You are on page 1of 361

Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.

866227
Advances in
Turbulence
Studies
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Edited by
Herman Branover
Yeshajahu linger
Ben-Gurion University of the Negev
Beer-Sheva, Israel

Volume 149
PROGRESS IN
ASTRONAUTICS AND AERONAUTICS

A. Richard Seebass, Editor-in-Chief


University of Colorado at Boulder
Boulder, Colorado

Technical papers from the Proceedings of the Sixth Beer-Sheva Interna-


tional Seminar on Magnetohydrodynamic Flows and Turbulence, Ben-
Gurion University of the Negev, Beer-Sheva, Israel, February 25-March 2,
1990, and subsequently revised for this volume.
Published by the American Institute of Aeronautics and Astronautics, Inc..
370 L'Enfant Promenade, SW, Washington, DC, 20024-2518
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Copyright © 1993 by the American Institute of Aeronautics and Astronautics, Inc. Printed in
the United States of America. All rights reserved. Reproduction or translation of any part of
this work beyond that permitted by Sections 107 and 108 of the U.S. Copyright Law without
the permission of the copyright owner is unlawful. The code following this statement indicates
the copyright owner's consent that copies of articles in this volume may be made for personal
or internal use, on condition that the copier pay the per-copy fee ($2.00) plus the per-page fee
($0.50) through the Copyright Clearance Center, Inc., 21 Congress Street, Salem, Massachu-
setts 01970. This consent does not extend to other kinds of copying, for which permission
requests should be addressed to the publisher. Users should employ the following code when
reporting copying from this volume to the Copyright Clearance Center:

1-56347-018-7/93 $2.00 + .50

Data and information appearing in this book are for informational purposes only. AIAA is not
responsible for any injury or damage resulting from use or reliance, nor does AIAA warrant
that use or reliance will be free from privately owned rights.

ISSN 0079-6050
Progress in Astronautics and Aeronautics
Editor-in-Chief
A. Richard Seebass
University of Colorado at Boulder
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Editorial Board
Richard G. Bradley John L. Junkins
General Dynamics Texas A&M University

Alien E. Fuhs John E. Keigler


Carmel, California General Electric Company
Astro-Space Division
George J. Gleghorn
TR W Space Daniel P. Raymer
and Technology Group Conceptual Research
Corporation
Dale B. Henderson
Los Alamos National Laboratory Martin Summerfield
Princeton Combustion Research
Carolyn L. Huntoon Laboratories, Inc.
NASA Johnson Space Center
Charles E. Treanor
Reid R. June Arvin / Calspan
Boeing Military Airplane Company Advanced Technology Center

Jeanne Godette
Director
Book Publications
AIAA
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

This page intentionally left blank


Preface

This volume contains a collection of papers devoted to modern trends in the


research of turbulence, which were presented at the Sixth Beer-Sheva Inter-
national Seminar on Magnetohydrodynamic (MHD) Flows and Turbulence.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

The Beer-Sheva Seminars have been held every three years since 1975,
when a group of researchers from different countries, interested in both
MHD flows and different aspects of turbulence in electroconductive as well
as nonconductive fluids, came together for a week-long seminar. All partic-
ipants felt that the meeting was very fruitful in detecting promising areas of
research and developing new ideas. Therefore, the seminar was repeated in
1978 and ultimately became an ongoing event, taking place once every three
years. It also became a tradition to publish the papers presented at the
Beer-Sheva Seminar after a thorough review and editing. The first two
volumes were published by John Wiley and Sons, and the subsequent
volumes were published by AIAA (Volumes 84, 100, 111, and 112 in the
Progress in Astronautics and Aeronautics series).
The Beer-Sheva Seminars became well known in the international scien-
tific community, and the number of researchers willing to participate kept
growing steadily. This forced the organizers, who wanted to preserve the
intimate atmosphere of the seminars and to keep substantial time for
informal discussions, to introduce more and more severe criteria in the
paper selection process. The intention was to keep the number of partici-
pants in each event at approximately 100. The Sixth Seminar, held from
February 25 to March 2, 1990, hosted 128 participants from 18 countries.
For the first time, East European countries participated, including massive
delegations from the CIS. The number of papers that passed all of the
reviews reached 72, which is many more than at previous seminars. In
addition, a number of extended invited review papers were presented.
Therefore, the papers on MHD and on turbulence are published in two
separate volumes.
The papers of the present volume, Advances in Turbulence Studies, cover
many important contemporary trends in both experimental and theoretical
turbulence research. They give a concise and comprehensive picture of the
present status in most of the areas in which extensive studies are being
carried out. The authors of many of the articles in this volume are recog-
nized leaders in turbulence research.
To better acquaint the reader with the content of this volume, we will
briefly describe several experimental as well as several theoretical papers.
Hussain et al. address the question of reconnection of vortex tubes in
viscous flows as an example of topological transformation in fluid mechan-
ics. Numerical simulations of the crosslinking of two antiparallel vortex
tubes are presented to illustrate the various steps of the process (inviscid

The names of authors who actually presented their papers are printed in italics.
induction, stretching, and bridging), which is believed to play an important
role in mixing and turbulence production.
Kraichnan discusses the intermittency of small-scale turbulence. He pre-
sents a new closure approximation centered about the systematic approxi-
mation of joint probability distribution for fields and their gradients at
single points in space. These distributions are sampled along a fluid-element
path, and closure is obtained by nonlinear mapping of Gaussian fields into
non-Gaussian fields according to a particular procedure. The resulting
closure approximations are wholly nonperturbational. They yield intermit-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

tency phenomena as the combined effect of straining and molecular dissipa-


tion. An application to Burger's equation predicts a nearly exponential tail
to the velocity gradient distribution that is verified by simulations.
Fiedler et al. describe measurements taken in the plane mixing layer
formed between two parallel streams with different velocities and densities.
A comprehensive set of velocity-velocity, velocity-density, and density-den-
sity statistical correlations is shown offering an interesting test case for the
development of stress-mass flux turbulence models. Comparison with the
results of a numerical direct simulation is also discussed. Special attention
is focused on the differences between the two configurations of cogradient
and countergradient flows. The growth rate of the second is found to be
larger than that of the first, with pronounced effects at the lower density
side of the flow.
Paschereit and Wygnanski study the resonant subharmonic interaction
between two axisymmetric traveling waves in the shear layer of an axisym-
metric jet. The first part of the experiment entails forcing the jet at a single
frequency. Downstream, the phase speeds of the fundamental frequency
and its first subharmonic are sufficiently similar that a narrow band of
frequencies centered on the subharmonic results from the excitation. Here,
the amplification rate of the subharmonic wave is enhanced because the
excited waves are nondispersive, and the phase difference is constant.
However, no response is seen at the second subharmonic when the jet is
forced at the first subharmonic, because these waves do not propagate at
the same wave speed. The second part of the experiment studies the effect
of a phase difference between the fundamental wave and the subharmonic
on the growth of the subharmonic. In comparison to phase differences of 0,
180, and 270 deg, the maximum amplitude of the subharmonic wave is
suppressed for an initial phase difference of 90 deg. It is noted that the
relative amplitude of the subharmonic does not affect the maximum ampli-
tude observed.
Zimin and Kolpakov develop a closed system of equations for three-di-
mensional vortex MHD flows in thin layers at high Reynolds numbers
through local-equilibrium boundary-layer equations. Because fluid is ex-
changed between the boundary layer and the core flow, classical boundary-
layer theory is inappropriate. In essence, the three-dimensional equations
are reduced to two-dimensional equations, and then the full solution is
reconstructed from the two-dimensional results. The solution technique is
demonstrated on a problem involving an aluminum electrolyzer. The results
of the local-equilibrium boundary-layer method are in reasonable agree-
ment with a numerical solution, with the exception of the axial velocity
component. This discrepancy can be rectified by using higher-order approx-
imations to the boundary-layer velocity components.
Politano, Pouquet, and Sulem present direct numerical simulations of
two-dimensional MHD incompressible flows at resolutions up to 10242 grid
points. An inertial range is observed, with spectral properties depending on
the velocity-magnetic field correlation. At very small scales, resistive tearing
destabilizes current sheets resulting from the inertial dynamics and leads to
the formation of small-scale magnetic islands.
The application of the renormalization group method is presented in the
papers of Sukoriansky et al., Staroseisky and Sukoriansky, and Talmage et
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

al. The latter describes the hydrodynamic behavior of the sliding liquid-
metal electric contact. Turbulence is simulated by the scalar eddy viscosity
obtained by the RNG method. The effects of turbulence in the presence of
a magnetic field are shown, and the importance of the adjustment of the
turbulence modeling to the suppression of turbulence by the electromag-
netic interactions is emphasized. The flow formed between the fast moving
rotor and the stationary stator in the presence of both an open surface and
an axial magnetic field is also numerically calculated for the laminar flow
regimes.
Dynamo effects are discussed in several papers, including those coau-
thored by Shtern (separately with Goldshtik, with Sergeev, and with
Petrunin). Sulem, Galanti, and Gilbert show that an inverse cascade of
magnetic field is not confined to turbulent MHD. Using the information
that the fluid is driven by a given steady body force, they systematically
derived equations for the evolution and saturation of dynamo instabilities.
The growth of an initially weak magnetic field is first dominated by the
linearly most unstable mode, and later the magnetic excitation is transferred
to larger and larger scales. They also consider a time-dependent anisotropic
forcing for which, in addition to the dynamo action, a linear instability of
large-scale velocity, called AKA instability, is present.
Another paper that discusses geophysical phenomena is that ofHenoch et
al. The analogy between large-scale strongly anisotropic (almost two-di-
mensional) turbulence caused by a magnetic field in flows of electroconduc-
tive fluids and by density stratification in geophysical flows, mainly ocean
flows, is analyzed. The analogy appears to be very profound. In both cases
there is an inverse energy transfer towards the large-scale turbulent move-
ments. In certain cases energy is transferred from the disturbances to the
mean flow, which can be interpreted as manifestation of negative eddy
viscosity. The analogy is also followed through by comparing energy spec-
tra. The above analogy can be used for experimental simulation of large-
scale ocean flow turbulence phenomena by small-scale liquid-metal MHD
experiments.
Vlasyuk and Shcherbinin discuss the numerical calculation of the axisym-
metric vortical flow in molten metal induced by an axial electric current.
Whereas the whole bottom plate of the vertical container serves as an
electrode, the second electrode occupies but a small part of the top of the
container. An azimuthal (axisymmetric) magnetic field is self-induced and
interacts with the turbulent flow. A comparison is made between two flow
regimes (laminar and turbulent) and between two mathematical models of
turbulence (the k-e model and the k-W model). The possibility that the
circulation level is augmented by the magnetohydrodynamic turbulence is
suggested.
Naot et al. manipulate the standard k-e turbulence model equations and
obtain a transport equation for the high Reynolds number dissipation
length. The geometric nature of this equation is discussed for open channel
flows, closed duct flows, and the horizontal slug flow. The physical situa-
tions in which the equation for the dissipation length becomes geometric,
depending on the field geometry and boundary conditions only, are dis-
cussed, casting a new interpretation in terms of length scale dynamics on the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

choice of k-e model coefficients.


Ruzmaikin develops a framework for describing fluctuating magnetic
fields in a turbulent flow. He addresses such issues as the temporal growth
threshold in the subcritical region where the Reynolds number is less than
some critical value, the role of boundaries, and the creation of intermittency
by a fast dynamo effect. Since the magnetic field structure in a random flow
is intermittent in the large Reynolds number limit, the behavior of the
magnetic field cannot be characterized solely by the mean field and the
second moment. A ratio of the moments is required, as is the distribution
function of the random field, where possible.
The two following papers deal with instabilities and are not in the main-
stream of this volume.
Mond and Hameiri present a "ballooning" mode or pressure-driven
instability theory for both classical and plasma flows. In classical flows,
ballooning modes entail waves that propagate along stream lines, whereas
in plasma flows, ballooning modes entail Alfen and slow magnetosonic
waves that propagate along magnetic field lines. The governing problem is
reduced to a one-dimensional ordinary differential equation for the propa-
gation of waves along either stream lines or magnetic field lines. Results
illustrate the importance of the dynamic pressure and the curvature in
determining the stability of the flow.
Rutkevich applies a modified method of geometrical acoustics to nonuni-
form flows of low-temperature plasmas through MHD channels to analyze
the stability of these flows with respect to high-frequency perturbations.
The instability modes in supersonic channels have been previously de-
scribed. Here, the instability mode associated with the cutoff phenomena of
acoustic wave propagation in nonuniform subsonic MHD flows is investi-
gated.
Without going into detail, we should mention here that a number of
papers deal with the results of direct numerical simulation of both hydrody-
namic and magnetohydrodynamic turbulence.
In their totality, the papers of this volume referred to above as well as
those not mentioned here present a quite complete picture of the contempo-
rary trends in turbulence research. They present a substantial amount of the
most recent results. The extensive references given in most papers of this
volume should be an additional aid to the reader who wishes to make
himself familiar with the present status of the field.
The authors and editors are very pleased that AIAA is continuing to
publish the proceedings of the Beer-Sheva International Seminars on MHD
Flows and Turbulence in their Progress in Astronautics and Aeronautics
series.
The editors would like to express their sincerest gratitude to Dr. Martin
Summerfield for his continuing enthusiastic interest in the Beer-Sheva Sem-
inars and for his permanent guidance, and to Ms. Jeanne Godette, Director
of the Progress in Astronautics and Aeronautics series, and her staff mem-
bers for their continuing creative assistance, as well as to the entire team of
the Progress in Astronautics and Aeronautics series.
We now look forward to the next volumes of this series, and we invite the
reader to participate in the 7th Beer-Sheva International Seminar on MHD
Flows and Turbulence, which is planned for February 1993.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Herman Branover
Yeshajahu Unger
November 1992
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

This page intentionally left blank


Table of Contents

Preface
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Vortex Reconnection, Cascade, and Mixing in Turbulent F l o w s . . . . . . . . . . 1


F. Hussain, D. Virk, and M. V. Melander, University of Houston, Houston, Texas

Intermittent TYirbulence from C l o s u r e s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17


Robert H. Kraichnan, 369 Montezuma 108, Santa Fe, New Mexico

Plane Mixing Layer Between Parallel Streams of Different Velocities


and Different Densities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
H. E. Fiedler, M. Lummer, and K. Nottmeyer, Technische Universitdt Berlin,
Berlin, Germany

Instabilities in the Axisymmetric Jet: Subharmonic Resonance . . . . . . . . . 53


C. O. Paschereit and I. J. Wygnanski, University of Arizona, Tucson, Arizona

Three-Dimensional Vortex MHD Flows at High Reynolds Numbers


in Thin Layers of Conducting Incompressible Fluid . . . . . . . . . . . . . . . . 65
V. D. Zimin and N. Ju. Kolpakov, Institute of Continuous Media Mechanics,
Perm, Russia

Tearing Instabilities in Two-Dimensional MHD Turbulence . . . . . . . . . . . . 81


H. Politano, A. Pouquet, and P. L. Sulem, Observatoire de la Cote d'Azur, Nice,
France

Axisymmetric Hydromagnetic D y n a m o . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
M. A. Goldshtik and V. N. Shtern, Siberian Branch of the Academy of Sciences,
Novosibirsk, Siberia

Bifurcations of Self-Oscillating and Almost Periodical Regimes in an


Azimuthal MHD J e t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
K. Sergeev and V. N. Shtern, Siberian Branch of the Academy of Sciences,
Novosibirsk, Siberia

Bifurcations in MHD Flow Generated by Electric Current Discharge... 116


A. A. Petrunin and V. N. Shtern, Siberian Branch of the Academy of Sciences,
Novosibirsk, Siberia

Homogeneous MHD Turbulence at Weak Magnetic Reynolds Numbers:


Approach to Angular-Dependent Spectra . . . . . . . . . . . . . . . . . . . . . . . . 131
C. Cambon, Ecole Centrale de Lyon, Ecully, France
Inverse Cascades Generated by Alpha Dynamo and Anisotropic Kinetic
Alpha Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
P. L. Sulem and B. Galanti, Observatoire de la Cote d'Azur, Nice, France, and
A. D. Gilbert, Cambridge University, Cambridge, England

Renormalization Group Analysis of MHD Turbulence with Low


Magnetic Reynolds Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
S. Sukoriansky, I. Staroselsky, B. Galperin, S. Roy, and S. A. Orszag,
Princeton University, Princeton, New Jersey
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Renormalization Group Approach to Two-Dimensional Turbulence


and the e-Expansion for the Vorticity Equation . . . . . . . . . . . . . . . . . . . 159
I. Staroselsky and S. Sukoriansky, Princeton University, Princeton, New Jersey

Liquid-Metal Flows in Sliding Electric Contacts: Solution for Turbulent


Primary Azimuthal Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
G. Talmage and J. S. Walker, University of Illinois at Champain-Urbana, Urbana,
Illinois, S. H. Brown and N. A. Sondergaard, David Taylor Research Center,
Annapolis, Maryland, and H. Branover and S. Sukoriansky, Ben-Gurion
University of the Negev, Beer-Sheva, Israel

Anisotropic Turbulence: Analogies Between Geophysical and


Hydromagnetic Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
C. Henoch and M. Hoffert, New York University, New York, New York, and
H. Branover and S. Sukoriansky, Ben-Gurion University of the Negev,
Beer-Sheva, Israel

l\irbulent Electrically-Induced Vortical Flows . . . . . . . . . . . . . . . . . . . . . . 210


V. N. Vlasyuk, Vinica Institute of Teaching, Vinica, Ukraine, and E. V.
Shcherbinin, Latvian Academy of Sciences, Riga-Salaspils, Latvia

Dissipation Length Scale Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221


D. Naot, Center for Technological Education Holon, Holon, Israel, and
N. Yacoub and D. Maron Moalem, Tel Aviv University, RamatAviv, Israel

Towards Quasi-Isotropic Algebraic Stress Model for


Magnetohydrodynamic Channel F l o w . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
D. Naot and J. Tanny, Center for Technological Education Holon, Holon, Israel

Two-Phase Grid Turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254


Th. Panidis and D. D. Papailiou, University ofPatras, Patras-Rion, Greece

Abridged Octave Wavenumber Ring Models for Two-Dimensional


Turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
J. Lee, Flight Dynamics Laboratory, Wright-Patterson Air Force Base, Ohio

Solving Partial Differential Equations via Boolean Automata: Statistical


and Deterministic Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
K. Dang Tran, A. Cosnuau, J. Ryan, and Y. Morchoisne, Office National
d'Etudes et de Recherches Aerospatiales, Chdtillon, France
Rag Theory of Magnetic Fluctuations in Turbulent Flow . . . . . . . . . . . . . 309
A. A. Ruzmaikin, Institute of Terrestrial Magnetism, Ionosphere, and Radio Wave
Propagation, Troitsk, Russia

Ballooning Instability in Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 317


M. Mond, Ben-Gurion University of the Negev, Beer-Sheva, Israel,
and E. Hameiri, New York University, New York, New York

Instabilities of the Nonuniform Flows of a Low-Temperature Plasma in


MHD C h a n n e l s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

I. M. Rutkevich, Russian Academy of Sciences, Moscow, Russia

Author Index for Volume 149 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342

List of Series Volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343


Table of Contents for Companion Volume 148

Preface

Chapter 1. Metallurgical Technologies


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Metallurgical Applications of M a g n e t o h y d r o d y n a m i c s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Yves Fautrelle, Madylam-Institut National Polytechnique de Grenoble, St. Martin d'Heres, France

Research and Development in the Field of MHD Devices Utilizing Liquid Working Medium
for Process Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
V. M. levlev and N. N. Baranov, Russian Academy of Sciences, Moscow, Russia

Application of MHD Facilities to Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32


Yu. M. Gelfgat, Latvian Academy of Sciences, Riga-Salaspils, Latvia

Electromagnetic Modulation of Molten Metal Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50


R. M. Slepian, Westinghouse Science & Technology Center, Pittsburgh, Pennsylvania, P. A. Davidson,
Imperial College, London, England, and A. R. Keeton, Westinghouse Science & Technology Center,
Pittsburgh, Pennsylvania

Superconducting MHD Devices: Parametric Study of an Electromagnetic Pump Performance . . . . . . . . . . 66


