentioned, the dominant contribution to dissipation by
954: Ohmic losses in MHD. This hyperdiffusion is required in the
955: sub-filter scales to accurately model the total energy dissipated at
956: the unresolved scales. This was demonstrated by our experiments
957: with a modified \lamhd, where we (non-consistently) removed the
958: hyperdiffusive term and found the resulting model to fail as a LES.}
959:
960: \resp{}{Yet} another way to understand the differences between \lansa
961: (for incompressible isotropic and homogeneous flows) and \lamhda is to
962: consider the derivation of these models \cite{H02a} using the
963: generalized Lagrangian-mean (GLM) formalism \cite{AM78}. This form of
964: Lagrangian averaging describes wave, mean-flow interactions. For the
965: case of weak turbulence, where the nonlinear transfer is dominated by
966: waves, GLM requires in principle no closure. As a result, GLM gives an
967: exact closed theory for the evolution of the wave activity. On the
968: other hand, when there are no waves (as in incompressible
969: Navier-Stokes) or when eddies dominate the transfer, a closure is
970: required. One possible closure assumes that fast fluctuations are
971: just advected by the mean flow (basically, Taylor's frozen-in
972: hypothesis for the small scale turbulent fluctuations) and leads to
973: the several "$\alpha$-models" that include \lansa and \lamhd. In this
974: context, it is not surprising for subgrid models based on GLM to
975: perform better in the presence of Alfv\'en waves (for \lamhd) or
976: Rossby and gravity waves (for the Lagrangian-averaged primitive
977: equations \cite{HeHoPe+2008}). The more relevant the waves are to the
978: dynamics, and to the non-linear coupling of modes in the system, the
979: less relevant is the hypothesis behind the closure. Furthermore, the
980: $\alpha$-model equations can then be expected to be a better
981: approximation to the problem at hand, that is, to be closer to an
982: exact closure of the original system of equations.
983:
984: \vier{In the fluid case, the application of the ``Taylor'' closure
985: that smaller-than-$\alpha$ scale fluctuations are swept along by the
986: large-scale flow results in the fluctuations having greatly reduced
987: interactions. This allows for a reduction in computational expense
988: and leads to the super-filter-scale bottleneck by quenching
989: spectrally non-local interactions. In the \lamhda case, the
990: small-scale $\vec{z}^+$ ($\vec{z}^-$) fluctuations are swept along
991: by the large-scale $\bar{\vec{z}}^-$ ($\bar{\vec{z}}^+$) flow.
992: Small-scale fluctuations advected by two different fields may now
993: collide and nonlinearly interact. The second part of the model is
994: the preferential hyperdiffusion of Alfv\'en waves
995: with wavelengths shorter than $\alpha$. This damps rather than
996: quenches nonlinear interactions among the small scales. This more
997: gentle suppression of the transfer of energy to smaller scales
998: reduces the numerical resolution requirements without forming a
999: bottleneck.}
1000:
1001: It was noted in \cite{MMP05b} when assessing the properties of \lamhda in the dynamo context that the overall temporal evolution was satisfactory, e.g. with a correct growth rate, although
1002: the growth of the magnetic seed field started slightly earlier in the \lamhda run than in the DNS. One can speculate as to whether this delay is linked to the super-bottleneck effect of \lansa (which prevails when the magnetic field is negligible compared to the velocity, the two modeling approaches, \lamhda and \lans, being dynamically consistent). This point is left for future work; one could determine as well at what ratio of magnetic to
1003: kinetic energy the overshooting of spectra in \lansa disappears for \lamhd.
1004:
1005: Also deserving of a separate study is to investigate the behavior of \lamhda when anisotropies that appear at small scales \cite{MiPo2007a} are present; this would be essential when a uniform magnetic field is imposed to the overall flow. The evaluation of the behavior of the model when computing spectra in the perpendicular and parallel directions (with respect to a quasi-uniform magnetic field, computed by locally averaging the field in a sphere of radius comparable to the integral scale) remains to be done but is somewhat time consuming.