P. Del Vecchio, A. Geri, and G. M. Veca, University "La Sapienza," Rome, Italy

Induction Electromagnetic Pumps for Alkali Metals: Status and P e r s p e c t i v e s . . . . . . . . . . . . . . . . . . . . . . . . . 76


S. Poinsot and R. Rossignol, Centre d'Etudes de Cadarache, Saint Paul Lez Durance, France, F. Werkoff
and A. Marechal, Centre d'Etudes de Grenoble, Grenoble, France, and J. Rapin, FRAMATOME-Direction
NOVATOME, Lyon, France

Physical Model for Electromagnetically Driven Flow in Channel Induction Furnaces . . . . . . . . . . . . . . . . . . 92


Rene Ricou and Charles Vives, Universite d'Avignon, Avignon, France

Grain Refinement in Aluminum Alloys by an Electromagnetic Vibrational Method . . . . . . . . . . . . . . . . . . . 107


Charles Vives, Universite d'Avignon, Avignon, France

Dendrite Growth of Solidifying Metal in DC Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125


Eiichi Takeuchi, Ikuto Miyoshino, Kouichi Takeda, and Yutaka Kishida, Nippon Steel Corporation,
Chiba, Japan

MHD Means for Affecting Hydrodynamics, Heat Transfer, and Mass Transfer at Single Crystal
Melt G r o w t h . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Yu. M. Gelfgat and L. A. Gorbunov, Latvian Academy of Sciences, Riga-Salaspils, Latvia

Inverse Electromagnetic Shaping Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158


T. P. Felici and J. P. Brancher, Institut National Polytechnique de Lorraine, Nancy, France

Chapter 2. Energy Conversion

Major Engineering Physics for Optimization of the Seawater Superconducting Electromagnetic


Thruster . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
J. C. S. Meng, Naval Underwater Systems Center, Newport, Rhode Island

Liquid-Metal MHD Research and Development in I s r a e l . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209


H. Branover, Ben-Gurion University of the Negev, Beer-Sheva, Israel
OMACON Technology for Seawater Desalination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
A. Barak and E. Greenspan, Israel Atomic Energy Commission, Tel-Aviv, Israel

MHD Generator for Waste Heat Utilization in Northern Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239


I. M. Kirko, Perm Institute of Technical Physics, Perm, Russia

Recent Results on LMMHD Induction Generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244


Ph. Marty, L. Leboucher, A. Alemany, and A. Pilaud, Institut de Mecanique de Grenoble, Grenoble, France,
Ph. Masse, Madylam-Institut National Polytechnique de Grenoble, St. Martin d'Heres, France, and
H. Branover, A. El Boher, and Y. Kaplan, Ben-Gurion University of the Negev, Beer-Sheva, Israel

Analytical and Experimental Studies of End Effects in an LMMHD Generator . . . . . . . . . . . . . . . . . . . . . . 261


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

T. K. Thiyagarajan, P. Satyamurthy, N. S. Dixit, and N. Venkatramani, Bhabha Atomic Research Center,


Bombay, India, and V. K. Rohatgi, University ofPoona, Pune, India

Theoretical Magnetic Field Distributions Eliminating End Losses in Linear High Magnetic
Reynolds Number MHD C h a n n e l s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
L. Blumenau, Ben-Gurion University of the Negev, Beer-Sheva, Israel

Embrittlement of Steels by Lead . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310


Rangarajan Venkataraman, Michael D. Baldwin, and Glen R. Edwards, Colorado School of Mines, Golden,
Colorado

Materials Compatibility of Mercury for Practical Applications at Elevated Temperatures . . . . . . . . . . . . . 335


Noel W. Kirshenbaum, Placer Dome U.S. Inc., San Francisco, California

Status of MHD Energy Conversion Research in Poland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343


B. Zaporowski, Technical University ofPosnan, Posnan, Poland

Open Cycle Disk Generator Operating Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348


H. K. Messerle, University of Sydney, Sydney, Australia

Conceptual Design of an MHD Retrofit of the Corette Plant in Billings, Montana . . . . . . . . . . . . . . . . . . . . 361
R. Labrie and N. Egan, MHD Development Corporation, Butte, Montana, and F. Walter, Montana Power
Company, Butte, Montana

Constricted Discharges in Ar-Cs MHD G e n e r a t o r s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373


W. F. H. Merck, A. P. C. Holten, A. Veefkind, and E. M. van Veldhuizen, Eindhoven University of Technology,
Eindhoven, The Netherlands, V. A. Bityurin and A. P. Likhachev, USSR Academy of Sciences, Moscow, Russia,
and B. Stefanov and L. Zarkova, Institute of Electronics, Sofia, Bulgaria

Pseudo Two-Phase Flow in an Open-Cycle MHD Generator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398


V. A. Bityurin and A. P. Likhachev, USSR Academy of Sciences, Moscow, Russia

Analysis of Flow Parameters in MHD Channel at Various Load Conditions . . . . . . . . . . . . . . . . . . . . . . . . . 410


B. Zaporowski and K. Sroka, Technical University ofPosnan, Posnan, Poland

Simulation and Comparison with the Experiment: The Dynamic Processes in an MHD Facility
Flow T r a i n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
A. M. Levints, V. R. Satanovsky, and V. N. Zatelepin, Russian Academy of Sciences, Moscow, Russia

Acceleration of Gas-Liquid Piston Flows for Molten-Metal MHD G e n e r a t o r s . . . . . . . . . . . . . . . . . . . . . . . . 431


A. Kolesnichencko and V. Malakhov, Ukranian Academy of Sciences, Kiev, Ukraine

Recent Developments in Liquid-Metal MHD Thermoacoustic E n g i n e s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 441


D. Hamann, Dresden University of Technology, Dresden, Germany, and G. Gerbeth, Central Institute for
Nuclear Research, Rossendorf, Germany

Chapter 3. Magnetohydrodynamic Flows

Interfacial Instabilities in the Presence of Electric Current and Magnetic Field . . . . . . . . . . . . . . . . . . . . . . 457
Sylvain Pigny and Rene Moreau, Madylam-Institut National Polytechnique de Grenoble, St. Martin
d'Heres, France
Survey of Liquid-Metal MHD Activities in Dresden . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
G. Gerbeth and G. Uhlmann, Central Institute for Nuclear Research, Rossendorf, Germany, and D. Hamann,
Dresden University of Technology, Dresden, Germany

Experiments with a Superconducting Magnet on an InGaSn L o o p . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476


O. Lielausis, E. Platacis, I. Platnieks, M. Pukis, and A. Shishko, Latvian Academy of Sciences, Riga-Salaspils,
Latvia

Comparison of the Core Flow Solution and the Full Solution for MHD Flow . . . . . . . . . . . . . . . . . . . . . . . . 482
Lutz Lenhart, Kernforschungszentrum Karlsruhe GmbH, Karlsruhe, Germany, and Kathy McCarthy, Idaho
National Engineering Laboratory, Idaho Falls, Idaho
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Hydrodynamics and Heat Transfer of Thin Liquid-Metal Films in a Magnetic Field . . . . . . . . . . . . . . . . . . 500
I. A. Evtushenko, E. M. Kirillina, S. Y. Smolentzev, and A. V. Tananaev, Leningrad Polytechnic Institute,
St. Petersburg, Russia

Effects of a Vertical Magnetic Field on Rayleigh-Benard Convection in Mercury . . . . . . . . . . . . . . . . . . . . . 509


Joel Stavans, Weizmann Institute of Science, Rehovot, Israel

MHD Flow Around a Cylinder in an Aligned Magnetic F i e l d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 519


J. Josserand, Ph. Marty, and A. Alemany, Institut de Mecanique de Grenoble, Grenoble, France,
and G. Gerbeth, Central Institute for Nuclear Research, Rossendorf, Germany

Electromagnetically Driven Flow Around a Cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 535


A. Thess and G. Gerbeth, Central Institute for Nuclear Research, Rossendorf, Germany, and Ph. Marty,
Institut de Mecanique de Grenoble, Grenoble, France

New Results for MHD Drag C o e f f i c i e n t s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551


G. Gerbeth, Central Institute for Nuclear Research, Rossendorf, Germany

Heat Transfer in an MHD Flow Inside a Channel with Walls of Finite Thickness . . . . . . . . . . . . . . . . . . . . . 566
Sergio Cuevas, Instituto de Investigaciones Electricas, Cuernavaca, Mexico, and Eduardo Ramos,
Laboratorio de Energia Solar, Temixco, Mexico

Instability of a Liquid-Metal Surface in a Low-Frequency Alternating Magnetic Field . . . . . . . . . . . . . . . . 580


Jean-Marie Galpin, Madylam-Institut National Polytechnique de Grenoble, St. Martin d'Heres, France,
Alfred Sneyd, University ofWaikato, Hamilton, New Zealand, and Yves Fautrelle, Madylam-Institut National
Polytechnique de Grenoble, St. Martin d'Heres, France

Chapter 4. Two-Phase Flows

Natural Convection over a Vertical Heated Flat Plate with Gas Injection and in the Presence
of a Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 601
Paul S. Lykoudis and Akira T. Tokuhiro, Purdue University, West Lafeyette, Indiana

Nucleate Boiling of Mercury in the Presence of a Horizontal Magnetic F i e l d . . . . . . . . . . . . . . . . . . . . . . . . . 626


Paul S. Lykoudis and M. Takahashi, Purdue University, West Lafeyette, Indiana

Direct Contact Heat Transfer in Two-Phase Gas Liquid Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635


C. Pisoni, C. Schenone, and L. Tagliafico, University of Genoa, Genoa, Italy

Heat and Kinetic Energy Transfer in Two-Phase Flow: Theoretical Aspects . . . . . . . . . . . . . . . . . . . . . . . . . 649
R. Mathes and A. Alemany, Institut de Mecanique de Grenoble, Grenoble, France

Nuclear Void Fraction Gaging in Large Two-Phase Organic Liquid-Metal MHD G e n e r a t o r s . . . . . . . . . . . 662
A. P. Kushelevsky, Ben-Gurion University of the Negev, Beer-Sheva, Israel

Liquid-Metal Magnetohydrodynamic Two-Phase Flow E x p e r i m e n t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667


J.-P. Thibault, B. Seek, and A. Cartellier, Institut de Mechanique de Grenoble, Grenoble, France
Investigation of Two-Phase Liquid Gas Mixers for MHD Energy Conversion S y s t e m s . . . . . . . . . . . . . . . . . 678
D. Farchi, A. El Boher, S. Lesin, Y. Unger, and H. Branover, Ben-Gurion University of the Negev, Beer-Sheva,
Israel

Two-Phase Flow Measurements in Reaction Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705


Yaakov M. Timnat, Technion—Israel Institute of Technology, Haifa, Israel

Author Index for Volume 148 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725

List of Series Volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 727


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

This page intentionally left blank


Vortex Reconnection, Cascade, and Mixing in
Turbulent Flows
F. Hussain,* D. Virk,t and M.V. Melander$
University of Houston, Houston, Texas 77004
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Abstract

Vortex reconnection is explored in order to understand coherent


structure evolution and interactions, as well as its role in fundamental
turbulence phenomena such as cascade and mixing. Due to lack of
measurement techniques for such topology-changing events,
particularly for capturing the instantaneous 3D velocity and vorticity
fields, we rely on well-resolved direct numerical simulations of the
Navier-Stokes equations. The reconnection process is found to consist of
a set of generic stages including bridging, threading and head-tail
formation. Reconnection is incomplete as the self-induced motions of
bridges and threads suggest a physical mechanism of cascade to smaller
scales via successive bridging and threading: the number of stages
increases with Reynolds number. These observations lead us to suggest
that reconnection processes play an important role in cascade,
entrainment and mixing in turbulent flows.

Introduction

Modern experimental turbulence research has focussed on three


different approaches: statistical, structural and dynamical. The
statistical approach began in the forties and was prompted mostly by
Kolmogoroff's local equilibrium hypothesis (K41). Statistical
turbulence has seen bursts of resurgence; the second major thrust was
provided by Kolmogoroff's third hypothesis (K62). In parallel, there
has been hectic activity in documenting stochastic measures such as
profiles of mean velocity, pdf and moments of fluctuations, spectra,
correlation, etc. in many basic and applied flows. Interest in structural

Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Professor, Department of Mechanical Engineering.
fResearch Assistant, Department of Mechanical Engineering.
^Associate Professor; currently, Department of Mathematics, Southern Methodist
University, Dallas, Texas.
1
2 F. HUSSAINETAL.

turbulence started in the early seventies with the discovery of large-


scale organized motions - popularly called coherent structures
(forerunners among the studies are Kline et all; Crow & Champagne2;
Brown & Roshko^). The eighties saw the birth of dynamical turbulence
with the advent of chaos as a new field of physics. It emerged,
however, with less fanfare than structural turbulence and seems to
have stayed in a simmering state for lack of any direct connection with
nonlinear dynamical systems. Dynamical turbulence seems to have
some applicability to closed flows (Rayleigh-Benard convection and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Taylor-Couette flow) in comparison to open flows. Historically, 'open


flows' are those which cross a defined flow boundary and 'closed flows'
are those which do not. We find this classical definition to be rather
limiting and have redefined these not in terms of mass crossing a
boundary, but in terms of information crossing a boundary 4 . While some
ideas have emerged under the category of 'convective chaos'5, no clear
breakthrough is in sight because of the exceedingly numerous degrees of
freedom involved in any developed turbulent flow". The dimension is
indeed so high that no clear benefit would follow from the dynamical
approach to turbulence.
We have indeed pursued all these three approaches to turbulence
(statistical, structural and dynamical) to varying degrees, but the
majority of our own efforts has been focussed on structural turbulence.
Here our approach has had a distinctly different flavor. The
overwhelming majority of studies in structural turbulence have been
based on flow visualization; however, we have long contended that
flow visualization can not only be confusing but also grossly
misleading 7 . Even at unity Schmidt number, passive scalar can
faithfully mark vorticity boundaries of structures only in the case of 2D
flow. If, however, the flow is 3D, and in particular turbulent,
visualization of vortex interactions and coherent structure dynamics
may present incorrect information because marker gets depleted from
regions of intense vortex stretching but accumulates in regions of
minimal stretching. As a result, dynamically dominant events can
easily be missed by flow visualization. Flow visualization, a highly
useful tool for qualitative flow interpretation, must be supplemented by
quantitative data.
In our continuing pursuit of structural turbulence, we build here on
our long-standing assertion that coherent structures are characterized
by coherent vorticity - the instantaneous space-correlated vorticity
underlying the random vortical field that we call turbulence - and that
vortex dynamics is a (perhaps the most) tractable avenue for
understanding the evolutionary dynamics of coherent structures, their
role in turbulent transport phenomena, their mutual interactions (such
as pairing, tearing, and reconnection) and their coupling with fine-scale
turbulence. The last is the subject of a more recent investigation8, but we
limit our discussion here to the material discussed during the seminar.
VORTEX RECONNECTION
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 1 I CO I surfaces at 30% of the initial peak: a) t = 0; b) t = 2.25; c) t - 35; d) t


= 4.75; and e) t = 6.0.
F. HUSSAIN ET AL

Our interest in the reconnection process was prompted by our


intuitive claim that it is a frequent event in turbulent flows 7 , that it
plays an important role in cascade and mixing, that it is an important
source of aeroacoustic noise in hydrodynamics 9 , and that, while
viscosity is crucial to reconnection, the phenomenon occurs at a
convective, rather than a diffusive, timescale. Our curiosity was also
piqued by our interest in the topological changes of vortex lines during
reconnection. As an experimental technique that can provide well-
resolved spatial and temporal data on this highly 3D vortex
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

interaction is not available, we considered fully resolved direct


numerical simulation of this interaction in an idealized configuration.
In addition to being the simplest, this configuration was motivated by
our experimental studies of reconnection of elliptic vortex rings in high
aspect ratio elliptic j

Simulation Method and Initial Conditions

The flow is simulated using a spectral method dealiased by a 2/3


k-space truncation and with a fourth-order predictor-corrector
algorithm for time stepping. Periodic boundary conditions are
appropriate because reconnection is a strong local vortex interaction,
insensitive to nonlocal effects such as the influence of adjacent boxes.
Tests were undertaken to check that the evolution details remained
unchanged when the domain was doubled. The initial condition
consists of two antiparallel vortices with mutually inclined symmetric
sinusoidal perturbations (Fig. la). The sinusoidal perturbation is
provided only to ensure that the two vortex tubes collide so as to induce
and sustain reconnection; this perturbation is not central to the flow
physics. Initially, each vortex has a circular core with a compact
support Gaussian vorticity distribution. We also studied an asymmetric
initial condition as a perturbation to Fig. la. No change was noticeable
in the details of the evolution of the vortex reconnection process, thus
confirming that imposition of symmetry is not a constraint of the
simulation (see later). The symmetric initial condition is particularly
useful for clearly identifying the effects of viscosity because the two
symmetry planes are preserved in time and are material planes in the
absence of viscosity. For clarity, the xy-plane is called the 'symmetric'
plane or 7ts, while the yz-plane is the 'dividing' plane or TUtf. In both
planes, the normal velocity and the tangential vorticity vanish at all
times. Furthermore, xCOz < 0 everywhere in 71s; this eliminates the
problem of self-annihilation within each vortex which occurs in the
simulation of Meiron et al.^. Consequently, the circulations Fs and Fj
in the half planes 7U S + (x>0) and TUd" (z<0) change exclusively by
viscous vorticity annihilation. We also know that dF^/dt = - dF s /dt,
since an arbitrary vortex line must intercept either 7ts or Tltf-
VORTEX RECONNECTION 5

Three Phases of Reconnection

The evolution of the vortices is shown in Fig. 1 for the Reynolds


number Re=r/V=1000. Fig. 1 shows wire-frame plots of the vorticity
norm Icol at t=0, 2.25, 3.5, 4.75 and 6; the level surface here represents
30% of the initial peak vorticity CO max (0), and t is time, t ,
nondimensionalized by C0 ma x(0) such that t=t*CO ma x(0)/20. Fig. 2
shows the evolution of circulation Ts. On comparing Fig. 1 with Fig. 2
it is clear how reconnection evolves through three phases.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

The first phase is governed by inviscid vortex dynamics as almost


no circulation is transferred from 7CS to 71^- During this phase, while
the two vortices move upward by mutual induction, self-induction
presses the vortices against each other, flattening them and forming a
contact zone with steep vorticity gradients. In the second phase, a
dramatic change in topology takes place as a rapid transfer of
circulation from 7ts to K& (Fig. 2) occurs due to viscous vorticity
annihilation (by cross diffusion) in the contact zone. This phase,
marked by 'bridging' (growing humps in Figs. lc,d), is the heart of the
reconnection process (discussed later). The growing bridges acquire a
curved horseshoe shape, causing them to pull apart by self-induction.
Simultaneously, the induced flow by the bridges stretches the dipole in
7TS and sustains their annihilation. The third phase is the evolution of
the remnants of the original dipole after the rapid circulation transfer.
The slender vortices, which we call 'threads', are well defined because
of the stretching by the bridges. These threads arise as a consequence of
the incompleteness of the reconnection and are the longest lasting
features of the reconnection.

1.2

i.o

0.8

r/r
o 0.6 Phase I
Inviscid advection
0.4

0.2

J___I
2 3
t

Fig. 2 Circulation in 11$+ (x > 0); phases of recognition.


F. HUSSAIN ET AL.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

b)

-8
d)
-7

-6

-8
SOOi -7

Fig. 3 I CO I contours in %; contour increments A CO = 4: a) t = 0; b) t = 3.0;


c) t = 3.75; d) t = 4.5; and e) t = 6.0.
VORTEX RECONNECTION

I I I !

a) -7

-6

—— 9
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

b)

c)

d)

Fig. 4 I CO I contours in KJ; contour increments AO) = 4: a) t = 3.0; b) t = 3.75;


c) t = 4.5; and d) t = 6.0.

The simulation with initial asymmetries produced virtually the


same wireframe plots as in Figs, l(a-e) even at t=6, thus vindicating
both the analysis9 and the high-resolution simulation12 performed by
invoking symmetry. The Re dependence of the crosslinking process was
tested and found to consist of the same sequence of stages with identical
successive spatial patterns, except that the corresponding timescales
increase with decreasing Re.
F. HUSSAIN ET AL

Bridging Mechanism

The rapid circulation transfer from 71s to 71 d begins when the


vortices come into contact near 7ls ^ ^d (Figs. lb,3b) and form a contact
zone C(t) with steep vorticity gradients. As the flow is highly viscous
around C and nearly inviscid elsewhere, we may think of the vortex
lines far away from C, merely for the sake of discussion, as material
lines. We use this idea schematically in Figs. 5a-c, where each vortex
is represented by three vortex lines; one vortex line from each vortex
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

intercepts C (Fig. 5a), and shortly afterwards these vortex lines


annihilate each other at x=z=0 by cross-diffusion. This produces two
reconnected lines intercepting Tltf at z=±zo (Fig. 5b). The Tl^-intercept of
zo quickly retreats away from x=z=0, because of strong diffusion and
nearly antiparallel vorticity lines emerging from cusps.
A cross section AB (Fig. 5c) at the edge of contact zone C reveals
the two-dimensional dipole structure of the original vortices. In a co-
translating frame the streamline pattern becomes the well-known
recirculation bubble of a vortex dipole (Fig. 5d). Although this frame is
not precisely defined for an unsteady dipole, the existence of two
stagnation points — at the front and back of the dipole — Sp and Sg can
be inferred. The large velocity between Sp and SB advects zo towards
Sp, ahead of the dipole (Fig. 5c). The relative upward advection of zo
stops at Sp/ and a reconnected vortex line unfolds along the diverging
separatrix a-a (Fig. 5d). This mechanism is schematically illustrated
in a perspective view of one half of the two vortices in Fig. 5e, which
shows a vortex line at five different instants (1-5). Far away from C, a
vortex line is simply swirled around the vortex tube. However, in C,
vortex lines get reconnected as soon as they come into contact after 1, but
before instant 2 and are swept upward by the pumping motion of the
dipole. At instant 3, the upward motion of the reconnected vortex line
stops at Sp, and it gets stretched and begins to wrap around the dipolar
vortex pair near Sp (vortex line at instants 4 and 5). Meanwhile, the
circulatory motion in the dipole continuously pumps fresh non-
reconnected lines into C near SB (Fig. 5d) which then in turn go through
the same sequence of reconnection. Due to this mechanism, reconnected
vortex lines accumulate ahead of the antiparallel vortices and align
orthogonal to the latter.
During bridging, a complex interaction of three competing effects
governs the rate of annihilation of vorticity by changing its gradients
in C. These effects are diffusion, stretching in the z-direction, and local
'self-induction'. Diffusion counteracts the steepening of vorticity
gradients and can only decrease the peak vorticity in the symmetry
plane C0m, while any increase in C0m is due to stretching (normal to 7ls).
This 'axial 1 stretching is generated mainly by the vorticity distribution
e(t) away from C (£ is shown in Fig. 5a; note that £ includes the
bridges), and in part by self-induced lengthening of vortex lines near C
VORTEX RECONNECTION

Contact Zone
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

External Upward
Distribution e velocity

SYMMETRIC PLANE

Fig. 5 Explanation of bridging mechanism (a-c), streamline pattern In 1% (d), and


schematic of the perspective view of one half of the two vortices; a vortex line is
sketched at 5 different instants to illustrate the bridging mechanism (downwash
ath).
10 F. HUSSAIN ET AL

(analogous to the axisymmctric collision of vortex rings). Depending on


the relative positions of £ and the dipole in 7ts, the external stretching
is either positive (vortex stretching) or negative (vortex compression):
positive if the dipole has fallen behind e (Fig. 5f), and negative if the
dipole has advanced ahead of £. The self-induction on each side of the
sheet-like region C is in 7ls, and is directed toward Tttf whenever the
non-reconnected vortex lines in C curve upward, and away from lift when
they curve downward. The motion toward 71$ increases gradients across
the y-axis by compressing the vortex cores into a characteristic head-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

tail structure (Fig. 3d), while motion away from TCd decreases the
gradients by separating the vortex core in 7ts. The above three effects -
diffusion, stretching and self-induction - dominate the evolution in the
following order. At first, the vortex lines in C curve upward, and the
gradients across 71^ increase as self-induction toward K& presses the
dipole cores into a head-tail shape; meanwhile, the external
stretching being negative, the increase in C0m is small. However, as the
dipole in 7CS falls behind 8 (discussed later) such that the vortex lines
in C curve downward, the axial stretching picks up and produces a large
increase in C0m, and consequently large gradients. Finally, towards the
end of the rapid circulation transfer from 7CS to n& (Fig. 2), the local
self-induction away from n^ arrests the cross-diffusion by producing a
near balance between axial stretching, diffusion and separation of the
dipole cores in 7TS> causing the threads to linger.

Curvature reversal and reconnection cascade

The curvature reversal, which causes the incomplete


reconnection, is a characteristic and crucial event. It occurs because the
upward motion of the dipole in 71 s slows down relative to that of £
(Figs. 3,4; note that the y-position can be compared using the numbers on
the right side of each panel); but why does the dipole in 7CS slow down?
A downward velocity field in C — downwash — opposes the dipole's
own upward motion in K$ (Fig. 5f). Furthermore, the dipole loses
circulation through vorticity annihilation; its upward motion weakens
and it relaxes into a head-tail shape. The downwash is induced by the
growing bridges as well as by non-reconnected vortex lines in £, so
downwash will be present even at Re=<^>. The relaxation into a head-
tail shape, an inviscid process, virtually decreases the dipole
circulation as the dipole propagation velocity is determined almost
exclusively by the head. Although this inviscid effect is weak at
Re=1000, it may well be the primary mechanism for curvature reversal
at higher Re. An indication of this would be that the area A of the
nearly self-similar head decreases with increasing Re, such that
COmAl/2 (propagation speed of the head to leading order) is bounded as
Re —><>o. This indicates the possibility that the amount of circulation
VORTEX RECONNECTION 11
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

b)

Fig. 6 Helicity density h at a) t = 4.5,20 % of peak value, and at b) 50% level.


Black regions correspond to negative h, and white regions to positive h.

transferred from 7ts to Ttd tends to zero as Re tends to infinity while the
transfer is still occurring within a convective timescale.
The above gives us some interesting new results. The vortex
reconnection is not complete in a single step; as the bridges move apart
by self-induction, the thread-dipole advects forward in the presence of
the weakening 'downwash 1 caused by the bridge-dipole. This way, the
threads reverse their curvature again and then collide at the midpoint
by self-induction, starting the next burst of reconnection. This is what
we call reconnection cascade. Clearly, the frequency of reconnection
bursts will increase and circulation transfer will decrease as the
Reynolds number increases.

Dynamical Interactions between Bridges and Threads

In order to provide further insight into the reconnection


mechanism, topology, and dynamics of interacting vortices, we discuss
the evolution of helicity, enstrophy production and vortex lines. Even
though helicity density (h) is a frame-dependent quantity and hence
not a suitable measure of topology (unlike the Helicity integral), in the
present idealized configuration it does point out some important
features. Helicity in the initial condition and during the first phase of
reconnection is very small and vortex lines inside the vortex tubes are
almost parallel. However, the helicity distribution after the
reconnection (e.g. t=4.75 and t=6) is particularly interesting, as it
consists of intertwined regions of positive and negative helicity. Apart
from the threads, where the helicity stands out clearly (Figs. 6a,b),
the structure is an enigma until one examines bundles of vortex lines.
12 F. HUSSAIN ET AL

Figs. 7a,b show that vortex lines emerging from the threads do not wrap
all the way around the bridges as one would expect from Icol surfaces,
e.g. expanded views of Figs. ld,e. Instead, they form asymmetric
hairpin-like structures (see A and B, Fig. 7a). The vorticity is high in
one hairpin leg and low in the other. Vortical structures of this kind
are unfortunately not faithfully reflected in the surface plots of Icol.
The intense leg (A) is the extension of the thread (C) and its high
vorticity makes its helicity stand out in Fig. 6. The diffuse leg is
difficult to identify in the helicity distribution as the Icol there is low
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

and the angle between vorticity and velocity is larger. The intense leg
induces a flow away from 1Z& on the outside of the bridges causing vortex
stretching in the outer part of the bridges (see point A in Figs. 8b,d).
Inside the bridge, the vortex lines have a slow helical twist near
Ttd (barely discernible in Fig. 7b). Contrary to the intense hairpin leg,
the twist produces an axial flow toward TUtf at the center of the bridge.
Since this axial flow is toward 7l<j from both sides, it results in a
negative enstrophy production, as seen in a cross section of bridges
parallel to TC^ (dotted lines in Figs. 8b,d). The twisting of vortex lines is
the result of a skewed (nonconcentric) vorticity distribution in the
bridges (Figs. 4b-d). The peak vorticity in the bridge is not at the
location of the geometric centroid of the low level vorticity contours,
but is far away from the contact zone, where the vorticity is lower than
in the bridges - a consequence of earlier stretching by the dipole during
bridging. Figs. 7a,b show that a vortex line which neatly follows the
centerline of the vortex far away from 71^ ceases to do so in the bridge.
This effect clearly results in vortex line twisting and thus in an induced
axial flow (Fig. 7c). The direction of the axial flow is determined by
the orientation of the twist, which in turn is determined by the highest
swirling velocity along the vortex. The bridges, having the highest
peak vorticity, also have the highest swirling velocity. Hence the
axial flow is toward K& from both sides, resulting in vortex compression
(i.e. negative PQ in Figs. 8b,d) which decreases the peak vorticity and
thereby also the twisting rate. This constitutes a new inviscid
mechanism, clearly distinct from diffusion, for smoothing the vorticity
intensity along the reconnected vortex lines.
Interacting vortex tubes relax to a head-tail structure as shown in
Fig. 3d. This is essentially an inviscid mechanism and was also
observed in the head-on collision of inviscid vortex rings-^/14. The
principal reason for head-tail formation is the strain in rcs due to self-
induced motion of the vortex tubes toward n$. Because of higher
vorticity concentration, the annihilation is higher between the heads
than between the tails. The tail decays more by viscous diffusion than
by annihilation. The higher circulation in the head also causes it to
move faster due to mutual induction and around t=4.5, we see a
separation of the head from the tail (Fig. 3e). Such deformation of a
circular vortex core into a head-tail structure increases the surface area
VORTEX RECONNECTION 13

A ,C
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

thread
b)

twist

ccntcriinc

Fig. 7 Vortex lines in a quarter of the computational domain; also shown are the
plane cross sections of the vorticity norms in %, KA and in the box side. A sketch
of the vortex has been overlaid to orient the reader: a) t = 4.75; b) t = 6.0; and
c) axial flow due to vortex line twist.
14 F. HUSSAIN ET AL

peak
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

c) d)

peak

Fig. 8 Plane cross section parallel to 714 but moved Ax = 0.196: a) 10)1 at t = 4.75;
b) enstrophy production P^, at t = 4.75 (chain-dotted lines are I CO I contours);
c) I CO [ at t = 6.0; and d) P at t = 6.0 (chain-dotted lines are I CO I).

over which vorticity diffuses to make the surrounding fluid rotational,


and thus, increases entrainment. This is in addition to the increase in
surface area due to the formation of bridges which also contribute to
entrainment via strong 'downwash'. In the present simulations, one of
the vortices was marked with a scalar with Schmidt number unity.
Even under such ideal conditions the scalar does not continue to
faithfully mark the vortical fluid (compare Figs. 1 and Fig. 9). At
t=4.75, vortex stretching has diluted the scalar in the contact zone in A
(Fig. 9d) while the vorticity, being accentuated, remains large, with
the result that a key feature of the vortical structure, the threads, is
VORTEX RECONNECTION 15
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 9 Scalar at 5% of the initial peak: a) t = 0; b) t = 225; c) t = 35; d) t = 4.75;


and e) t = 6.0.

invisible in experimental flow visualization (Fig 9e; also see


Schatzle 15 ). This serves as another example supporting our long-
standing warning against relying too heavily on flow visualization in
studying 3D vortex interactions and turbulence dynamics.

Concluding Remarks

The reconnection process thus represents successive stages of


bridging and threading. In addition, the original vortex dipole with
circular cross section deforms into the head-tail structure. Because of
the much higher circulation, the heads pull ahead faster, leaving the
tail behind. Combined with the vorticity annihilation by viscous
cross-diffusion, the head-tail structure causes both a decrease of vortex
16 F. HUSSAIN ET AL

size (the threads) as well as increase in vortex surface. Generation of


bridges orthogonal to the original vortices also effectively mixes the
surrounding fluid. Thus reconnection is associated with both cascade
and mixing. Thus, we present here a physical mechanism for cascade,
entrainment and mixing, which, in spite of their widely acknowledged
significance in turbulence phenomena, have remained elusive so far.

Acknowledgment
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

This research was supported by the Office of Naval Research


and the Department of Energy.

References
1
Kline/ S.J., Reynolds, W.D., Schraub, F.A., & Runstadler, P.W., "The
structure of turbulent boundary layers", /. Fluid Mech., 30, 1967, p741.
^Crow, S.C. & Champagne, F.H.,"Orderly structure in jet turbulence", /.
Fluid Mech., 48, 1971, p547.
^Brown, G.L. & Roshko, A., "On density effects an large structure in
turbulent mixing layers", /. Fluid Mech., 64, 1974, p775.
4
Broze, J.G., "Chaos in an 'open1 flow: experiments in transitional jets", PhD
thesis, U. of Houston, 1992.
-^Dcissler, R.J., & Kancko, K., " Velocity-dependent Lyapunov exponents as
a measure of chaos for open-flow systems", Phys. Lett. A, 119, 1987, p397
°Keefc, L., Moin, P. & Kim, ]., "The dimension of an attractor in turbulent
Poiscuillc flow", Bull. Am. Phys. Soc., 32, 1987, p2026.
'Hussain, F., "Coherent structures and turbulence", /. Fluid Mech., 173,
1986, p303.
8
Melander, M. and Hussain, F., "Vortex core dynamics, helical waves and
organization of fine-scale turbulence", (submitted) 1992.
^Takaki, R. and Hussain, F., "Recombination of vortex filaments and its
role in aerodynamic noise",Turb. Shear Flows V, Cornell U., 1983, p3.19.
^Hussain, F. and Husain, H., "Elliptic jets. Part 1. Characteristics of
unexcited and excited jets.",/. Fluid Mech., 208, 1989, p257.
-^Meiron, D.I., Shelley, M.J., Ashurst, W.T. and Orszag, S.A., "Numerical
study of vortex reconnection", in Mathematical Aspects of Vortex Dynamics,
SIAM, Ed R. Caflisch, 1989, p.183.
-^Kerr, R.M. and Hussain, F., "Simulation of vortex reconnection", Physica
D, 37, 1989, p474.
l%hariff, K., Leonard, A., Zabusky, N., and Ferziger, J. "Acoustics and
dynamics of coaxial interacting vortices", Fluid Dyn. Res., 3, 1988, p337.
-^^Stanaway, S., Shariff, K. and Hussain, F., "Head-on collision of viscous
vortex rings", NASA report CTR S-88,1988, p287.
^Schatzle, P.R., "An experimental study of fusion of vortex rings", PhD
thesis, Cal. Inst. of Tech., 1987.
Intermittent Turbulence from Closures
Robert H. Kraichnan*
569 Montezuma 108, Santa Fe, New Mexico 87501
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Abstract

Turbulent flows universally exhibit intermittency of small scales. Analyt-


ical turbulence theories that deal with moments of the velocity distribution
have been unable to capture the phenomenon. This paper reviews a new
theoretical development that manipulates entire probability distributions
rather than moments. The key tool is the construction of a nonlinear map
of a Gaussian reference field in order to produce a tractable stochastic
model of a dynamically evolving non-Gaussian field. The resulting closure
approximations are systematic and wholly nonperturbative. They yield in-
termittency phenomena in a natural way. The intermittency mechanism
is effective at low as well as high Reynolds numbers. An application to
Burgers' equation predicts a nearly exponential tail to the velocity-gradient
distribution and a power-law tail to the fluid-density distribution. Both are
verified by simulations. Implications for Navier-Stokes flow are discussed.

1. Introduction

Most turbulent flows exhibit intermittency of small scales.1"18 This is


true even if the Reynolds number is not large. The intermittency in the
interior of a turbulent flow is evidenced by spatial spottiness of dissipation,
associated with sparse distributions of intense vorticity in the form of sheets,
tubes, or more complex structures.
A number of heuristic models of intermittency have been based on the el-
ementary vortex-stretching mechanism: stretching (antistretching) gives in-
tensification (diminution) of vorticity. Repeated, stochastically distributed,
strainings are expected to yield intermittent vorticity by random composi-
tion of stretching ratios. This picture leads to log-normal type distributions
in the absence of viscosity. Saffman19 has given a simple model:

= a(t)u (1)

Copyright © 1992 by Robert H. Kraichnan. Published by the American Institute


of Aeronautics and Astronautics, Inc., with permission.
* Researcher.
17
18 R. H. KRAICHNAN

where u is the magnitude of vorticity in a fluid element and a(t) models the
fluctuation in time of the stretching term. The predictions of this model
are strongly affected if viscous damping is added and the nonlinearity of
the stretching process is taken into account.
Analytical approximations that attempt to proceed systematically from
the Navier-Stokes (NS) equation have, in the past, failed to capture the
intermittency of small scales. An exception is the work of Qian 20 which
predicts non-Gaussian fourth-order moments at the cost of violating realiz-
ability constraints. Stochastic modeling, like the direct-interaction approx-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

imation (DIA), has succeeded in describing certain non-Gaussian statistics:


the skewness of velocity derivatives and a depression of the nonlinear term
in the NS equation.21 But neither this approach nor any other based on
moment approximations has yielded the intermittency of dissipation.
A common feature of the systematic moment approximations that have
been offered is a dependence on perturbative treatment. Stochastic models
like DIA can be formulated nonperturbatively, but the final equations are
derivable only because of tricks that enable exact* treatment of the model
by perturbation methods. Moreover, the final equations retain an intimate
relation to perturbation expansion of moments of the exact velocity field
about a Gaussian distribution. 22 The intermittency of small scales involves
strong departure of the statistics of velocity gradients from Gaussian form.
Perturbation approaches seem intrinsically unsuited to this problem, and
it is not surprising that they have failed.
I shall attempt here a brief pedagogical review of a new analytical ap-
proximation method 23 ' 24 that seems well-suited to the exploration of in-
termittency. It is intrinsically non-perturbative, and it deals with entire
probability distributions (pdf's) rather than with low-order moments. The
key concept is the modeling of non-Gaussian fields by a simple mapping
of a Gaussian field. The latter is called the reference field. Statistics of
the non-Gaussian field can then be evaluated by exploiting the explicitly
available statistics of the reference field. This leads to systematic closure
approximations. The modeling technique can easily handle fields that have
highly singular or violently intermittent pdf's.
Later in this paper, an analytical mapping closure is developed for Burg-
ers' equation. An heuristic model for intermittency in incompressible NS
turbulence can be written by drastically simplifying this closure and remov-
ing some effects that arise from the compressiblity of the Burgers' fluid.
This heuristic model is qualitatively different from (1). It is potent even
at low Reynolds numbers, where there is no opportunity for a sequence of
random stretchings.
Suppose that the initial velocity field is multivariate-Gaussian, so that
the components of the rate-of-strain tensor also are Gaussian. The heuristic
model equation for the evolution of a strain component s is

* = J(so)*o (2)
INTERMITTENT TURBULENCE 19

dJ/8t= \s0\J2 -T]k2dJ3 (3)


Here SQ is the initial, Gaussianly distributed, value of s. J is an overall
effective stretching ratio for regions of the field with given SQ, ^ is kinematic
viscosity and the parameter k^ is a characteristic dissipation wave number.
J > 1 represents the intensification of gradients.
The interpretation of (3) is as follows. The first term on the right side
represents nonlinear straining effects. J appears as square because the non-
linearity is quadratic. Alone, this term would lead to extreme intermittency,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

and a singularity at finite time for any given value of SQ . The second term
on the right side represents viscous decay with a time constant inversely
proportional to the square J 2 of the stretching ratio. The third J factor
in the viscous term comes from (2). The ratio (so) 1 / 2 / 7 ?^ *s an en°ec^ve
Reynolds number. The only other parameter for the model is the normal-
ized evolution time (s 2 ) 1 / 2 ^ The decay of velocity amplitudes is ignored
here, so the model is properly applicable only over times short compared to
overall decay times. (Velocity decay effects are included in the systematic
closures to be presented in § 3.)
At large enough values of \SQ\, (3) quickly leads to a near equilibrium in
which J grows to make the two terms on the right side balance; thus, at
large s,

.7 « so/to**), M««?/fo*2). ./«H 1 2 /(^) 1 2 (4)


The pdf of 5 is
(5)

where
Po(So) = (2^ S 2 0 ))- 1 / 2 exp(-i^ /(s 2 )) (6)

is the Gaussian pdf of SQ and { ) denotes ensemble average.