1006: An analysis of the structures that develop in the highly turbulent \lamhda flow studied in the preceding section is also left for future work; of particular interest is the occurrence of Kelvin-Helmholtz like roll-up of current sheets as observed at high resolution \cite{MiPo2007a}; however, the choice of the parameter $\alpha$ in the present paper was made on the basis of questioning the existence or lack thereof of a rigid-body high-wavenumber $k^{+1}$ spectrum and, thus, was not optimized for the study of the inertial range properties of the flow for which a much smaller value of the length $\alpha$ could be used.
1007:
1008:
1009: Finally, how far resolution can be reduced when using \lamhda as a LES for
1010: various statistics of interest will also require further detailed study. The present study shows that, to reproduce the super-filter-scale energy spectrum in three dimensions, gains by a factor of 1300 in computing time can be achieved. The need to reproduce higher order statistics can decrease these gains. As an example, in two-dimensional MHD, it was shown that gains when using \lamhda as a subgrid model depend for high order moments on the order that
1011: one wants to see
1012: to be accurately reproduced \cite{PGHM+06}.
1013:
1014: %The lack of strong contamination of super-filter-scales in \lamhda when compared with LANS can be associated with the interaction between the magnetic and the velocity field, which breaks the formation of rigid bodies in the flow. Associated with this process, the Lorentz force in \lamhda breaks the conservation of the circulation which is preserved in LANS. We can expect similar behavior of the Lagrangian-averaged models for other flows with mechanisms that can change the circulation, such as rotating or baroclinic flows.
1015:
1016:
1017:
1018: \begin{acknowledgments}
1019: Computer time was provided by GWDG, NCAR, and the National Science
1020: Foundation Terascale Computing System at the Pittsburgh Supercomputing
1021: Center. The NSF Grant No. CMG-0327888 at NCAR supported this work in
1022: part and is gratefully acknowledged. \add{}{PDM is a member of the
1023: Carrera del Investigador Cient\'{\i}fico of CONICET.} \resp{}{The
1024: anonymous referees are gratefully acknowledged for improving the
1025: clarity of the discussion of our results.}
1026: \end{acknowledgments}
1027:
1028:
1029: %\bibliographystyle{plain}
1030: %\bibliography{phd}
1031: \begin{thebibliography}{10}
1032:
1033: \bibitem{MK00}
1034: C.~{Meneveau} and J.~{Katz}.
1035: \newblock {Scale-Invariance and Turbulence Models for Large-Eddy Simulation}.
1036: \newblock {\em Annual Review of Fluid Mechanics}, 32:1--32, 2000.
1037:
1038: \bibitem{PFL76}
1039: A.~{Pouquet}, U.~{Frisch}, and J.~{Leorat}.
1040: \newblock {Strong MHD helical turbulence and the nonlinear dynamo effect}.
1041: \newblock {\em Journal of Fluid Mechanics}, 77:321--354, 1976.
1042:
1043: \bibitem{Y87}
1044: A.~{Yoshizawa}.
1045: \newblock {Subgrid modeling for magnetohydrodynamic turbulent shear flows}.
1046: \newblock {\em Physics of Fluids}, 30:1089--1095, 1987.
1047:
1048: \bibitem{ChLe1981}
1049: J.-P. {Chollet} and M.~{Lesieur}.
1050: \newblock {Parameterization of Small Scales of Three-Dimensional Isotropic
1051: Turbulence Utilizing Spectral Closures.}
1052: \newblock {\em Journal of Atmospheric Sciences}, 38:2747--2757, 1981.
1053:
1054: \bibitem{BaPoPo+2008}
1055: J.~{Baerenzung}, H.~{Politano}, Y.~{Ponty}, and A.~{Pouquet}.
1056: \newblock {Spectral modeling of magnetohydrodynamic turbulent flows}.