Eqs. (4) and (5) imply that P(s) at large enough \s\ has the form

The nearly exponential tail exhibited in (7) recalls the exponential-like


tails that have been observed in incompressible NS simulations.5'9'11"13 Fig-
ure 1 shows the pdf of transverse velocity derivative found in an isotropic-
turbulence simulation by Vincent and Meneguzzi13 at a macroscale
Reynolds number of about 1000. Figure 2 shows the pdf (5), evolved
at t = 0.3 under (2) and (3) from a start at t = 0 with s = s 0 » ($o) = 1>
rjk'j = 1.5. The model reproduces not only the broad skirt of Fig. 1 but
also the characteristic upward curvature of the skirt. It is notable not only
that the match is possible but that it is obtained with values of the two
20 R. H. KRAICHNAN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-16 -14 -12 -10 -

FlG. 1 Normalized pdf of transverse velocity derivative in a simulation of


isotropic turbulence at macroscale Reynolds number about 1000. After Vincent
and Meneguzzi. 13

NORMALIZED AMPLITUDE

FlG. 2 Normalized pdf (s2)1/2P(s) for model (2), (3) vs normalized rate-of-
strain amplitude s/(s2)1'2. See text.
INTERMITTENT TURBULENCE 21

FlG. 3a The function ^o(x) = sin(2Tx) or ^o(^) = sin(27T2r)(k


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

FlG. 3b The function ilj(x) produced by the transformation (8), (9) on the -
of Fig. 3a.

FlG. 3c The function rf(x) produced by the transformation (11), (15) on the
tpo of Fig. 2a.

model parameters that correspond to a Reynolds number and a normalized


evolution time both of order unity. This hints that the essential intermit-
tency mechanism in NS dynamics may not require high Reynolds number, a
possibility that will be discussed further in § 4. The unrealistic sharp peak
at the pdf maximum in Fig. 2 is the result of some extreme simplification
in writing (3). This is discussed after Eq. (52).

2. Mapping Closures
The mapping transformation of a Gaussian field involves two elementary
operations that may be used singly or in combination: first, the alteration
22 R. H. KRAICHNAN

of the field amplitude at each point in x space according to some few-


parameter functional relation; second, squeezing, stretching or other dis-
tortion of the field by a transformation of the space in which the Gaussian
field lives. These operations will be illustrated here by simple examples that
lead up to two applications with spatially homogeneous statistics: the de-
cay of a stochastically distributed scalar field under the heat equation, and
the evolution of a one-dimensional velocity field under Burgers' equation.
The simplest amplitude mapping of a scalar field makes the transformed
amplitude -0(x) depend only on the amplitude ^o(x) of the reference field
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

at the same point in space:

V> = X(1>0) (8)

where dX/d^Q > 0 in order to make the mapping single- valued. For exam-
pie, if
= |lMl + W) (9)

then the sinusoidal i^o(x) shown in Fig. 3a is transformed to the shape


shown in Fig. 3b.
The coordinate mappings to be considered here carry a reference field
^o(z) into a field V>(x) where x is position in laboratory coordinates. If
the space dimensionality is one (all that will be discussed here) and the
coordinate transformation at each point in space depends only on ^o and
£o = di/jQ/dz, the transformation may be taken in the form

where J > 0 to assure single- valuedness. Eq. (10) implies

£o) (ii)
If the transformations (8) and (10) are applied together, then (11) is re-
placed by

* = toJ(1>o,to)dX/d1>0 = Y(tfo,&). (12)


Figure 3c illustrates the effect of a space distortion on the sinusoidal field
of Fig. 3a. The reference field is taken as

V> 0 (z) = sin(27T£) (13)

where z is coordinate value in the reference space, and the distortion is


defined by

x = Q (z = 0), J=f(£0>0), J = 2(£0<0) (14)


INTERMITTENT TURBULENCE 23

The pdf induced by (8) is

(15)

where Po(V'o) is the pdf of the reference field. On the other hand, if (8) and
(10) are applied to a Gaussian field, the joint pdf of ^ and £ is
-i
1
(16)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where Qo(£o) is the Gaussian pdf of £ 0 , and N(t) normalizes P(^>, £) to


unity. Recall that amplitude and gradient of a homogeneous, Gaussian field
are statistically independent. The square bracket in (16) is the Jacobian of
the transformation between ^>o,£o and V>,£ while the 1/J factor represents
the weighting of probability associated with squeezing and stretching. If J
is independent of ^o then the marginal pdf P(VO is independent of J. If,
in addition, X — ^o, then P(if>) = PO(^O)-
Figure 4 shows the pdf induced by (9) acting on a Gaussian field with
(^o) = 1- As in Fig. 2, the quadratic relation between ^ and ^0 induces

normalized x

FIG. 4 Normalized pdf (^ ) P(^) vs </>/<^ 2 } 1/2 after applying the transfor-
2 1/2

mation (8), (9) to a Gaussian field i/>Q.


24 R. H. KRAICHNAN

exponential tails on the pdf. Figure 5 shows the marginal pdf

Q(t) = (17)

induced by (14) acting on the same Gaussian field. (With this «/, N —
1). Comparison of Fig. 3 with Figs. 4 and 5 illustrates that the effects of
transformation on the pdf can be stronger than might be guessed by looking
at the change in segments of individual realizations.
The heat equation
«VV (18)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

is linear, and it was solved completely in the 18th century. Nevertheless,


(18) presents a fundamental closure problem at the pdf level. If the evolu-
tion of the one-point pdf P(VJ) is sought, one must deal with the fact that
the statistics of the diffusive term /cV2^ depend on the three-point pdf.
But it is intuitively plausible, and easy to show,23 that the evolution of any
single-point function g($) under (18) depends only on the conditional mean
of the diffusive term at given Vs which will be denoted here by tf[V 2/ 0] £:</,.
This mean can be evaluated straightforwardly if an approximation is made:
namely, that the evolved field ^(x) is related to a Gaussian reference field
by a transformation of the form (8). Closure is then obtained by choosing

normalized x

FIG. 5 Normalized gradient pdf (£2}1/2Q(£) vs f/{£ 2 ) 1/2 after


applying the
transformation (11), (15) to a Gaussian field ^>o«
INTERMITTENT TURBULENCE 25

X to yield, at each 2, the P(V>) that evolves under (18). That is,

The evaluation of the conditional mean goes as follows. Differentiation of


(8) with respect to x yields

f\ -T/" Q •yr ^2 V

V ip = ——V yjQ H~ ~z—o" v^ol


(j J\. rt Oy\. n O Y\ . ,r\ /c\r\ 1 \
\\p — ——V^o> (20a, b)
Ul/JQ dt^Q V^Q
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Now the conditional mean of the right side of (20b) can be evaluated
immediately by using the fact that conditional mean at fixed \l> is equivalent
to conditional mean at the value of V>o given by (8). The conditional means
of the reference field are fixed by homogeneity and the fact that statistics
of a Gaussian field are wholly determined by covariances. Thus, V^o is
statistically independent of V>o at the same x because (^oV^o) — 0 by
homogeneity. Also,
(VWVo) = -(|VV>o| 2 ) (21)
which implies
[V2V>o]c:^ = -V-o(|VV>o|2)/{V>02) (22)

All of this, together with (19), gives the final closure equation

(23)

that fixes the evolution of X. The evolution of P(^) then follows from (15).
In some cases the closure can be improved, at low cost, by including
a degenerate form of the J transformation in which J depends solely on
time. Then N(t) = J(t), (16) reduces to (15), (19) is unaltered, a factor J
appears in the right side of (20a) and a factor J 2 appears in the right sides
of (20b) and (22). Eq. (23) then becomes

^ = /cJ 2 (|V^o| 2 )(-7^^ + ^ ) (24)

= -2«{(V2V02) (25)

which follows from (18), upon differentiation and partial integration. If


(22), with the J factors included, is substituted into (25), the resulting
equation for J 2 is

(26)
dt dt
26 R. H. KRAICHNAN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

SCALAR AMPLITUDE

FIG. 6 Relaxation of pdf (C 0 ) 1/2 P(^) vs ^/(C 0 ) 1/2 from double-6-function


initial form as given by Pope's analytical solution of (23) (see text). Times
shown are / « 0.4/ c , t = tc, t w 2.0*c, t « 3.9*c. The pdf value at ^ = 0
rises monotonically with t. The value at |^| = {^o}1^2 is infinite for t < tc and
zero for t > tc.

where
(27)

(28)

and () in (26) and (27) denotes average over PQ(I/>Q).


The transformation X defines a trajectory in ij> space. It must be em-
phasized that this trajectory does not represent the ip history of a typical
fluid particle that starts with ^ = V'o- Instead, X is an effective mean
trajectory that yields self- consistent evolution of -P(VO- Ari initially mul-
tivariate Gaussian distribution remains multivariate Gaussian under the
heat equation. This remains true under either of the closures (23) or (24)-
(28). The appropriate initial transformation is simply X(ipo,t = 0) = -00-
Closure (24)-(28) has the additional property that the evolution of P(V > , t)
(decay of the Gaussian amplitudes) then is exactly correct, provided that
the wave-number spectrum of ip has the self-preserving form

= 0) oc£ n exp[-|6(<)*2] (n > -1) (29)


INTERMITTENT TURBULENCE 27

In this case, (24)-(28) reduce to

X(i/>Q, t) = r(*)t/> 0 , dr/dt = -Kj2rCi/C0 (30)

(31)

The initial value


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(</><>< 0)

gives the singular initial pdf

,t = 0) = |[«ty - (y>02>1/2) + *(V> + (V>o2}1/2)] (33)

Under this initial condition, (23) admits an analytical solution, due to


Pope:25
t) = C01/2erf[a(t)t/;o/C01/2] (34)
with
[a(<)]2 = i[exp(2««71/C0) - I]'1 (35)
As X evolves, the initial abrupt step at t/>o = 0 smooths and diffuses. The
corresponding evolution ofP(ij>,t) is shown in Fig. 6, at a series of successive
times. At the time tc such that [a(tc)]2 — 1/2, P(^,t c ) is perfectly flat
between the end points zb^g) 1 / 2 . Evolution remarkably similar to Fig. 6
appears in three-dimensional computer simulations with an approximation
to (33) as initial condition.26 However, there is evidence that the shapes of
the true relaxation from the 6-function initial condition depend on space
dimensionality and on initial wave-number spectrum. It seems possible
that the degree of dependence on initial spectrum decreases with increase
of space dimensionality. Spectrum form and space dimensionality can affect
the pdf shapes given by the mapping closure if the J(V'O)^o) transformation
is employed in addition to the X(ipo) transformation. The analysis will not
be given here.
Two noteworthy features of the mapping closure are illustrated by Pope's
analytical solution. The first is that the closure gives stable relaxation
toward an eventual Gaussian pdf. This property of the heat equation has
been widely observed in the laboratory and in computer simulations but has
hitherto proved notoriously difficult to capture in closure approximations.27
Another feature, also hard to keep in closures,27 is that the amplitude
\l) never seeps outside the initial extremal values, here i^ 2 ) 1 / 2 - Both
features can be predicted from the nature of the mapping representation.
At each t, (8) represents the actual field by a realizable model field. In
each interval dt this model field evolves under the exact heat equation.
Physically impossible results therefore cannot arise. Also, since the model
28 R. H. KRAICHNAN

field is obtained by a local distortion of the Gaussian reference field, it is


plausible that relaxation to a Gaussian pdf survives in the closure.

3. Burgers1 Equation
Burgers' equation is
V^/Vt = rpl>9X (36)
where (for consistency with preceding notation) i/s(xyt) is the the one-
dimensional velocity field, T>/T>t = d/dt -f ^d/dx and 77 is kinematic vis-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

cosity. Differentiation of (36) with respect to x yields

Vt/Vt = -e + tfxx (37)

where £ = i/>x. Burgers' equation describes an infinitely compressible flow,


in which the density p(x,t) obeys

Vp/Vt = -Sp (38)

The initial condition p(x,Q) = I is assumed.


The dominant behavior under Burgers' equation is the steepening of
velocity gradients into viscosity-limited shocks.28 During this process an
initially multivariate-Gaussian field $ retains a nearly Gaussian univari-
ate distribution P(ip) while the univariate distribution Q(f) of the gradi-
ent £ becomes highly intermittent. Such behavior can be tracked via the
single-point joint pdf P(if>,£,t). Mapping closures are most naturally ap-
plied to this problem if P(V>,£,2) is constructed by sampling along particle
trajectories.24 If this is done, V/T>t in (36)-(38) is read as d/dt. But, since
the flow is compressible, an initially uniformly spaced set of trajectories will
tend to migrate into the shocks, so that the Lagrangian pdf PL(^^^) so
computed will differ from the Eulerian pdf PE(^^^} measured in a fixed
coordinate system. The relation between the two pdf's is

PE(1>,t,t) = NELPL(4>,t,t)/(p(i(>,t,t)}c:M (39)

where NEL normalizes PE(^, f , 2 ) to unity and [p(Vs£>^)]c:t/>f is ^ ne mean


conditional on given values of both t/> and f.
It is natural to treat Burgers' equation by the combined transformation
(8), (10), (12). Then (16) is

(40)

It is now necessary to evaluate the quantities [Vw]c:t/>£ and [£par]c:^- This


proceeds in the same fashion as illustrated above for [i/>Xx]c:if>i but there is
INTERMITTENT TURBULENCE 29

some additional complication because the statistical dependence of higher


space derivatives of the reference field on both I^Q and £o must be included.
The analysis is greatly simplified by the approximation that PE(^) is
Gaussian and that t/> and £ are statistically independent. The first part
of the assumption is supported by direct computer simulation of Burgers'
equation. The second part is more suspect and cannot be wholly consis-
tent within the closure method. The error in these approximations can be
estimated a posteriori, but this analysis will not be discussed here. Thus it
will be assumed that
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(41)

(42)

Here r(t) measures the decay of velocity amplitudes under viscosity and
the second part of (42) follows from (12). The energy-balance equation

(43)

yields
dr/dt = -rr,(e)/(ti) (44)
Now only [^Jc^, the conditional mean at given f is needed. To obtain
closure, the requirement is made that, over each dt, the stretching factor J
change so that the change in PE (*!>>£) given by (40) and (41)-(44) matches
the change under the equations of motion. This may be assured as follows:
The reduced Liouville equation for QE(£) implied by (37) is

(45)
where the divergence term on the right side expresses the effects of trans-
formation from Lagrangian to Eulerian coordinates.
On the other hand, (40)-(42) give

This implies the reduced Liouville equation

dt d£ \ dt J

where
30 R. H. KRAICHNAN

The right side of (47) expresses the effects of squeezing and stretching by
the J transformation. Subtraction of (45) from (47) and integration over £
now yields

OY_
dt = -e+i?K«]C:f + TT^T f M*') - n^m' (49)
VMs) J-oo

where the constant of integration is chosen to give finite results at ±00. In


this connection, note that homogeneity and the definition of N imply
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

CO fOO

tQB(t)dt = I «(OQiKOde = 0 (5°)


/ •00 J— 00

Eqs. (42)-(44) and (49), lead to the following somewhat complicated


equation of motion for J:

and evaluation of [£r:r]c:f by the techniques that gave (23) or (24) gives

with
Ci = ((dto/dz)2), k\ = C2/{£02)
The first term on the right side of (51) and the first term on the right side
of (52) will be recognized as like the terms on right side of the simplified
model (3). The J 2 term in (51) comes directly from the — £ 2 term in
(37). The derivative terms on the right side of (52) come from consistent
treatment of £xx under the space distortion represented by J. The integral
term in (51) arises from the effect on PE(I/>,£) of the N/J factor in (16).
This term makes (51) an integro-differential equation that must be solved
iteratively. The derivative and integral terms [omitted in the heuristic
model (3)] play an essential role in shaping the pdf near its maximum. The
behavior of J at large £ is dominated by the terms like those in (3), so
that the nearly exponential tail to the pdf implied by (3) is also a property
of QE(£) under (51) and (52). [In contrast to s in (3), £ is asymmetrical,
and the nearly exponential tail now appears only for negative £.] The tail
appears to be a robust property and not a consequence of the independence
approximations made to obtain (51).
Figures 7-12 compare the marginal pdf QE(£) obtained from solution of
(40)-(42), (44), (48), (51) and (52) with the results of a series of simula-
INTERMITTENT TURBULENCE 31

tions of Burgers' equation.24 The simulations started from a multivariate-


Gaussian distribution of i/>(z,0) and a spectrum of the form (29). Each
simulation involved 105 points unit-spaced on a cyclic line segment. Sets
of simulation runs were carried out for two parameter assignments: (Case
A) n = 0, 17 = 5, C0 = 1, Ci = 0.005, C2 = 7.5 x 10~5; (Case B) n = 2,
77 = 2.5, C0 = 1, Ci = 0.015, C2 = 3.75 x 10~4. The same parameter
values were used in integrations of the closure equations. Case A corre-
sponds to a spectrum of form exp(—k 2 /k%) and case B to a spectrum of
form k2 exp(—& 2 /fc 2 ), where k0 is a characteristic wave number.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Figures 7 and 8 show the evolution of velocity-gradient skewness S —


(£ 3 )/{£ 2 ) 3 ^ 2 - S is a measure of the production of (£ 2 ) by nonlinear effects.
The closure results capture two of the qualitative differences between Cases
A and B: the initial rapid rise in S is concave downward in Case A and
concave upward in Case B; and, after the initial rapid rise, 5 continues to
increase in Case A but decreases in Case B. However, the closure misses
the sharp overshoot exhibited by the Case B simulations. Figs. 9 and 10
show simulation and closure results for (£ 2 ) and (£ 2 )/{£ 2 ) in Case A. The
simulation curves are the means of all runs. Again, the closure captures
qualitative behavior.
Figures 11 and 12 show the full pdf's of £ for Cases A and B at several
times of evolution. There is increasing departure from Gaussian shape

FlG. 7 Velocity-gradient skewness 5 vs time for Case A, as given by 12 simu-


lation runs (data points) and by closure (solid line).
32 R. H. KRAICHNAN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

FlG. 8 Velocity-gradient skewness 5 vs time for Case B, as given by 10 simula-


tion runs (data points) and by closure (solid line).

with time in both Case A and Case B. Case B shows stronger departure at
given values of S and, in particular, displays a marked upward concavity
of the skirt reminiscent of Figs. 1 and 2. The near-exponential tail at large
negative £ predicted by the closure seems well-supported by the simulations.
There are no adjustable parameters or functions in the closure.
Under Burgers' equation, the fluid density becomes highly concentrated
in the shock regions. The zero-velocity points in the shocks are accumula-
tion points for fluid particles. Unlike the velocity field, the fluid density is
not diffused by viscosity. Consequently the density pdf becomes extremely
intermittent.
Eq. (38) gives the evolution of fluid density along the particle paths. The
latter, considered as trajectories in f, are well approximated at early times
of evolution by the trajectories induced under the J mapping. Later, when
viscosity is more important, the two kinds of trajectory can be significantly
different. This makes some complication in the analytical treatment of the
density pdf. Manipulations like those that led to (49) yield the following
equation for the effective evolution of p along a trajectory with fixed £o*

(53)
INTERMITTENT TURBULENCE 33
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

FlG. 9 Velocity-gradient variance vs time for Case A, as given by mean of 12


simulation runs (dotted line) and by closure (solid line).

Here 7 = Inp' and PE(J) is the Eulerian pdf of 7. The integral term
in (53), like those in (49) and (51), corrects the Lagrangian equation of
motion for differences between particle trajectory and mapping trajectory.
The integral is the same quantity in all three cases.
Figure 13 shows a log-log plot of PE(P) — PE(J)/P for simulation and
closure for Case A at t = 10. The most striking feature is the power-law
tail at large p, which indicates that moments of PE(P) do not exist above
some critical order. This behavior can be understood in a simple way. The
qualitative behavior of 7 at large negative £ is controlled by the — £ term
in (53). Since £ rapidly reaches quasi-equilibrium at large negative £, the
qualitative behavior of 7 in the tail is then 7 w \£\t. Thus the pdf of 7 has
an exponential-like tail resembling that of £, and the power-law tail for p
follows immediately.

4. Discussion

The mapping closures illustrated by the examples in this paper can be


varied in many ways and appear to be applicable to a rich variety of prob-
lems. Systematic sequences of improved approximations can be envisaged,
in which joint pdf's with more arguments are included in the analysis [e.g.,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

FlG. 10 The quantity (£*}/{£2) (a measure of dissipation of velocity-gradient


variance) vs time for Case A, as given by mean of 12 simulation runs (light dotted
line) and by closure (solid line). The heavy dotted line shows what the evolution
would be if only viscosity acted.

NORMALIZED AMPLITUDE

FIG. lla Normalized velocity-gradient pdf (?)l/2QE(£) vs £/(?)1/2 for Case


A, as given by closure (solid line) and 12 simulation runs (data points) at t = 10.
Also shown is a normalized Gaussian pdf for reference (dotted line).
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

NORMALIZED PDF NORMALIZED PDF

m
z
H
H
D
DO
C
m
o
m
o

CO
en
36 R. H. KRAICHNAN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

NORMALIZED AMPLITUDE

FIG. 12a Normalized velocity-gradient pdf {£2}1/2Q£($) vs £/(£ 2 } 1/2 for Case
B, as given by closure (solid line) and 10 simulation runs (data points) at t = 10.
Also shown is a normalized Gaussian pdf for reference (dotted line).

NORMALIZED AMPLITUDE

FlG. 12b Same as Fig. 12a, but for t = 20.


INTERMITTENT TURBULENCE 37
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

FlG. 13 Pdf PE(P) of fluid density p for Case A, as given by closure (solid line)
and 12 simulation runs (data points) at t = 10.

x^xx)]- This requires that the additional arguments appear also


in the transformation functions. The closures also lend themselves well to
empirically motivated modifications and parameterizations. The mappings
illustrated above all represent nonstochastic transformations of a stochastic
field. More generally, the mappings can themselves be stochastic, provided
that they are simple enough to yield explicit expressions for the needed
conditional means.
Single-point mapping closures yield moments of spectra, in the form of
variances of field amplitudes and their space derivatives, but they do not
predict full spectra. These closures therefore are complementary to the
more familiar spectral closures, rather than a replacement for them.
The exponential distribution for velocity gradient found in the appli-
cation to Burgers' equation raises hope that mapping closures have the
power to predict well the intermittency observed in the small scales of NS
turbulence. The shock-formation mechanism for velocity-gradient growth
in Burgers dynamics is substantially different from the vortex stretching
that arises from the NS equation. A fluid element once caught in a forming
shock in the Burgers case is there forever, while an element of vorticity
in the NS case may be subjected to a series of statistically uncorrelated
stretchings before it is diffused by viscosity. However, the exponential tail
to the pdf of velocity gradient plausibly is present also in NS .dynamics.
38 R. H. KRAICHNAN

Only the final stretching, the one that narrows a vortex to the point where
viscosity acts strongly on it, is likely to be important in determining the
asymptotic skirts of the pdf. If this argument is valid, fractal cascade pro-
cesses do not play the crucial role in fixing the pdf of velocity gradients: the
essential ingredients for intermittency already are present at low Reynolds
numbers. A qualitative argument to this end was presented some years
ago.29
The one-point pdf of velocity gradient does not capture the appearance
of structures with spatial extent. Certainly the sharply-defined rope-like
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

dissipative structures of high-Reynolds-number turbulence are not present


at low Reynolds numbers. If the preceding argument is correct, the presence
or absence of such clearly-defined structures does not have a strong effect
on the one-point pdf of velocity gradient and therefore does not have a
strong effect on the total dissipation, which is a second-order moment of
velocity gradients.
Mapping closures for N-S dynamics require that the J mapping be con-
sistently formulated for space dimensionality > 1. It is also necessary to
evaluate conditional means of spatial derivatives of pressure. This can be
done for compressible flow with an arbitrarily nonlinear equation of state
by following the evolution of the fluid density and its spatial derivatives via
the continuity equation and its derivatives [cf. (38)]. Conditional means of
density derivatives can thereby be found in terms of conditional means of
velocity derivatives. Conditional means of pressure derivatives then follow
from the equation of state. The temperature or entropy field must also
be followed. Incompressible flow may be treated as a limit of compressible
flow, or handled directly.

A cknowledgment s
H. Chen, S. Chen, T. Gotoh, C. E. Leith and S. B. Pope have made
essential contributions to the development of mapping closures. I am in-
debted to a number of colleagues for valuable discussions of intermittency
and, in particular, to J. R. Herring, M. Meneguzzi, Z-S. She, H. Shen,
K. R. Sreenivasan and V. Yakhot. A. Vincent and M. Meneguzzi have
kindly given permission for the reproduction of Fig. 1. This work was
supported by the National Science Foundation, Division of Atmospheric
Sciences, under Grant ATM-8807861 and the Department of Energy under
Contract W-7405-Eng-36 with the University of California, Los Alamos Na-
tional Laboratory. During part of this work, the author served as consultant
to the Theoretical Division of Los Alamos National Laboratory.

References
*G. K. Batchelor and A. A. Townsend, Proceedings of the Royal Society (London)
A199, 238 (1949).
INTERMITTENT TURBULENCE 39

2
C. H. Gibson, G. R. Stegen and R. B. Williams, Journal of Fluid Mechanics 41,
153 (1970).
3
A. Y. Kuo and S. Corrsin, Journal of Fluid Mechanics 56, 447 (1972).
4
C. W. Van Atta and T. T. Yeh, Journal of Fluid Mechanics 71, 417 (1975).
5
F. Anselmet, Y. Gagne, E. J. Hopfinger and R. A. Antonia, Journal of Fluid
Mechanics 140, 63 (1984).
5
E. D. Siggia, Journal of Fluid Mechanics 107, 375 (1981).
7
R. M. Kerr, Journal of Fluid Mechanics 153, 31 (1985).
8
R. M. Kerr, Physical Review Letters 59, 783 (1987).
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

9
K. Yamamoto and I. Hosokawa, Journal of the Physical Society of Japan 57,
1532 (1988).
10
P. Tong and W. I. Goldburg, Physics of Fluids 31, 2841 (1988).
11
1. Hosokawa and K. Yamamoto, Journal of the Physical Society of Japan 58,
20 (1989).
12
I. Hosokawa, Physics of Fluids A, 1, 186 (1989).
13
A. Vincent and M. Meneguzzi, Journal of Fluid Mechanics 255, 1 (1991).
14
S. Kida and Y. Murakami, Fluid Dyn. Res. 4, 347 (1989).
15
Y. Gagne and B. Castaing, Compte Rendu Academie Sciences (Paris) 312, 414
(1991).
16
J. C. McWilliams, Journal of Fluid Mechanics 198, 199 (1989).
17
J. R. Herring and O. Metais, Journal of Fluid Mechanics 202, 97 (1989).
18
O. Metais and J. R. Herring, Journal of Fluid Mechanics 202, 117 (1989).
19
P. Saffman, Physics of Fluids 13, 2193 (1970).
20
R. H. Kraichnan, Journal of Fluid Mechanics 62, 305 (1974).
21
H. Chen, J. R. Herring, R. M. Kerr and R. H. Kraichnan, Physics of Fluids A,
1, 1844 (1989).
22
R. H. Kraichnan in Current Trends in Turbulence Research, edited by H. Bra-
nover, M. Mond and Y. Unger, (American Inst. of Aeronautics, Washington,
1988), p. 198.
23
H. Chen, S. Chen and R. H. Kraichnan, Physical Review Letters 63, 2657 (1989).
24
T. Gotoh and R. H. Kraichnan, submitted to Physics of Fluids A.
26
S. B. Pope, private communication
26
V. Eswaran and S. B. Pope, Physics of Fluids 31, 506 (1988).
27
S. B. Pope, Progress in Energy Combustion Science 11, 119 (1985).
28
J. M. Burgers, in Statistical Models and Turbulence, edited by M. Rosenblatt
and C. Van Atta (Springer-Verlag, New York, 1972), p. 41.
29
R. H. Kraichnan, Physics of Fluids 10, 2081 (1967).
Plane Mixing Layer Between Parallel Streams
of Different Velocities and Different Densities
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

H. E. Fiedler,* M. Lummer,| and K. Nottmeyert


Technische Universitat Berlin, Berlin, Germany

Abstract
The influence of a density difference between the two streams of a mi-
xing layer was investigated experimentally and numerically for the cases
of parallel and inverse gradients of velocity and density. In the course of
the numerical simulation a stability analysis of the inhomogeneous flow was
done. For the spread of the inhomogeneous layer an empirical spread rela-
tion is given. Both, experimental and numerical results show reduced flow
stability in the case where higher density is on the low velocity side and vice
versa.

Introduction
The problem investigated is frequently encountered in technical applicati-
ons, yet only little attention has been devoted to its scientific investigation in
the past. The inhomogeneous free jet was investigated, e.g., by Thring and
Newby1. However, because of strong entrainment, the jet's volume flux is
rapidly increased downstream and as a consequence density effects become
insignificant already after a few diameters. To study the basic and non-
decaying effect of density inhomogeneity in a turbulent flow over sufficient
length of the flow the parallel mixing layer is best suited. Investigations of
this flow are, however, scarce 2>3 and, so far, to some extent ambiguous (see,
e.g., Ref.4).
In the following we describe an investigation of the effect of a density dif-
ference on the two-stream mixing layer between streams of different velocity
and different density. We use notations as sketched in Fig. 1.
The influence of density inhomogeneities is twofold: 1) via buoyancy ef-
fects and 2)via inertia effects. The investigations described here were aimed

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Professor, Hermann-FSttinger-Institut.
fScientist, Hermann-Fottinger-Institut
40
DENSITY MIXING LAYER 41

_.___— -——-—--""
^-1-
1
I
1 "•"•-^-~— ^__^
^~~~~~^~- -___
b
—1 1J 1 ^~~-^__^
1
1 ~~~~~

PI
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

r =

P =
P2

Fig. 1 Schematic of flow situation and parameters.

at studying only the latter, whereas the former, by appropriate choice


of parameters, was sufficiently suppressed. Some global effects, e.g., the
spread behavior, are known for a few parameter combinations from the
work of Refs. 2 and 3, the former used as a basis for comparison.
A first attempt at describing the integral spread behavior of inhomoge-
neous mixing layers by a density weighted transformation, in analogy to
the uAbramovich-Sabin" relation (see, e.g., Ref. 5) was suggested by Ya-
kovlevsky 6 as b = &oo(l + s)(l —r )/2(l + sr), where r and s denote the
velocity and density ratios, respectively, and 6 gives the spreading rate (600
for r — 0 and s = 1). This relation fails for Xp = — 1, where it gives infinite
spread. An improved version was earlier suggested by the first author as
6/600 = A u (l — 0.6AP\/A^"). Detailed knowledge about turbulence structure
was, however, still missing (see Fig. 12).
The investigation was done experimentally and, in parallel, by direct nu-
merical, two-dimensional simulation of the turbulent flow, some details of
which will be described in the following.

Experimental Arrangement and Results

The major body of experiments was done in a specially designed blow-


down facility producing a two-(seperate) stream mixing layer with 60 x
60 mm cross section and 680 mm length, allowing for variable velocity and
density ratios. For all measurements taken, the separating boundary layer
at the trailing edge was laminar at the low velocity side and turbulent at
the high velocity side. Fully computerized experimentation control allowed
data to be taken in the available running time of around 40 s.(See Fig. 2).
42 H. E. FIEDLERETAL

60mm
60mm

gas reservoir
settling chamber
fans i nozzle test section exit
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Nozzle exit: 0.3/0.2 % turbulence level


1.2/2.0 % profile non-uniformity

Fig. 2 Experimental facility.

Measurements were done for flows between CO2 and air and for helium and
air. Characteristic values are given in Table 1.
Simultaneous density and velocity data were obtained with combinations
of hot-wires with hot-wire-aspiration probes, as well as with combination
hot-wire probes consisting of wires with different diameters that are opera-
ted at different overheat ratios. Fig. 3 shows two hot-wire probe configurati-
ons as used in the course of this work. We found the temporal as well as the
spacial resolution of the combined aspiration-hot-wire-probe to be unsatis-
factory for the task at hand. All subsequent measurements were, therefore,
done with probes of type A or B, whereas measurements with the aspiration

Table 1 Characteristic values of experimentation

C02 He

U2 (m/s) 12.5- 15.0 7.2- 8.8


3.1- 3.8 105 3.1- 5.1 104
RimaX 1.2- 2.3 10~5 4.5-8.3 10-2

Microscales, homogeneous case: A u>y = 2.8 mm,A U)< j = 3.9 mm


DENSITY MIXING LAYER 43
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 3 Hot-wire probe configurations: A: u — Q — probe, wire 1: 9 //m


diameter, 1.6 overheat ratio; wire 2: 5 /im diameter, 1.2 overheat
ratio. B: u — v — £-probe, wire 1: 9 ^m diameter, 1.6 overheat
ratio; wires 2 and 3: 5 //m diameter, 1.2 overheat ratio.

probe only served for corroboration of tendencies. Moreover, both probes A


and B were unsatisfactory for measurements in helium-air mixtures, owing
to the thermal characteristics of helium. Data obtained with helium-air
are, therefore, not presented here. Their tendency, however, supports the
findings from the experiments with CC>2 and air.
Parallel to probe measurements, the flow was visualized using a schlieren
apparatus in combination with conventional video technique, followd by
digitization and semiquantitative evaluation.
Most data were obtained at a fixed downstream position of xm = 480 mm,
and a velocity ratio of r = 0.4 (A u = .43), with 2.2 104 < Re$ = 6Aui,2/v <
2.4 104. These values are slightly in excess of Konrad's 7 "diffusion-critical"
value of 2 104.
Three cases were considered and juxtaposed: 1) countergradient = higher
density on the low speed side (Xp)min = —0.22); 2) homogeneous = no
("active") density gradient, and; 3) cogradient = higher density on high
speed side (A p)maa; = 0.22).
Fig. 4 shows the corresponding mean velocity and density distributions.
The p'lt'-correlation is a particularly sensitve indicator of the density influ-
ence, suggesting stronger organization and stability of coherent structures in
the cogradient case (positive A^), where values of 0.8 — 0.95 over most of the
turbulent flow regime are found. The distributions are shown, together with
the p'v1-values, in Fig. 5. This observation is corroborated by the different
spread behavior, as seen in Fig. 4 and, more specifically, seen later in Fig.
12. Other results of significance are the standard deviations (rms-values) of
the velocity fluctuations u' and v' and the density fluctuation // as shown
in Fig. 6. We observe extraordinary high p' values on the low velocity side
in the countergradient case, suggesting strong mixing. Reduced mixing is,
on the other hand, obvious from the unusually low v' values in cogradient
case. Fig. 7 presents intermittency distributions 7 and the probabilities wl
and w2 for the presence of gas 1 (high velocity side) and gas 2 (low velo-
city side), where 7 = 1 — (wl -f w2). We find the major difference on the
44 H. E. FIEDLERETAL

1.0 jAp=-0.22 ;X P =0.22

-Js
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0.1 -W-'
Q.

O-5
...X....I

-0.1 0.1 -0,1 0.1

•Y]=. y-y0
x-x n x-x n
Fig. 4 Mean velocity and density profiles for air-CO2 combination, A u
0.43. Data presented were measured with different probes.

1.0 1.0

id
^r*
-1.0
-0.1 0.1 -0.1 0.1

Fig. 5 Velocity-density correlation coefficients, A u = 0.43.


DENSITY MIXING LAYER 45

0.3 U.J i ! ! =
;
V-0.22! ! i iPrms ! Xp=0.22
A '• A •

1 1 i i I J ; iPrms ;
A i :A j
0.2 ...... ..I......]... ...1. ...J........ !»....,.... ...j.. ......... ......|....... 0.2 : A
,-HAp ...i .......;.......
• I A

......r!......T.......
: i AJ
r.....A,. .
I : - ' : *

!
i
i
;
•!
I A « i* k ;U
rms
0.1
! ^t""^ i
0.1
....... I w xf.Au --!•--
1
1 i *
i jj»
i
.......:.......,...................
1
i
i1 Vrrriis!
!71?j« *^.u A
; : ; JX
.Vr_. ...............
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

I. i
: !
I^U 1 !
I Au : j* :i??ff* Au j tt..i !

-0.1 0.1 -0.1 0 0.1


1)

Fig. 6 Effective (RMS-) values of turbulent velocity components and den-


sity fluctuations.

low velocity side of the flow, while in both cases there is negligible overlap
from both sides. Fig. 8 shows distributions of unmixedness for the cases
considered. The unmixedness was obtained as

where t\ = t(c > c),<2 = t(c < c) and c — concentration. Again the
distribution is asymmetric in the countergradient case, with a maximum on
the low velocity side. On the average, however, the unmixedness is in both
cases around a value of 0.5. The mean density inside the structures may be

0.1

Fig. 7 Intermittency distributions, A u = 0.43.


46 H. E. FIEDLERETAL
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-0.1 0 0.1 -0.1 0 0.1


Fig. 8 Distributions of unmixedness DM.

estimated as
- _ +
PCS

where VEI andJ/£2> the entrainment rates are estimated from VE « u y(j =
0,5). We find p^s/ I Apl,2 |= 0.47 and 0.60 for the countergradient case
and the cogradient case respectively. These values are in good agreement
with the location of platforms in some of the density profiles.
To obtain better information about the characteristic structures and pro-
vide a basis for understanding and interpretation of the statistical distribu-
tions, the flow was periodically excited at a very low level to stabilize the
structures. Characteristic distributions and contours, thus obtained, serve
as a basis for explaining the differences in the /?V-correlation distributions
for the different configurations as a consequence of structural asymmetry in
the countergradient case.

Results from Numerical Simulation

Numerical investigation of this flow was done by linear inviscid stability


analysis as well as by direct two-dimensional numerical simulation, both for
temporal (tdf) and spacial (sdf) development. The simulation was done on
the basis of the following:
1) vorticity-sreamfunction formulation of the balance equations of mass
and momentum (incompressible flow: Atx = 0);
2) full-time explicit code with Fourier, Chebeychev, or spline approxima-
tion of the flowfield;
3) spectral or multigrid solution of the Poisson equation for streamfunc-
tion and pressure;
4) time integration with forth-order Runge-Kutta scheme, and;
5) rectangular grid with 128 x 65 points (tdf) or 129 x 33 points (sdf),
respectively.
DENSITY MIXING LAYER 47

O -0.7575
A -0.5681
+ -0.3788
-OC;
x -0.1894
o 0
+ 0.1894
0.3788
0.5681
0.7575
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 9 Characteristics of flow stability. Top: maximum amplification of


spatial perturbation versus 1/AU for various X e . Bottom: spatial
amplification -a{ versus frequency /? for A u = 1 and various A c .

The computations were done on a Cray XMP machine of the Konrad-


Zuse-Zentrum fur Informationstechnik Berlin. Major results obtained are
as follows.