1057: \newblock {\em \pre}, 78(2):026310--+, August 2008.
1058:
1059: \bibitem{AMP05a}
1060: A.~{Alexakis}, P.~D. {Mininni}, and A.~{Pouquet}.
1061: \newblock {Shell-to-shell energy transfer in magnetohydrodynamics. I. Steady
1062: state turbulence}.
1063: \newblock {\em \pre}, 72(4):046301--+, 2005.
1064:
1065: \bibitem{MAP05}
1066: P.~{Mininni}, A.~{Alexakis}, and A.~{Pouquet}.
1067: \newblock {Shell-to-shell energy transfer in magnetohydrodynamics. II.
1068: Kinematic dynamo}.
1069: \newblock {\em \pre}, 72(4):046302--+, 2005.
1070:
1071: \bibitem{ZSG02}
1072: Y.~{Zhou}, O.~{Schilling}, and S.~{Ghosh}.
1073: \newblock {Subgrid scale and backscatter model for magnetohydrodynamic
1074: turbulence based on closure theory: Theoretical formulation}.
1075: \newblock {\em \pre}, 66(2):026309--+, 2002.
1076:
1077: \bibitem{GaNaNe+2000}
1078: S.~{Galtier}, S.~V. {Nazarenko}, A.~C. {Newell}, and A.~{Pouquet}.
1079: \newblock {A weak turbulence theory for incompressible magnetohydrodynamics}.
1080: \newblock {\em Journal of Plasma Physics}, 63:447--488, 2000.
1081:
1082: \bibitem{GoSr1995}
1083: P.~{Goldreich} and S.~{Sridhar}.
1084: \newblock {Toward a theory of interstellar turbulence. 2: Strong alfvenic
1085: turbulence}.
1086: \newblock {\em \apj}, 438:763--775, 1995.
1087:
1088: \bibitem{I64}
1089: P.~S. {Iroshnikov}.
1090: \newblock {Turbulence of a Conducting Fluid in a Strong Magnetic Field}.
1091: \newblock {\em Soviet Astronomy}, 7:566--+, 1964.
1092:
1093: \bibitem{K65}
1094: R.~H. {Kraichnan}.
1095: \newblock {Inertial-range spectrum of hydromagnetic turbulence}.
1096: \newblock {\em Physics of Fluids}, 8:1385--1387, 1965.
1097:
1098: \bibitem{MaCaBo2008}
1099: J.~{Mason}, F.~{Cattaneo}, and S.~{Boldyrev}.
1100: \newblock {Numerical measurements of the spectrum in magnetohydrodynamic
1101: turbulence}.
1102: \newblock {\em \pre}, 77(3):036403--+, 2008.
1103:
1104: \bibitem{MiPo2007a}
1105: P.~D. {Mininni} and A.~{Pouquet}.
1106: \newblock {Energy Spectra Stemming from Interactions of Alfv{\'e}n Waves and
1107: Turbulent Eddies}.
1108: \newblock {\em Physical Review Letters}, 99(25):254502--+, 2007.
1109:
1110: \bibitem{PoMiMo+2008}
1111: Annick {Pouquet}, Pablo {Mininni}, David {Montgomery}, and Alexandros
1112: {Alexakis}.
1113: \newblock {Dynamics of the Small Scales in Magnetohydrodynamic Turbulence }.
1114: \newblock In Yukio {Kaneda}, editor, {\em Proceedings of the IUTAM Symposium on
1115: Computational Physics and New Perspectives in Turbulence, Nagoya University,
1116: Nagoya, Japan, September, 11-14, 2006}, volume~4, pages 305--312.
1117: Springer-Verlag, 2008.
1118:
1119: \bibitem{AMK+01}
1120: O.~{Agullo}, W.-C. {M{\"u}ller}, B.~{Knaepen}, and D.~{Carati}.
1121: \newblock {Large eddy simulation of decaying magnetohydrodynamic turbulence
1122: with dynamic subgrid-modeling}.