From Linear, Inviscid Stability Analysis (Parallel Flow


Assumption)
1) Temporal development: No influence of the magnitude oi Xu or ot the
sign of A, on the instability behavior was found;
2) Spacial development: The influence of the density ratio increases with
decreasing convection velocity, where the instability of the flow is stronger
for the case of smaller density on the high velocity side (-countergradient
case) (See Fig. 9).
48 H. E. FIEDLERETAL

From Direct Simulation


1) Temporal development: Strong dependence of dynamical flow deve-
lopment from initial conditions renders transformation from temporal to
spacial development difficult.
2) Spacial simulation (129 x 33 grid, Re « 70): In this case slight ex-
citation of the flow at the trailing edge was needed to trigger dynamical
development. Strong influence of A u on spread is observed/Clear influence
of A p on structural shape and development is demonstrated in Figs. 10, 11.
Comparison of the two simulation schemes demonstrates a considera-
ble equivalence concerning the major structural characteristics of the flow.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Thus, the considerably less extravagant temporal simulation appears to be


sufficient for many cases. Difficulties arise only in assessing global aspects,
e.g., the overall spread of the flow.

General Discussion

The numerical simulation, although providing interesting and rather rea-


listic flow details, does not yield to useful results for the spread behavior

21.3

0.0 = -0.75

-21.3

Fig. 10 Vorticity contours (isoverts), spatial simulation, \u = 0.45.


DENSITY MIXING LAYER 49

-0.46

0.00
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0.46

-0.46

X p =-0.75

0.00

0.46

-0.46

; P 2 (>Pi)

0.00

; Pi
0.46

-0.46

0.00

0.46

Fig. 11 Temporal simulation: a: isoverts, and b: contour lines of barocli-


nic torque.
50 H. E. FIEDLERETAL

0.30
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0.00

Fig. 12 Spread rate of inhomogeneous mixing layer (Xu = 1; measure-


ments and suggested relationship.
A Helium, measurements;
+ CO2, measurements;
• Helium, optical study (schlieren);
D data from Ref. 2;
o data from Ref. 10;
A data from Ref. 6;
O data from Ref. 8.

(present),
(Ref. 6),
(present),
= 0.145; mQ = -0.091.

even in the case of spacial simulation. This well be attributed to the two-
dimensionality of the simulation scheme.
An interesting and characteristic detail is the persistent //t/'-correlation
maximum on the high-density side, where we have maximum concentration
and steepest gradients of vorticity. On the low-density side, where the
correlation changes sign in the countergradient case, regions of vorticity
of inverted sign show in the numerical simulation. The p'v1-correlation is
consistently of opposite sign with the mean density gradient.
The obvious question as to the quantitative influence of the A^-parameter
on the turbulent structure at a given Xu finds no conclusive answer on the
DENSITY MIXING LAYER 51

basis of this investigation. From a comparison with the data obtained in


helium/air (A^ = ±0.75) it seems that, not only the general tendency, but
also the absolute values of, say, the it'p'-correlation are equal to those in the
CO2/air (Xp — ±0.22) case. The same holds, with a grain of salt, for the
other statistical values presented.
A general result, concerning the dependence of the mean spread of the
density profiles as a function of the density parameter A p as obtained from
the experiments, is shown in Fig. 12 which also contains data from other
sources. All data plotted in this diagram have been reduced to the case
\u — 1, using the Abramovich-Sabin relation. Extended relations to fit the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

data in this diagram, accounting for A u , as well as for A^, are suggested and
included.
From Fig. 9 it is clear that the cogradient case provides a more stable
situation than the countergradient case. Consequently the coherent struc-
tures have a reduced growth rate in this configuration, undergoing fewer
amalgamations, which, as opposed to the countergradient case, leads to re-
duced spread. In any case, as indeed was already observed by Abramovich8,
the influence of density differences on the spread is found to be more pro-
nounced on the low-density side of the flow. (See Fig. 11).

A cknowledgment
Support of this investigation by Deutsche Forschungsgemeinschaft (DFG)
is gratefully acknowledged.

References
M. W., Newby, M. P., 4th Symposium on Combustion, Cam-
bridge, MA, 1952, p. 789.
2
Brown, G. L., and Roshko, A., "On Density Effects and Large Structure
in Turbulent Mixing Layers," Journal of Fluid Mechanics, Vol. 64, 1974, p.
775-816.
3
Rebollo, M. R., "Analytical and Experimental Investigation of a Turbu-
lent Mixing Layer of Different Gases in a Pressure Gradient," Ph.D. Thesis,
California Institute of Technology, 1972.
4
Birch, S. F., and Eggers, J. M., "A Critical Review of the Experimental
Data for Developed Free Turbulent Shear Layers," NASA SP-321, 1973, p.
11-40.
5
Dziomba, B., and Fiedler H. E., "Effect of Initial Conditons on Two-
Dimensional Free Shear Layers," Journal of Fluid Mechanics, Vol. 152,
1985, p. 419-442.
6
Yakovlesky, O. V., "The Problem of the Thickness of the Turbulent Mi-
xing Zone on the Boundary Between Two Gas Streams of Different Velocity
and Density," Izv. Akad. Nauk SSSR, Otd. Tekhn. Nauk, Vol. 10, 1958.
7
Konrad, J. H., "An Experimental Investigation of Mixing in Two-Di-
mensional Turbulent Shear Flows with Application to Diffusion-Limited
Chemical Reactions," Ph. D. Thesis, California Institute of Technology,
1976.
52 H. E. FIEDLER ET AL

8
Abramovich, G. N., "The Theory of Turbulent Jets", M.I.T. Press, Cam-
bridge, MA, 1963.
9
Abramovich, G. N., Yakolevsky, O. V., Smirnova, I. D., Sekundov, A.
N., and Krashenikov, S. Y., "An Investigation of Turbulent Jets of Different
Gases in a General Stream," Astronautika Ada, Vol. 14, 1969, p. 775-816.
10
Bogdanoif, D. W., "Interferometric Measurement of Heterogeneous
Shear-Layer Spreading Rates," AJAA Journal, Vol. 22, 1984, p. 1550-1555.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227
Instabilities in the Axisymmetric Jet:
Subharmonic Resonance
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

C. O. Paschereit* and I. J. Wygnanski t


University of Arizona, Tucson, Arizona 85721

Abstract
A resonant subharmonic interaction between two axisymmetric traveling-
waves was induced in a shear layer of an axisymmetric jet by superposing
on it controlled sinusoidal perturbations of two frequencies separated by one
octave. Wherever the excited waves are nondispersive, (J. Cohen and I.
Wygnanski, Journal of Fluid Mechanics, Vol. 176, 1987, pp. 221-235), they
interact in a manner that enhances the amplification rate of the subharmonic
wave train. The distribution of the phase-locked velocity fluctuations was
measured across the flow at various streamwise locations at one frequency
pair, and the data were used to assess the evolution of the two waves in the
direction of streaming. The rate of amplification of the subharmonic wave
is increased by the nonlinear interaction with the fundamental wave in the
range of 0.4 < x/D < 0.7. Beyond this distance the shear layer becomes
sufficiently wide for dispersion, and the energy transfer to the subharmo-
nic is terminated. Depending on the overall forcing level, the initial phase
difference between the two excited waves may either suppress or have no
influence on the subharmonic resonance. The influence of the initial am-
plitude ratio between the subharmonic wave and the fundamental wave is
discussed.
Introduction
The initial evolution of an axisymmetric jet has been investigated ever
since jet propulsion became technologically feasible. Nevertheless, the me-
chanisms affecting mixing, combustion, and the generation of noise are not
completely understood, nor can they be fully predicted or controlled.
The spreading rate of the jet can be influenced by external excitation,
which also enhances the mixing on the molecular scale and therefore the
rate of chemical reaction.1 The range of distances at which mixing is en-
hanced depends on the frequency and the amplitude of the excitation as
well as on some characteristic parameters of the unexcited motion. The

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Also at Hermann-Fottinger-Institut, Technische Universitat, Berlin.
t Also Faculty of Engineering, Tel-Aviv University, Tel-Aviv, Israel.
53
54 C. O. PASCHEREIT AND I. J. WYGNANSKI

degree of control is therefore limited as long as only one frequency has been
excited. Crow and Champagne2 were the first to observe that the instability
wave increases the rate of spread of the jet only if it grows in the direction of
streaming. Increasing the forcing amplitude beyond a certain saturation le-
vel, while maintaining all other flow parameters constant, does not enhance
the growth of the fundamental instability wave beyond its saturation level
and therefore has no additional effect on the mixing process (see also Oster
and Wygnanski3 , Weisbrot and Wygnanski4 and Fiedler and Mensing5).
Further enhancement of the mixing can be achieved by a concomitant
excitation of a second instability wave, e.g., the subharmonic. When the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

fundamental wave becomes neutrally stable, its subharmonic, which then


amplifies at approximately its maximum rate, enhances the mixing. The
subharmonic wave may resonate with the fundamental and give rise to a
secondary instability mechanism leading to a strong augmentation of the
subharmonic wave6.
Cohen and Wygnanski1 derived the conditions under which two distur-
bances can resonate in an axisymmetric shear layer. Not surprisingly, they
found that, for a fundamental wave to interact with its subharmonic, both
waves must propagate with identical phase speeds, allowing sufficient time
for the fundamental high-amplitude wave to transfer energy to the subhar-
monic wave.
The present study focuses on the details of the interaction of fundamen-
tal and subharmonic wave by forcing both waves at the same time. The
influence of the initial phase difference between the two waves, the rela-
tive forcing level of the subharmonic with respect to the fundamental, and
the total forcing amplitude of both waves are currently being investigated.
Since there is more than one way in which a resonance may be generated,
one has to be careful in mapping the correct physical mechanism involved.
Third-order effects may become important and perhaps even dominate the
flow, if one of the wave trains exceeds a certain threshold level. For this
reason an overall forcing level of the excitation of less than 3.2% (maximum
amplitude at ^ = 0.25) was maintained throughout the investigation.

Experimental Apparatus

A schematic of the airjet facility is shown in Fig. 1. The jet flow is supplied
by a turbo blower equipped with an air filter. The flow rate is controlled
by the frequency of the three-phase motor power supply. The air enters the
apparatus at the top of a large (1.25 m high, 1.0 m wide) rectangular plenum
chamber, from which it enters a smaller cylindrical settling chamber of 0.3 m
diam and approximately 0.8 m height, placed axisymmetrically inside the
rectangular one. To reduce the turbulence level of the flow, the cylindrical
plenum chamber is equipped with a honeycomb and three screens, and the
rectangular plenum chamber is fitted with perforated plates. The air passes
from the rectangular to the cylindrical chamber via a second set of high-
density automotive air filters.
AXISYMMETRIC JET INSTABILITY 55

Nozzle
_ _ p| Azimuthal Array of
U—M—U Speakers
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

3-Air Supply

Perforated
Plates
\—Acoustic Liner

Honeycomb

Automobile
Air Filter

Plenum Speaker

Fig. 1 Air jet facility.

The flow exits through a spun aluminium nozzle with a contraction ratio
of 36:1 and an exit diameter of 50.8 mm. The exit velocity used in the
present experiment is 8 m/s, and the corresponding Reynolds number, based
on the nozzle diameter, is 28,000. The exit flow has a "top hat" velocity
profile with only a thin laminary boundary layer at the nozzle wall. At 8
m/s, the freestream turbulence level is 0.097%. The entire jet is enclosed in
a large cage of 1.6-mm mesh screen to minimize any effect of room drafts.
The streamwise component of velocity was measured with a circumferen-
tial array of eight hot wires. The eight probes are mounted on separate
micrometer screws and could be adjusted independently in both the radial
and the axial direction. Axial alignment of the probes is achived with the
aid of an optical cathetometer. Radial alignment is accomplished by adju-
sting each of the eight hot wires to the position of one-half of the freestream
velocity at x/D = 0.2. The traversing mechanism is then capable of moving
all sensors simultaneously an equal distance in the radial direction. The
locally built hot-wire anemometers are operated in the constant tempera-
56 C. O. PASCHEREIT AND I. J. WYGNANSKI

ture mode at an overheat ratio of 1.8. The signals are amplified to achieve
the maximum dynamic range of the analog-to-digital converter, operating
between -5 V and +5 V. They are low pass filtered at 10 kHz. The coherent
motions were extracted by phase averaging. Controlled excitation of the
flow is accomplished by a loudspeaker at the base of the plenum chamber,
which generates axisymmetric disturbances at the exit plane of the nozzle.

Some Experimental Results


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

In the initial stages of the experiment the jet was forced at a single fre-
quency of 368 Hz, which approximately corresponds to the preferred fre-
quency of the jet at the exit velocity of 8 m/s. From a downstream distance
of x/D > 0.4 on, the phase velocities of the forced (fundamental) frequency
and its subharmonic are nearly identical; therefore, a narrow band of fre-
quencies centered around the subharmonic frequency was found to be ampli-
fied in response to the excitation (Fig. 2). Conversely, forcing at a frequency
of 184 Hz did not produce any significant response at the subharmonic fre-
quency of 92 Hz. In this case the two waves do not propagate with the same
phase speed; hence, there is insufficient time for an exchange of energy from
the fundamental wave to its subharmonic (Fig. 3). These results were veri-
fied for different excitation levels, jet velocities, and frequencies. In the next
step the jet was forced with both, the fundamental and the subharmonic
frequency. At different x/D locations, phase-locked data were taken with
the reference signal based on the subharmonic frequency. The maximum

10-*

10+01-:

Q
00 10+°°-,
CL-

IO-01-,

10" I \ I I I I I I I I I I I I I I I I I
0. 100. 200. 300. 400. 500. 600. 700. 800. 900. 1000.
Frequency in Hz
Fig. 2 Power spectrum, / = 368 Hz, x/D = 0.8.
AXISYMMETRIC JET INSTABILITY 57

10 +0

10 +0

o
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(/)
CL

10- i 1 i I I I I i I I I I I I T I I I I
0. 100. 200. 300. 400. 500. 600. 700. 800. 900. 1000.
Frequency in Hz

Fig. 3 Power spectrum, / = 184 Hz, x/D = 0.8.

amplitude of velocity fluctuations of the subharmonic, the fundamental and


the first harmonic were determined at each x location and are shown in
Fig. 4. For this case the initial amplitude ratio between subharmonic and
fundamental was chosen to be 1.6, and the initial phase difference was 0
deg.
Considering only the maximum amplitude at a given cross section of the
flow is not a very accurate method of assessing the downstream amplifiation,
but since the overall shape of the eigenfunction does not change dramati-
cally in the region of interest, this criterion can be used to display the
growth of the different modes. Verification of the validity of the mechanism
assumed by Cohen and Wygnanski for the enhancement of growth of the
subharmonic is achieved if the phase velocities are found to be equal. We
have indeed found that the phase difference between the fundamental and
the subharmonic is constant, i.e.,

2/7, - /?/ = const.


This is demonstrated in Fig. 5 for an input phase difference A3> = 0 deg.
The filled symbols in Fig. 5 represent this phase difference, and the open
symbols display the phase advance of the subharmonic. The results clearly
indicate that during the resonance interaction the phase difference between
the two waves remains constant.
One of the purposes of this study was to determine the influence of the
phase difference between the fundamental and the subharmonic on the gro-
wth of the subharmonic, with all other parameters kept constant. The
phase difference A3> is defined with respect to the fundamental. The ma-
58 C. O. PASCHEREIT AND I. J. WYGNANSKI

0.20 E subharmonic 184 Hz


+fundamental 368 Hz E
<S> harmonic 736 Hz a
0
0.15
0

0.10
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0.05

"o 0.2 0.4 0.6 0.8 1.0 1.2

X/D
Fig. 4 The variation of the maximum amplitude of the subharmonic, fundamental
and first harmonic wave with x/D, r = 1.6, A$ = 0°.

180 180

90 90
O)
0)
T5 O)
C 05

J °
'^3
0 .£

GO0

CO
-90 -90

-180
0.2 0.4 0.6 0.8 1.0 1.2

X/D
Fig. 5 The variation of the phase difference 2j3s — /?/ and the phasejidvance of the
subharmonic wave with x/D, r = 1.6, A$ = 0°, (uf)max/Uci = 1.6% at
x/D = 0.25.
AXISYMMETRIC JET INSTABILITY 59

a subharmonic 184 Hz
• fundamental 368 Hz
^> harmonic 736 Hz

^ . . , , . , , . | , . , , , , , ,0, ,,,, , , , , , , , , _
0.20
A<D=90° ~

0.15 13
_ a J
IB ^

A
V
0<1
°
«^
<*~*\
a •
" ^ w**i*" •
* * m
:
-i
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

f
0.05
r / ****v * ~ / ^^^ * -
:
0 , , , , i £<**^> • . iT^.*T, ,«, ,t, A ,5 . . . . I 'vjx^tf** , I ,^t, ,o, ,0, A ,5
() 0.2 0.4 0.6 0.8 1.0 12 C) 0.2 0.4 0.6 0.8 1.0 1.

X/D X/D
_• i i i j • i i • | i i i i | i i i i | i i i i | i i i i_ _i • •i | • •i i | i i i i | i i i i$ i •i i | i •i i
0.20
A0>=1800 o ~ A<D=270° H ~
OJ3 •
E H a i
,0.15 H EJ _
-
jx£?
% nH 3 a 0
Si
E 0.10 ^***^ H ^ **** '-
//* aa- * » * ::
3 ^Ti * :
V / J8 *
0.05 f B * 4 J
:
^ y^O^**^^

, , , .A^rfT,, i , .^r,,*, ,*,,«, ,3 :


» ^«% 3
. ...it,^* , ,, i . * . T . A . t . A . 4
n
0.2 0.4 0.6 0.8 1.0 1.2 0 0.2 0.4 0.6 0.8 1.0 1.2

X/D X/D

Fig. 6 The variation of the maximum amplitudes for four different phase angles
with x/D, r = 0.4, (uf)maz/Uci = 1.6% at x/D = 0.25.

ximum amplitude of the subharmonic wave was measured at various values


x/D and is shown in Fig. 6. The amplitude ratio between subharmonic and
fundamental is in this case 0.4. For the initial phase difference of 90 deg,
the maximum amplitude of the subharmonic was supressed, as opposed to
the cases with phase differences of either 0, 180, or 270 deg. To determine
whether this effect depends on the imposed amplitude ratio between the
subharmonic and the fundamental, the experiment was repeated at three
additional amplitude ratios of r = 0.1, 0.8, and 1.6 and at four different
initial phase angles. Once again the lowest maximum amplitude of the sub-
harmonic wave was observed at A$ = 90 deg, irrespective of the imposed
amplitude ratio. For example, for an initial amplitude ratio of r = 0.1 the
maximum amplitude ratio at x/D w 0.8 between A3> = 90 and 0 deg, was
measured as 65%. The previously given ratio changed by a mere 4% when
the imposed amplitude ratio of the subharmonic was increased by a factor of
16 (i.e., r = 1.6), whereas the overall excitation level was retained (Fig. 7).
An increase in the combined forcing level reduces the effect of the initial
phase difference on the maximum amplitude of the subharmonic attained.
The effect of doubling the amplitude of the combined wave for r — 0.1 is
shown in Fig. 8, where the downstream evolution of phase-locked amplitudes
of fundamental and subharmonic is plotted. In this case A$ = 90 deg had
no effect on the maximum amplitude attained. It should be noted that the
60 C. O. PASCHEREIT AND I. J. WYGNANSKI

phase-locked forcing level in the middle of the shear layer did not exceed
3.2% at x/D = 0.25. Previous works of Mankbadi8 and of Arbey and
Ffowcs Williams9 stress that the amplitude of the subharmonic depends
highly on the initial phase difference between these two waves. Their results
are based on measurements along the centerline only carried out at extremly
high forcing levels (i.e., (u/)/Uci > 7%). To compare the results, additional
measurements were also made along the centerline at higher forcing levels
at three different pairs of frequency: 1) 184 Hz and 368 Hz, 2) 62 Hz and
124 Hz, and 3) 31 Hz and 62 Hz. For all three cases the forcing level of
the subharmonic was 10% of the forcing level of the fundamental. The
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

attained phase-locked forcing level of the fundamental at x/D = 0.2 is


(uj)/Ud = 0.5% for case (1), 1.7% for case (2) and 2.7% for case (3).
Figure 9 shows the downstream evolution of the subharmonic for case 1) for
two different initial phase angles. The solid line represents the phase-locked
amplitude of the subharmonic for an initial phase difference of A$ = 0 deg,
and the dashed line represents that for an initial phase difference A$ = 90
deg. The maximum amplitude attained for A3> = 90 deg, is about 93% of
the maximum amplitude attained for A3> = 0 deg, however, the difference
is not significant. Forcing the jet at 62 and 124 Hz results in the same ratio
for the maximum amplitudes as that shown in Fig. 10. For 31 and 62 Hz no
a subharmonic 184 Hz
• fundamental 368 Hz
0 harmonic 736 Hz

0.20 -

1.2

Fig. 7 Comparison of the amplitude evolution for the initial forcing ratios r =
0.1 and r = 1.6 for two different phase angles A$ = 0° and A$ = 90°,
(uf}max/Uci = 1.6% at x/D = 0.25.
AXISYMMETRIC JET INSTABILITY 61

3.0
0.20
s subharmonic 184HzEs 2.5
+fundamental 368 Hzo
<S> harmonic 736 Hz H 2.0

1.5 i
j

0.05
0.5

0
0.2 0.4 0.6 0.8 1.0 1.2 0 0.2 0.4 0.6 0.8 1.0 1.2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

X/D x/D
180

90 . .•jo*

-180
0.2 0.4 0.6 0.8 1.0 1.20 0.2 0.4 0.6 0.8 1.0 1.2

X/D X/D
Fig. 8 The variation of the maximum amplitudes and the phase with x/D for a
higher forcing level, r = 0.1, (uf}max/Ucl = 3.2% at x/D = 0.25. The solid
line represents the momentum thickness.

amplification of the subharmonic could be observed, regardless of the initial


phase angle: The two waves do not travel with the same phase speed in the
region where mode 0 is amplified linearly. They have, therefore, insufficient
time to exchange energy. At large x/D, where the two waves become non-
dispersive, the amplification of mode 1 becomes dominant and the described
resonance mechanism can, therefore, not take place.
Forcing at high amplitudes indicates a small dependency between attai-
ned amplitude and initial phase difference. The following explanation is
suggested: At very low forcing levels the flow is quasiparallel. For parallel
flow many theoretical works8,10 show a relation between attained amplitude
of the subharmonic and the initial phase difference between the two excited
waves. At higher forcing levels the flow can no longer be assumed to be
parallel, as displayed in Fig. 11 (the dashed line represents the_evolution of
the momentum thickness for the high forcing level, (uj}max/U c\ = 3.2%.
Therefore, the phase speed of the subharmonic and the fundamental is no
longer constant across the shear layer,11 and the influence of the initial phase
difference vanishes. At very high forcing levels at a certain distance down-
stream of the nozzle, the flow again becomes parallel in a defiened region,
due to saturation.4 There the subharmonic can be suppressed or amplified,
depending on the phase angle between the two waves .
The presence of a strong subharmonic in the flow enhances the mixing
by enhancing the rate of spread of the mixing layer with x (Fig. 11). The
evolution of the momentum thickness for three cases, (1) natural jet, (2) flow
62 C. O. PASCHEREIT AND I. J. WYGNANSKI

0.020

0.015-

O
0.010-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

O.OOSi

0.000

Fig. 9 The variation of the subharmonic along the centerline with_x/Z> for A$ =
0° and A$ = 90°, // = 368 Hz, fs = 184 Hz, (u/} ma z/^ci = 0.5% at
x/D = 0.2.

0.20

0.15-

O
0.10-

0.05-

0.00
2 X/D

Fig. 10 The variation of the subharmonic along the centerline with x/D for A$ = 0°
and A^ = 90°, // = 124 Hz, /, = 62 Hz, ( u f ) m a t / U c f = 1.7% at x/D = 0.2.
AXISYMMETRIC JET INSTABILITY 63

2.5
-a— unforced
-A - • • combined, A4>=0°, <uf>max ^.Q 25/yd=1.6%
- x- - combined, AflM)°, <uf>max; ^loW^r3-2^'
2.0

E
c
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1.0

x/D
Fig. 11 The evolution of the momentum thickness with x/D.

forced at both 184 and 368 Hz for an initial forcing level of (uf)max/Uci =
1.6%, and (3) flow forced at both 184 and 368 Hz for an initial forcing level
of (uf)max/Uci = 3.2%, is plotted in Fig. 11.

Conclusions

The influence of the form of the initial excitation on the subharmonic


resonance was investigated. The results corresponding to a single-frequency
excitation confirm the existance of resonance conditions suggested by Cohen
and Wygnanski.12 Energy can be transferred from one wave to another,
provided both waves are non dispersive and have sufficient time for the
exchange.
The results obtained from a concurrent excitation at two frequencies in-
dicate that the maximum level of the subharmonic attained depends on the
imposed phase difference between it and the fundamental. For an initial
phase difference of 90 deg the maximum level of the subharmonic is gene-
rally suppressed. The amplitude ratios between the subharmonic and the
fundamental wave do not influence the damping significantly, as long as the
energy in the coherent motion is kept constant.
The growth of the subharmonic becomes independent of the initial phase
difference when the total, combined forcing level is doubled. The flow is no
longer quasiparallel and therefore the phase speed is not constant across the
shear layer. A phase-locked arrangement used by Monkewitz10 to explain
the amplification or suppressing is not possible over the whole width of the
64 C. O. PASCHEREIT AND I. J. WYGNANSKI

shear layer, because of the different phase speeds: thus, the influence of the
initial phase difference vanishes.
Additional measurements taken only in the centerline of the jet at high
forcing levels (i.e.,0.5% < (uj)/Uci < 1.7%) indicate a small dependency
between attained amplitude and initial phase difference. A mixing layer,
forced at high amplitudes, has after a region of strong spreading a region
where the width remains almost constant. In this parallel region a phase-
locked arrangement of vortices across the whole width of the shear layer is
possible and thus there is an influence of the initial phase difference.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Acknowledgments
This work was supported by National Science Foundation under Grant
MSE 8800086 and by Deutsche Forschungsgemeinschaft.

References
1
Cohen, J., Wygnanski, I., 1987. "The evolution of instabilities in the
axisymmetric jet. Part 1. The linear growth of disturbances near the
nozzle", Journal of Fluid Mechanics, Vol. 176, 1987, pp. 191-219.
2
Cohen, J., Wygnanski, I., "The evolution of instabilities in the axisym-
metric jet. Part 2. The flow resulting from the interaction between two
waves," Journal of Fluid Mechanics, Vol. 176, 1987, pp. 221-235.
3
Roberts, F.A., "Effects of a periodic disturbance on structure and mi-
xing in turbulent shear layers and wakes," Ph.D.-Thesis, California Institute
of Technology, 1984.
4
Crow, S.C., Champagne, F.H., "Orderly structure in jet turbulence,"
Journal of Fluid Mechanics, Vol. 48, 1971, p. 567.
5
Oster, D., Wygnanski, I., "The forced mixing layer between parallel
streams," Journal Fluid Mechanics, Vol. 123, 1982, pp. 91-131.
6
Weisbrot, I., Wygnanski, I., "On coherent structures in a highly excited
mixing layer," Journal of Fluid Mechanics, Vol. 195, 1988, pp. 137-159.
7
Fiedler, H.E., Mensing, P., "The plane turbulent shear layer with peri-
odic excitation," Journal of Fluid Mechanics, Vol. 150, 1985, pp. 281-309.
8
Kelly, R.E., "On the resonant interaction of neutral disturbances in
inviscid shear flows," Journal of Fluid Mechanics, Vol. 31, 1967, p. 789.
9
Mankradi, R.R., "On the interaction between fundamental and subhar-
monic instability waves in a turbulent round jet," Journal of Fluid Mecha-
nics, Vol. 160, 1984, pp. 385-419.
10
Michalke, A., "Instabilitat eines kompressiblen runden Freistrahls unter
Beriicksichtigung des Einflusses der Strahlgrenzschichtdicke," Zeitschrift fur
Flugwissenschaften, Vol. 9, 1971, pp. 319-328.
11
Monkewitz, P.A., "Subharmonic resonance, pairing and shredding in
the mixing layer," Journal if Fluid Mechanics, Vol. 188, 1988, pp. 223-252.
12
Caster, M., Kit, E., Wygnanski, I., "Large-scale structures in a forced
turbulent mixing layer," Journal of Fluid Mechanics, Vol. 150, 1985, pp.
23-39.
13
Arbey, H., Ffowcs Williams, J.E., "Active cancellation of pure tones in
an excited jet," Journal of Fluid Mechanics, Vol. 149, 1984, pp. 445-454.
Three-Dimensional Vortex MHD Flows at High
Reynolds Numbers in Thin Layers of Conducting
Incompressible Fluid
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

V. D. Zimin* and N. Ju. Kolpakovt


Institute of Continuous Media Mechanics, Perm, Russia

Abstract
A closed system of equations for three-dimensional vortex MHD flows in
thin layers at high Reynolds numbers through local-equilibrium boundary layer
equations is developed Because fluid is exchanged between the boundary layer
and the core flow, classical boundary layer theory is inappropriate. In essence,
Kolpakov reduced the three-dimensional equations to two-dimensional equations,
and then reconstructed the full solution from the two-dimensional results. The
solution technique is demonstrated on a problem involving an aluminum
electrolyzer. The results of the local-equilibrium boundary layer method are in
reasonable agreement with a numerical solution, with the exception of the axial
velocity component. This discrepancy can be rectified by using higher-order
approximations to the boundary layer velocity components.

Introduction
We consider three-dimensional (3-D) vortex MHD Hows of viscous con-
ductive liquid in a planar layer with solid boundaries which are induced by the
electromagnetic forces uniform through thickness of the layer. Reynolds number
of the vortex large-scale flows, which characteristic scales are essentially greater
than the thickness of the layer, is suggested large, that allows to use the method
growing together asymptotic decompose for investing the velocity field. The
local-equilibrium solutions have been found for the 3-D boundary layer on the
solid boundary, which parametrically depends on the differential characteristics
of inviscid flow nucleus. On this basis the equation has been obtained, which
describes the vortex structures dynamic in the layer and differs from the pre-

Copyright © 1989 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Chief of Laboratory.
fSenior Scientist
65
66 V. D. ZIMIN AND N. JU. KOLPAKOV

viously known more consequent accounting for the kinetic energy dissipation
mechanisms.
Accurate theoretical investigations of the 3-D boundary layer vortex flows
and their stability in closed volumes and fluid layers are not available in the
published literature. As for numerical analysis of such problems, it is at the
upper limit of modern supercomputer capacity.
Classical theory of boundary layers is not appropriate for analytical treat-
ment of fully developed nonlinear regimes of vortex flows in plane fluid layers,
because this situation involves violation of the Friedriks condition, which is as-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

sociated with a weak influence of the boundary layer on the character of the
external flow. In fact, the flow curvature in the inviscous flow nucleus involves
the appearance of pressure gradients (both in longitudinal and transversal
streamline directions) which are transferred in their initial form to the boundary
layer and induce the fluid motion toward the center of the external flow
curvature. In the local cylindrical coordinate system, such radial flows give rise,
due to continuity equation, the ascending axial flux from the boundary layer,
which in turn contributes to the generation of inviscid flow.
It seems rather inefficient to describe 3-D vortex structures in a thin layer
using "direct" variational methods (Kantorovich, Bubnova-Galerkin, contour in-
tegral method, and so on) which are based on reduction of 3-D problems to 2-D
by averaging the equations across the fluid layer. As a result of continuous shape
variation of vertical profiles of hydrodynamic fields, which takes place point by
point in a plane of the layer, the model equations developed by the averaging
method give inadequate dynamics of large-scale vortex flows in the layer.
Conversely, formation of the boundary layers reduces characteristic scales
of the shear layers and increases the flow stability (as compared to the
frequently used Poiseuille velocity profile) since stability is now defined by the
Reynolds number Re 8 estimated along the boundary-layer thickness rather than
along the fluid-layer thickness. In this case, for a wide range of Re^ values there
exists a problem of obtaining a self-consistent solution for inviscid flow nucleus
and boundary layers up to the flow turbulization across the layer. To this end,
one should take an advantage of general concepts of match asymptotic
expansion methods [1,2], though this computer implementation required
searching for at least approximate solutions for nucleus with parametric
dependence on the boundary-layer response and for boundary layers depending
parametrically on the inviscous flow nucleus. The object of the present work is
to develop a closed system of equations for the large-scale vortex flows in a
layer, based on the local-equilibrium solutions for 3-D boundary layers.

Model Equations
Let us consider 3-D vortex flows in a plane horizontal layer of conductive
fluid with solid boundaries, which are created by the body forces (electromag-
THREE-DIMENSIONAL VORTEX 67

netic forces, gravity forces, and so on) homogeneous through the fluid layer. It is
assumed that magnetic Reynolds numbers are low, and the flow effect on the
distribution of electromagnetic forces f(r) is negligible. In Cartesian coor-
dinates related to the lower boundary of the layer (z axis has an upward
direction) the fields of velocity v=(u,v,w) and pressure p(x,y,z) are described by
the relation.1

3v r r l r rr r „.
—- + ( v V ) v = — V v + r A v g + f , divv = 0 (1)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

dt p
with no-slip conditions imposed on the layer boundaries z=0,h.
Let us assume that, for the flow regimes under consideration, the Reynolds
number Re h estimated through layer h is high and the viscosity far from the
boundaries of the region is negligible. The characteristic horizontal scale of the
flow L is much larger than the layer thickness. In this case, a transverse com-
ponent of the Navier-Stokes equation (1) is reduced to the hydrostatic equation
(3Z p = -pg), which upon integration over z jells layer distribution of pressure

p(x,y,z) = -pgz + P(x,y) (2)

where P(x,y) is the unknown coordinate function.


In inviscous flow nucleus the horizontal velocity is independent of the
transverse coordinate z. This assumption agrees with uniformity of electromag-
netic forces and the horizontal pressure gradient across the layer. Furthermore, a
vertical shear rate is inconsistent with the planar vorticity (Praudman-Taylor
theorem3'4). So, the velocity field takes the form

v = V(x,y,z)+v(x,y,z) (3)

where the inviscous flow V = (U(x,y,z), V(x,y,z), W) satisfies the Eulaer and
continuity equations

Ut+UUx+VUy=-Px/p+ F (4)

Vt+UVx+VVy=-Py/p +G (5)

Ux+Vy+W2=0 (6)

and component v defines the boundary-layer structure of the flow. Her the
indices t,x,y,z denote differentiation with respect the appropriate variables,
F(x,y) and G(x,y) and horizontal components of the electromagnetic forces.
In contrast with traditional asymptotic approach,5 6 the velocity field of the
inviscous flow will be considered along the entire thickness of the thin layer,
68 V. D. ZIMIN AND N. JU. KOLPAKOV

and boundary layers will be defined s regions with nonzero relative velocities
v r = v - V. Then, the sewing condition for solutions is transferred at the bound-
aries of the fluid layer,

u' = -U, v' = -V, wf = -W for z = 0, h (7)

whereas at the outer bounder of the boundary layer,

u' = 0, vf = 0, w' = 0 for z = 8j(x,y) (8)


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where 8i is the thickness of the lower (i=l) or upper (i=2) boundary layer. Using
Eqs. (4-6), 3-D boundary-layer equations can be written in terms of relative
velocity as,

u f t +u' Ux + v' Uy + u' x U+ u' y V + u' 2 W = vu'^ (9)

v' t +u! Vx + V V y + u!x U + v' y V+ v', W = vv f „ (10)

u' x +v' y +w' 2 = 0 (11)

In Eqs. (9) and (10), the terms (vV)v are neglected since all nonlinear terms in
the vicinity of the solid surface are small as compared with the viscous terms,
and on the outer boundary v —»0, sot hat the neglected terms are lower than
terms (vV)V and (vV)V.7
As follows from the analysis of Eq. (6), W is the linear function of the z
coordinate.

W(x,y,z) = -(Ux + V y )z + Wl(x,y) (12)

where Wl is the velocity of the fluid motion through the surface z=0 into the
flow nucleus. The relative velocity of the flow from the boundary layer at z=0 is
equal to w'=-Wl: this implies that the no-slip condition is satisfied at the solid
boundary. At low 81^ relation (12) can be neglected for the thickness of the
boundary layer assuming that at 0 < z < &i W=W1. Subtracting Eq. (6) from Eq.
(11) and integration over z from 0 to 8 gives

(13)

Thus the appearance of "fluid pumping" into inviscid flow at the boundary z=0
is due to divergence of relative flows across the boundary layer. The fluid ex-
change between boundary layers and flow nucleus provides the inverse flow of
THREE-DIMENSIONAL VORTEX 69

vorticity from the large-scale flows of the layer. It is to be noted that the above
considerations hold true for W2 at the upper boundary layer, which is equal to -
Wl by virtue of symmetry of the problem.

Local Equilibrium
Let us consider the local-equilibrium solutions for the system of Eqs. (7-11)
bearing in mind that essential variation of the planar velocity field (u,v) takes
place at distances on the order of L and at distances on the order of 81
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(8 « h « L). Values of x,y derivations of u,v are assumed to be constant and


the time of boundary-layer flow relaxation to the state of equilibrium with the
external flow is much less than the evaluation time of the latter L/U.7 In this
case, the stationary solution is to be sought in the form

uf = a(x,y)ex<x'y)z, v' = b(x,y)ex<x>y)z (14)


Functions a,b and c depend parametrically on the quantities u,v, and w and their
derivations, and in the vicinity of arbitrary point A with coordinates XQ, y0 the
can be extended in the Taylor series:

+ a x (A)(x-x 0 ) + a y (A)(y-y 0 )+... (15a)

y 0 )+... (15b)

y0)+... (15c)

Requirement of a large-scale range for local equilibrium solutions allows


restriction of a zero-order approximation to the first terms of expansions (15).
Then, substitution of Eq. (14) into Eqs. (9) and (10) gives the relation for
coefficients a and b; as far X this reach algebraic equation of the fourth degree,
with its roots written as
X n = [Wl ± (Wl2 + 2v(Q ± (Q2 - 2K))) 1/2 ] / 2v

where n = l~4, Q = Ux + V y , and K>U x V y -U y V x ,


The quantities Q(x,y) and K(x,y) are differential tensor invariants of the
planar velocity field in the inviscid flow nucleus. If for the specified velocity
field (u,v,w) there exist among ^ at least two district roots with negative real
parts, then using Eq. (14) one can obtain a solution that will satisfy boundary
conditions (7) and (8). According to Eq. (16), development of the local-equili-
brium boundary-layer solution depends on the values of invariants Q and K,
characterizing the inviscid nucleus at the given point, and Wl — rate of fluid
exchange between the boundary layer and the nucleus. Diagrams of all roots Xn
at the plane of parameters Q and K are shown Fig. 1.
70 V. D. ZIMIN AND N. JU. KOLPAKOV

-13
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

AOKAAbHO-PABHO&ECHblX
PEUJEHMM

Fig. 1 Map of local equilibrium solutions.

If Xj and Xk are the required roots [see Eq. (16)], the flow profile of the
boundary layer will be represented as

(17)

where

X k U(vX k -Wl)-UU,-VU y

and expressions for a^ and b^ follow from ai and bi after substitution


by Xk (X{). With regard to exponentially strong damping u,v as z increases one
obtains Wl from Eq. (13) by substituting the infinite limits of integration for the
upper ones:

xi7i / if -fv t. )dz


xj = ai avl d \ b \ bv1
-Wl=| (u x v - + -± -f T- - + T- (18)
0
THREE-DIMENSIONAL VORTEX 71

the implicit relation [Eq. (18)] may be regarded as the boundary condition for
Eq. (6) at z=0.
Reference 8 represents the local-equilibrium solutions as compared with a
number of accurate solutions of Navier-Stokes equations, which belong to the
von-K£rmdn class with the variable of self-similarity H = z/1. A fair argument
has been obtained between the approximate analytical equation (17) and the ac-
curate solution. Thus for a planar flow, the velocity components of potential
flow at the front critical point are proved to be-equal.5

U=cx, V=0, W=-cz+Wl


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where c and Wl are dimensional constants. The values of invariant Q = c > 0,


k = 0 define, according to Fig. 1, the type of boundary layer at the solid surface:
its local-equilibrium profiles are represented as:

u ' °'
The values of the u/U function of the coordinate i\ (dashed liens) are compared
in Fig. 2a with the accurate solution9 (solid lines).
For spatial flow in the vicinity of the critical point10

U=cx, V=0, W=-cz+Wl


there exists, respectively, Q = 2c> 0 and K = c2 > 0, which satisfy the relation
Q = 2-/K [the right-hand branch at the diagram of the local-equilibrium
solutions (Fig. 1)]. Profiles for such axisymmetric flow written in cylindrical
coordinates (r,cp,z)

e-^), v = v =0

w =v —

are shown along with the accurate solution10 in Fig. 2b. It should be noted that
the z component of the potential flow differs from W in Ref. 10 by the rate of
fluid exchange between the inviscid nucleus and the boundary layer
Wl = V4vc/3 (Wl = Vvc/2 is for the plane case). Here, the previously
infeasible 3 asymptotic condition of the velocity field transition to inviscid flow
at r| —» oo has been satisfied.
72 V. D. ZIMIN AND N. JU. KOLPAKOV
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

2/J ^
Fig. 2 Comparison of approximate analytical solution with known
accurate solutions: a) boundary layer at the solid surface; b) spatial
flow in the vicinity of the critical point; c) rotation motion of the
fluid over the fixed base.