1123: \newblock {\em Physics of Plasmas}, 8:3502--3505, 2001.
1124:
1125: \bibitem{TFS94}
1126: M.~L. {Theobald}, P.~A. {Fox}, and S.~{Sofia}.
1127: \newblock {A subgrid-scale resistivity for magnetohydrodynamics}.
1128: \newblock {\em Physics of Plasmas}, 1:3016--3032, 1994.
1129:
1130: \bibitem{HB06}
1131: N.~E.~L. {Haugen} and A.~{Brandenburg}.
1132: \newblock {Hydrodynamic and hydromagnetic energy spectra from large eddy
1133: simulations}.
1134: \newblock {\em Physics of Fluids}, 18:5106--+, 2006.
1135:
1136: \bibitem{MC02}
1137: W.-C. {M{\"u}ller} and D.~{Carati}.
1138: \newblock {Dynamic gradient-diffusion subgrid models for incompressible
1139: magnetohydrodynamic turbulence}.
1140: \newblock {\em Physics of Plasmas}, 9:824--834, 2002.
1141:
1142: \bibitem{LS91}
1143: D.~W. {Longcope} and R.~N. {Sudan}.
1144: \newblock {Renormalization group analysis of reduced magnetohydrodynamics with
1145: application to subgrid modeling}.
1146: \newblock {\em Physics of Fluids B}, 3:1945--1962, 1991.
1147:
1148: \bibitem{KM04}
1149: B.~{Knaepen} and P.~{Moin}.
1150: \newblock {Large-eddy simulation of conductive flows at low magnetic Reynolds
1151: number}.
1152: \newblock {\em Physics of Fluids}, 16:1255--+, 2004.
1153:
1154: \bibitem{PMM+05}
1155: Y.~{Ponty}, P.~D. {Mininni}, D.~C. {Montgomery}, J.-F. {Pinton}, H.~{Politano},
1156: and A.~{Pouquet}.
1157: \newblock {Numerical Study of Dynamo Action at Low Magnetic Prandtl Numbers}.
1158: \newblock {\em Physical Review Letters}, 94(16):164502--+, 2005.
1159:
1160: \bibitem{PPP04}
1161: Y.~{Ponty}, H.~{Politano}, and J.-F. {Pinton}.
1162: \newblock {Simulation of Induction at Low Magnetic Prandtl Number}.
1163: \newblock {\em Physical Review Letters}, 92(14):144503--+, 2004.
1164:
1165: \bibitem{H02b}
1166: D.~D. {Holm}.
1167: \newblock {Averaged Lagrangians and the mean effects of fluctuations in ideal
1168: fluid dynamics}.
1169: \newblock {\em Physica D Nonlinear Phenomena}, 170:253--286, 2002.
1170:
1171: \bibitem{H02a}
1172: D.~D. {Holm}.
1173: \newblock {Lagrangian averages, averaged Lagrangians, and the mean effects of
1174: fluctuations in fluid dynamics}.
1175: \newblock {\em Chaos}, 12:518--530, 2002.
1176:
1177: \bibitem{MP02}
1178: D.~C. {Montgomery} and A.~{Pouquet}.
1179: \newblock {An alternative interpretation for the Holm ``alpha model''}.
1180: \newblock {\em Physics of Fluids}, 14(9):3365--3366, 2002.
1181:
1182: \bibitem{MMP05a}
1183: P.~D. {Mininni}, D.~C. {Montgomery}, and A.~G. {Pouquet}.
1184: \newblock {A numerical study of the alpha model for two-dimensional
1185: magnetohydrodynamic turbulent flows}.
1186: \newblock {\em Physics of Fluids}, 17:5112--+, 2005.
1187:
1188: \bibitem{PGMP05}
1189: J.~Pietarila Graham, P.~D. {Mininni}, and A.~{Pouquet}.
1190: \newblock {Cancellation exponent and multifractal structure in two-dimensional
1191: magnetohydrodynamics: Direct numerical simulations and Lagrangian averaged
1192: modeling}.
1193: \newblock {\em \pre}, 72(4):045301(R)--+, 2005.
1194:
1195: \bibitem{PGHM+06}
1196: J.~Pietarila Graham, D.~D. {Holm}, P.~{Mininni}, and A.~{Pouquet}.
1197: \newblock {Inertial range scaling, K\'arm\'an-Howarth theorem, and
1198: intermittency for forced and decaying Lagrangian averaged magnetohydrodynamic
1199: equations in two dimensions}.
1200: \newblock {\em Physics of Fluids}, 18:045106, 2006.
1201:
1202: \bibitem{MMP05b}
1203: P.~D. {Mininni}, D.~C. {Montgomery}, and A.~{Pouquet}.
1204: \newblock {Numerical solutions of the three-dimensional magnetohydrodynamic
1205: {$\alpha$} model}.
1206: \newblock {\em \pre}, 71(4):046304--+, 2005.
1207:
1208: \bibitem{Mi2006a}
1209: P.~D. {Mininni}.
1210: \newblock {Turbulent magnetic dynamo excitation at low magnetic Prandtl
1211: number}.
1212: \newblock {\em Physics of Plasmas}, 13:056502--+, 2006.
1213:
1214: \bibitem{CHO+05}
1215: Alexey Cheskidov, Darryl~D. Holm, Eric Olson, and Edriss~S. Titi.
1216: \newblock {On a Leray$-\alpha$ model of turbulence}.
1217: \newblock {\em Proceedings of the Royal Society of London}, A461:629--649,
1218: 2005.
1219:
1220: \bibitem{PGHM+07a}
1221: J.~Pietarila Graham, D.~{Holm}, P.~{Mininni}, and A.~{Pouquet}.
1222: \newblock {Highly turbulent solutions of the Lagrangian-averaged Navier-Stokes
1223: alpha model and their large-eddy-simulation potential}.
1224: \newblock {\em \pre}, 76:056310--+, 2007.
1225:
1226: \bibitem{MiPoSu2008}
1227: P.D. {Mininni}, A.~{Pouquet}, and P.~{Sullivan}.
1228: \newblock {Two examples from geophysical and astrophysical turbulence on
1229: modeling disparate scale interactions}.
1230: \newblock In Roger {Temam} and Joe {Tribbia}, editors, {\em Summer school on
1231: mathematics in geophysics}. Springer-Verlag, 2008.
1232: \newblock to appear.
1233:
1234: \bibitem{GMD05b}
1235: D.~O. {G{\'o}mez}, P.~D. {Mininni}, and P.~{Dmitruk}.
1236: \newblock {MHD simulations and astrophysical applications}.
1237: \newblock {\em Advances in Space Research}, 35:899--907, 2005.
1238:
1239: \bibitem{GMD05}
1240: D.~O. {G{\'o}mez}, P.~D. {Mininni}, and P.~{Dmitruk}.
1241: \newblock {Parallel Simulations in Turbulent MHD}.
1242: \newblock {\em Physica Scripta Volume T}, 116:123--127, 2005.
1243:
1244: \bibitem{MaPoMi+2008}
1245: W.~H. {Matthaeus}, A.~{Pouquet}, P.~D. {Mininni}, P.~{Dmitruk}, and
1246: B.~{Breech}.
1247: \newblock {Rapid Alignment of Velocity and Magnetic Field in
1248: Magnetohydrodynamic Turbulence}.
1249: \newblock {\em Physical Review Letters}, 100(8):085003--+, 2008.
1250:
1251: \bibitem{MiPoMo2006}
1252: P.~D. {Mininni}, A.~G. {Pouquet}, and D.~C. {Montgomery}.
1253: \newblock {Small-Scale Structures in Three-Dimensional Magnetohydrodynamic
1254: Turbulence}.