For rotational motion of fluid over the fixed base,11 the inviscid velocity
field is solenoidal, and in cylindrical coordinates is written in the form6:

U = 0, = rQ, W = W1

thereby its invariants are assumed as Q = 0, K = Q2 > 0. Here Q is the angular


velocity of quasi-solid-type rotation of fluid. In this case, the local-equilibrium
flow profiles have the following form 8:

sinX2T|, )cosX27i

w= 2 *z cos X T i -
! + A,2
THREE-DIMENSIONAL VORTEX 73

where X t =0.43, X^ = 0.89 andi = z^/v . The approximate solutions to Fig.


2b (dotted lines) are compared with the accurate solution (solid lines) derived in
Ref. 11 by numerical methods. Elevation (vertical) profiles of the radial (1), azi-
muth (2), and axial (3) components of fluid velocity are depicted separately. To
reduce the difference between asymptotic values for the z component of velo-
city, it is necessary to allow for the following terms of expansion [Eq. (15)].

Equations for Stream Function and Potential


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Let us represent the inviscid velocity field as a sum of vortex and potential
component

U = -¥ y + <t> x , U = -¥.+<>, (19)

then, for the flow invariant Q and K, describing layer planar divergence and
circulation, there exist the relations

Q^+tfr (20a)

(20b)
rather than Eq. (16).
In a closed-flow region (flat cuvette), the flow is induced only be a vortex
component of electromagnetic forces. Then, the potential component of the
nucleus flow, which is due to fluid exchange between boundary layers and the
nucleus, is small as compared with the vortex component. Since the rate of ex-
change Wl (or W2) is of the order Qh, the velocity profiles in the boundary
layer will have the following form, under the assumption of satisfied inequality
4K » Q2. 4vVK » (QH)2 for Eq. (16):

. UUX 4- VUVL .
u' = -Ue coskz +————=— 2
e sinkz (21a)
2VK

UV + W
vf = -Ve"ta coskz + —^—r-2-*'* sinkz
(21b)

where KsK 1 / 4 /V2"V. In this case, the mechanism of boundary-layer


generation is similar to that of the Bedevadt boundary layer11 or Eckman layer in
rotating fluid.4 The boundary layer thickness is defined by a local balance of
frictional electromagnetic and centrifugal forces. However, unlike the Eckman
74 V. D. ZIMIN AND N. JU. KOLPAKOV

pumping,4 which changes its sign with alteration of sign of vortex rotation in
inviscid flow, the vertical velocity at the bottom (top) surface of the layer inside
any of the vortices in stationary layer is always positive (negative)., Indeed, after
substitution Eq. (21) into Eq. (18) the result can be written as a first approxi-
mation to nonuniformity of the planar velocity field in the inviscid nucleus

Wl = -W2 = K1/4 / V2v (22)

If the vortex and potential components of electromagnetic force are denoted as F


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

and I respectively

r = -Fy + G x , I = Fx+Gy (23)

then system of Eqs. (4-6) in terms of stream function *¥ and potential ty will be
written as

=T (24)

= I-p" 1 AP (25)

(26)

(27)

(28)

Here, closing relation (28) has been obtained using Eq. (12) at z=h and Eq. (22)
and the following notation has been adopted: Act = a x x +a y y ,
Va(3 = ax(3x -ctyPy, {a,fi} = a x P y -a y p x —are Poisson brackets. For
boundary conditions of the system of Eqs. (24-28) one assumes *¥\D = 0.
V<t>n|D = 0 at the contour of region D. The former implies that the planar
velocity field is decomposed into vortex and planar components, and the latter
means that the total flux over the contour of region D is equal to zero.
Thus, dynamics of large-scale vortex flows in a layer at high Reynolds
numbers is described by closed equations (24-28) which in contrast to
previously known relations allow for dissipation mechanism of kinetic energy in
a more consecutive manner. According to distribution of electromagnetic forces
over the layer plane equilibrium with F(x,y) in the vortex transfer equation (24)
can be achieved either due to {¥,£2} or to a potential component of the velocity
field despite its smallness with respect to a vortex component is proved to be a
necessary conclusion for existence of 3-D vortex structures. Potential <)>
describes the fluid on inflow in the boundary layer at the horizontal and vertical
borders, interaction of vortices of different intensities, wave processes, and so
THREE-DIMENSIONAL VORTEX 75

on. The present approach allows application of 2-D model equations (24-28) in a
wide range of Reynolds numbers to describe laminar regimes up to the closure
of boundary layers, and conversely, to describe evolution of the large-scale
turbulence in a plane layer up to appearance of instability of equilibrium
velocity profiles with consequent turbulization of flow across the layer.

Flows in Rectangular Regions


To illustrate the preceding statements, let us consider nonlinear regimes of
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

laminar MHD flow, induced in a rectangular cuvette wit sides L and D and
height h » L. D by the vortex electromagnetic force

F = -xC, G = yC, 1=0, F = const = C


Solution of the system of Eqs. (24-28) is given in terms of the elliptic vortex
with constitution scalar, defined as

^k^e-V + ey2), <t> = k2(t~lx2 + £y2), P = k3(e'1x2 + ey 2 ) (29)

where £ = L / D is eccentricity of the cuvette. Substitution of Eq. (29) into Eqs.


(24), (25), an d(28) gives according to the following expression for coefficient lq
(i=l,2,3);

k t - -(r1 + e)-2/3S1/3 Vf, k 2 = (e-1 + e)-47^'173 Vf (30a)

k 3 = -(£- l +£)- 7/3 S 4/3 F (30b)

where parameter S = h 2 Vf / v characterizes the intensity of induced motion.


According to Eqs. (12), (19), and (29), the velocity field in inviscid nucleus

U = 2e"1k2x - 2ekiy, V = 2z~lklx + 2ek 2 y (3 la)

W = k 2 (£- 1 +£)(h-2z) (31b)

defines superposition of double-layer meridian MHD flow with total differential


rotation. The flow is symmetric with respect to a plane z=h/2. While untwisting
around the left-hand spirals at F > 0 the fluid drops at the side walls and rises in
the center: Wl = k 2 (£ -1 + e) h > 0 over the entire solid surface z = 0.
Vertical flow profiles are depicted in Fig. 3a.
It is necessary to add the terms vA<3 and vAQ to account for the friction at
the side boundaries of the cuvette in the equation for the vortex equation (24)
and the potential equation (25), respectively, In this case, the system of Eqs. (24-
76 V. D. ZIMIN AND N. JU. KOLPAKOV
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 3 MHD How in rectangular cuvette: a) velocity profiles; b, c,


and d) isolines of stream function, potential, and vorticity, cor-
respondingly.
THREE-DIMENSIONAL VORTEX 77

28) requires numerical investigation. Calculations were made using the grid
method with the aid of the explicitly difference scheme and the knots number is
71 to 21. Solutions of Poisson equations (26) and (27) were calculated by the
Zeidel method, here are isolines of the flow function *F, vorticity ft, and
potential <|> for L=14 m, D = 4m, h = 0,3 m, v = 3.67 x 10 '7 m2/2. and F = 3 x
10'5 s"2 in Figs. 3b, c, d. Maximal motion speed was 0.1 m/s which is equivalent
to the planar Reynolds number ReL = vmaxxL / 2v ~ 2xl0 6 . Obtained stationary
regimes point at the fact, that in cuvette by constant-rotor force influence the
one-vortex counterclockwise (F > 0) flow creates. The vortex is non-parallel in
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

comparison with the elliptical equation (29) and its orientation is directed along
the diagonal of the motion, that is typical for the nonlinear regimes of the vortex
flows.
According to Eqs. (20) and (29) invariants of the one-vortex flow equal
Q = 2(e + e~ 1 )k 2 , K = k?,+kf « kf. Then the velocity profiles in the
boundary layer can be found from Eqs. (21) and (31), and for example, in a
point with coordinates x=0, y = D/2 looks such way:

sin KZ) (32a)

v EE V(l - e'*2 cos KZ + (4e(e"1 + e)l/3 S'273 -


--e-V1 + £)2/3S-2/Vkz sinK^ (32b)
4

W = k 2 (e" 1 + e)K-1[-2Kz -f 1 - e~ KZ (cosKz - sin KZ)] -


(32c)

where K = K1/4 / *j2v = V a / v . The pots of functions u/U, v/V, and w/W from
dimensionless coordinate t| = KZ for the different values of parameter S are
shown in Fig. 4.
In conclusion, it is worth noting that such a method can be used for
describing the vortex flows in different strong-flow technology processes and in
industrial MHD plants. As a typical example, let us consider calculation for
nonlinear regimes of electro- vortex motions in an aluminium electrolizer, which
takes place due to interaction between electrical current with its own magnetic
field. (For more mathematical modeling of hydrodynamically electrolizer pro-
cessing, see Ref. 12). In the electrolizer with the discontinuous anode the axis
electrical current density component is described by the expression

(33)
where j0 = I 0 /Id, I0 -total current through the electrolizer, 1, d-anode scales,
m ± (x) - asymmetrical unit stair functions. If we introduce the vector potential
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

78

0
L
V. D. ZIMIN AND N. JU. KOLPAKOV

10

Fig. 4 Velocity profiles in MHD boundary layer.


THREE-DIMENSIONAL VORTEX 79
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

b) 0/2

Fig. 5 Electrovortex motions in aluminium electrolizer: a) distrib-


ution of forces; b, c) isolines of stream function and potential.

of magnetic field A = (0,0, A), then from Eq. (23) we obtain that

|ior = (A, AA), ml = -(VA)2 - VA x V(VA) (34)

For the axis electrical current the electromagnetic force rotor in the layer is
different from zero only at a contour C, which is confined with the anode
projection. Distribution of the scaled T(x,y) and I(x,y) along contour C is shown
in Fig. 5a.
Numerical investigation of the system of Eqs. (24-28) was made using the
same calculation scheme as in the previous case. For example, calculations for
80 V. D. ZIMIN AND N. JU. KOLPAKOV

L = 14, D = 4, h = 0.3,1 = 13.4, d = 3 (meters) and the total current I0 = 160 kA


show that the stationary regimes with the system of four vortex exist in
electrolizer. Turning direction of diagonal vortex coincidence, does not depend
on direction of current through the volume and defines by net action of vortex
contour C and the region boundary. Isolines flow function and the velocity field
potential for four vortex flow for Reh = 105 are represented on Fig. 5b-c.

References
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

^hlichting, G., Teoriya pogranichnogo sloya, Nauka, USSR, 1974 (in Russian).
2
Shevelov, Ju. D., Trechmernie zadachi teorii laminarnogo pogranichnogo sloya. ,
Nauka, USSR, 1977 (in Russian).
3
Vazov, V., Asimptoticheskie razlogenia reshenii obiknovennich differencialnich
uravneniy, Mir, USSR, 1968 (in Russian).
4
Shevelev, Ju. D., Prostranstvennie zadachi bichislitelnoi aerogidrodinamiki, Nauka,
USSR, 1986 (in Russian).
5
Landau, L. D., Lifshic, E. M., Gidrodinamika, Nauka, USSR, 1986 (in Russian).
6
Pedloski, Dg., Geofizicheskaya gidrodinamika: v 2 tomach, Mir, USSR, 1984 (in Rus-
sian).
7
Zimin, V. D., Kolpkov, N. Ju., and Popova, E.V., "Vichrevie techeniya v tonkom sloe
provodyashei gidkosti pri bolshich chislach Reynoldsa," Magnitnaya godrodinamika,
No. 2, 1989, pp. 73-80 (in Russian).
8
Kolpakov, N. Ju., "Nelineynie effekti v trechmernich vichrevich MGD-techeniyach
nesgimaemoy gidkosti," Diss. ... kand. fiz,-mat. nauk, Perm, 1988 (in Russian).
9
Howarth, L. 'The boundary layer in three-dimensional flow, p. 2. The flow near a
stagnation point," Phil. Mag., Vol. 43, No. 43, 1951, pp. 1433-1440.
10
Fr6sling, N., Verdunstung. "Warmeiibertragung und Geschwindigkeit-sverteiling bei
zweidimensionaler und rotationssymmericher laminarer Grenzschichtstromung. Lunds
Univ., Arsskr. N.F. Avd. Vol. 2, No. 35, 1940, p. 4.
H
B6'derwadt, U. T., "Die Drehstrb'mung iiber festem Grund, " ZAMM, Vol. 20, 1940,
pp. 241-253.
12
Almuchametov, V., and Chripchenko, S.Ju., "Mechanizmi generacii elektrovichrevich
techeniy v vanne elektrolizera so sploshnim anodom," Magnitnaya gidroinamika, No. 3,
1987, pp. 101-104 (in Russian).
Tearing Instabilities in Two-Dimensional
MHD Turbulence
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

H. Politano,* A. Pouquet,t and P. L. Sulem$


Observatoire de la Cote d'Azur, Nice, France

Abstract
Direct numerical simulations of decaying two-dimensional
MHD flows at Reynolds numbers of several thousand are pre-
sented. An inertial range extending to about one decade is
observed. At very small scales, resistive tearing instabilities
destabilize current sheets generated by the inertial dynamics
and lead to the formation of small-scale magnetic islands,
which may then grow and reach the size of inertial scales.
High Reynolds number magnetohydrodynamic (MHD) flows
display both turbulence features like power law inertial ranges
and resistive instabilities of current sheets that develop at small
scales. These tearing instabilities, which lead to reconnection
of the magnetic field and changes of flow topology, are generally
studied by perturbing weakly a highly ordered magnetic pinch
in a flowless plasma.1'2 This magnetic configuration idealizes
the result of free evolution of a high Reynolds number MHD
flow with the initial energy located at much larger scales. The
nonlinear dynamics indeed leads to the formation near neutral
X-points of magnetic current sheets, corresponding to strongly
sheared magnetic fields.

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Laboratoire Giovanni Domenico Cassini.
t Laboratoire Giovanni Domenico Cassini; also at High Altitude Observatory,
NCAR, Boulder, Colorado.
J Laboratoire Giovanni Domenico Cassini; also at School of Mathematical Sci-
ences, Tel-Aviv University, Israel.
81
82 H. POLITANOETAL

An important question is whether a turbulent environ-


ment preserves the qualitative properties of resistive instabili-
ties. To address this question, we numerically simulate the free
decay of unit Prandtl number two-dimensional incompressible
MHD flows with large scale initial conditions. A resolution of
(1024)2 collocation points in a periodic geometry enables us to
precisely resolve all the scales, including the most dissipative
ones, within flows with Reynolds numbers of several thousand-
s. These simulations reveal both power law inertial ranges and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

tearing instabilities that destabilize current sheets formed near


neutral X-points.3
A prototype of two-dimensional MHD flows is the Orszag-
Tang (OT) vortex defined by the initial stream function
i(>Q(x,y) = 2(cosx + cosy) and magnetic potential ao(x,y) =
2 cos x + cos 2y. These initial conditions correspond to a ki-
netic vortex with a central stagnation point and two magnetic
vortices with a central neutral X-point. The viscosity and mag-
netic diffusivity are v — r/ = 10~3, in order to resolve precisely
all the scales of the motion. The initial kinetic and magnetic
Reynolds numbers are, thus, of order 104.
Figure la shows the energy flux HT(k) vs the wave-
number fc, at time t ~ 3.85. This flux is almost constant
in an "inertial range" extending approximately on 5 < k < 25.
The kinetic and magnetic energy spectra E% and E^1 are p-
resented in Fig. Ib. They display an (exponential) dissipation
range whose logarithmic decrement agrees with the dissipation
wave-number kj w 25 defined from the energy flux.
Figure 2 shows the contours of the stream function and
magnetic potential at t — 0 and t == 3.55. We note the pres-
ence of a magnetic current sheet located on the central kinetic
vortex. This eddy distorts the magnetic sheet and induces a
tearing instability. At times of tearing, strong bursts are visible
on the time evolution of the small scale kinetic and, above all,
magnetic modes. This amplification mainly affects the modes
with k > 80, which correspond to scales much smaller than
the dissipation scale. Figure 3 visualizes local snapshots of the
magnetic potential and of the stream function near the center,
at time t = 3.55 and t = 4.70; we see the formation of four mag-
netic islands, by reconnection. These islands may grow, reach
a sizeable fraction of the domain, and survive a very long time.
Small magnetic bubbles with a short lifetime are also generat-
ed. In this example, the tearing instability takes place in the
presence of a shear flow, and "cat's-eye" structures, correlat-
ed to the magnetic islands, are visible on the stream function.
TEARING INSTABILITIES 83
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0 10 20 ID fO SO 10 ID ID «) 100 UQ

Fig. la Energy flux IIT(fe) in the range 0 < k < 110 for the OT
vortex at t = 3.85.

10°
ur'

icr*

,0-10
,0-11

,0-13
IO-M
,0-15
10»

Fig. Ib Kinetic and magnetic energy spectra E% (solid line) and


E%* (dashed line at t = 3.85 for the OT vortex.
84 H. POLITANO ET AL.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

t = 0.00 t = 3.55
Fig. 2 Contours of constant magnetic potential (top) and stream
function (bottom) of the OT vortex at t = 0 and t = 3.55.

Other reconnection processes are seen, for example, when two


large magnetic vortices are pushed towards each other.
To test the general character of the observation made on
the OT vortex, we performed other numerical simulations with
various random initial conditions. When turbulence becomes
developed, the flows display a multiplicity of small scale activ-
ity regions in the vicinity of magnetic current sheets.
Figure 4 shows a cut of the current in such a random
flow computed on a (512)2 grid, when the flow has evolved
for roughly four eddy turn-over times. Although eddies are
space filling, the dissipative structures are more sparse.4 We
also observed the appearance and growth of magnetic islands,
TEARING INSTABILITIES 85
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

= 3.55 t = 4.70
Fig. 3 Local snapshots of the magnetic potential (top) and
stream function (bottom) near the central sheet of the OT vortex
at t = 3.55 and t = 4.70.

together with phenomena strongly evocative of impulsive


bursty reconnection.5
The simulations just reported thus demonstrate that the
basic features of resistive instabilities, which may generate sta-
ble coherent structures, survive in a turbulent environment.

Acknowledgments
The numerical simulations were performed on the CRAY 2 of
the Centre de Calcul Vectoriel pour la Recherche (Palaiseau,
France), and on the CRAY XMP of the National Center for At-
mospheric Research (Boulder, Colorado). The National Center
for Atmospheric Research is sponsored by the National Science
Foundation.
86 H. POLITANO ET AL.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 4 One-dimensional cut of the magnetic current for a random


flow.

References
Furth H. P., Killeen J., and Rosenbluth M. N., "Finite
Instability of a Sheet rinch," Physics of Fluids, Vol. 6,
1963, pp. 459-484.
Matthaeus W. H. and Lamkin S., "Turbulent Magnetic
Reconnection," Phys. Fluids, Vol. 29,1986, pp. 2513-2534.
Politano H., Pouquet A. and Sulem P. L., "Inertial Ranges
and Resistive Instabilities in Two-Dimensional MHD