1255: \newblock {\em Physical Review Letters}, 97(24):244503--+, 2006.
1256:
1257: \bibitem{FHT01}
1258: C.~{Foias}, D.~D. {Holm}, and E.~S. {Titi}.
1259: \newblock {The Navier-Stokes-alpha model of fluid turbulence}.
1260: \newblock {\em Physica D Nonlinear Phenomena}, 152-153:505--519, May 2001.
1261:
1262: \bibitem{LKT+07}
1263: E.~{Lunasin}, S.~{Kurien}, M.~{Taylor}, and E.~{Titi}.
1264: \newblock {A study of the Navier-Stokes-alpha model for two-dimensional
1265: turbulence}.
1266: \newblock {\em ArXiv Physics e-prints}, 2007.
1267:
1268: \bibitem{PP98b}
1269: H\'el\`ene {Politano} and Annick {Pouquet}.
1270: \newblock {Dynamical length scales for turbulent magnetized flows}.
1271: \newblock {\em Geophysical Research Letters}, 25(3):273--276, 1998.
1272:
1273: \bibitem{K67}
1274: R.~H. {Kraichnan}.
1275: \newblock {Inertial Ranges in Two-Dimensional Turbulence}.
1276: \newblock {\em Physics of Fluids}, 10:1417--1423, 1967.
1277:
1278: \bibitem{CHT05}
1279: C.~{Cao}, D.~D. {Holm}, and E.~S. {Titi}.
1280: \newblock {On the Clark {$\alpha$} model of turbulence: global regularity and
1281: long-time dynamics}.
1282: \newblock {\em Journal of Turbulence}, 6:19--+, 2005.
1283:
1284: \bibitem{DmGoMa2003}
1285: P.~{Dmitruk}, D.~O. {G{\'o}mez}, and W.~H. {Matthaeus}.
1286: \newblock {Energy spectrum of turbulent fluctuations in boundary driven reduced
1287: magnetohydrodynamics}.
1288: \newblock {\em Physics of Plasmas}, 10:3584--3591, 2003.
1289:
1290: \bibitem{Al2007a}
1291: A.~{Alexakis}.
1292: \newblock {Nonlocal Phenomenology for Anisotropic Magnetohydrodynamic
1293: Turbulence}.
1294: \newblock {\em \apjl}, 667:L93--L96, 2007.
1295:
1296: \bibitem{HSL+82}
1297: J.~R. {Herring}, D.~{Schertzer}, M.~{Lesieur}, G.~R. {Newman}, J.~P. {Chollet},
1298: and M.~{Larcheveque}.
1299: \newblock {A comparative assessment of spectral closures as applied to passive
1300: scalar diffusion}.
1301: \newblock {\em Journal of Fluid Mechanics}, 124:411--437, 1982.
1302:
1303: \bibitem{LMG95}
1304: D.~{Lohse} and A.~{M{\"u}ller-Groeling}.
1305: \newblock {Bottleneck Effects in Turbulence: Scaling Phenomena in r versus p
1306: Space}.
1307: \newblock {\em Physical Review Letters}, 74:1747--1750, 1995.
1308:
1309: \bibitem{MCD+97}
1310: D.~O. {Mart{\'{\i}}nez}, S.~{Chen}, G.~D. {Doolen}, R.~H. {Kraichnan}, L.-P.
1311: {Wang}, and Y.~{Zhou}.
1312: \newblock {Energy spectrum in the dissipation range of fluid turbulence}.
1313: \newblock {\em Journal of Plasma Physics}, 57:195--201, 1997.
1314:
1315: \bibitem{MAP06}
1316: P.~D. {Mininni}, A.~{Alexakis}, and A.~{Pouquet}.
1317: \newblock {Large-scale flow effects, energy transfer, and self-similarity on
1318: turbulence}.
1319: \newblock {\em \pre}, 74(1):016303--+, 2006.
1320:
1321: \bibitem{MT92}
1322: H.~K. {Moffatt} and A.~{Tsinober}.
1323: \newblock {Helicity in laminar and turbulent flow}.
1324: \newblock {\em Annual Review of Fluid Mechanics}, 24:281--312, 1992.
1325:
1326: \bibitem{T01}
1327: Arkady {Tsinober}.
1328: \newblock {\em {An Informal Introduction to Turbulence}}.
1329: \newblock Kluwer Academic Publishers, Dordrecht, 2001.
1330:
1331: \bibitem{CHM+99}
1332: S.~{Chen}, D.~D. {Holm}, L.~G. {Margolin}, and R.~{Zhang}.
1333: \newblock {Direct numerical simulations of the Navier-Stokes alpha model}.
1334: \newblock {\em Physica D Nonlinear Phenomena}, 133:66--83, 1999.
1335:
1336: \bibitem{PiGrHoMi+2008}
1337: J.~Pietarila Graham, D.~D. {Holm}, P.~D. {Mininni}, and A.~{Pouquet}.
1338: \newblock {Three regularization models of the Navier-Stokes equations}.
1339: \newblock {\em Physics of Fluids}, 20(3):035107--+, 2008.
1340:
1341: \bibitem{FrPoLe+1975}
1342: U.~{Frisch}, A.~{Pouquet}, J.~{L\'eorat}, and A.~{Mazure}.
1343: \newblock {Possibility of an inverse cascade of magnetic helicity in
1344: magnetohydrodynamic turbulence}.
1345: \newblock {\em Journal of Fluid Mechanics}, 68:769--778, 1975.
1346:
1347: \bibitem{HeHoPe+2008}
1348: M.~W. {Hecht}, D.~D. {Holm}, M.~R. {Petersen}, and B.~A. {Wingate}.
1349: \newblock {Implementation of the LANS-{$\alpha$} turbulence model in a
1350: primitive equation ocean model}.
1351: \newblock {\em Journal of Computational Physics}, 227:5691--5716, 2008.
1352:
1353: \bibitem{HaBrDo2003}
1354: N.~E.~L. {Haugen}, A.~{Brandenburg}, and W.~{Dobler}.
1355: \newblock {Is Nonhelical Hydromagnetic Turbulence Peaked at Small Scales?}
1356: \newblock {\em \apjl}, 597:L141--L144, 2003.
1357:
1358: \bibitem{Mi2007a}
1359: P.~D. {Mininni}.
1360: \newblock {Inverse cascades and {$\alpha$} effect at a low magnetic Prandtl
1361: number}.
1362: \newblock {\em \pre}, 76(2):026316--+, 2007.
1363:
1364: \bibitem{AM78}
1365: D.~G. {Andrews} and M.~E. {McIntyre}.
1366: \newblock {An exact theory of nonlinear waves on a Lagrangian-mean flow}.
1367: \newblock {\em J. Fluid Mech.}, 89:609--646, 1978.
1368:
1369: \end{thebibliography}
1370:
1371: \end{document}
1372:
1373: % LocalWords: Leray LANS DNS subgrid pseudospectral streamwise eps pdfs Yukawa
1374: % LocalWords: discretization Howarth Kraichnan advected Subdominance Eq IK
1375: % LocalWords: ABC Eqs Lyapunov MHD SGS NCAR Terascale Supercomputing CMG LAMHD
1376: % LocalWords: Sonnensystemforschung Katlenburg Lindau Departamento Facultad
1377: % LocalWords: Ciencias Exactas Naturales Universidad Buenos Ciudad diffusivity
1378: % LocalWords: Universitaria hyperdiffusivity hyperviscosity hyperdiffusion emf
1379: % LocalWords: Beltrami enstrophy BaPoPo Prandtl electromotive barotropic PDFs
1380: % LocalWords: Iroshnikov hydrodynamical baroclinic GLM Rossby
1381: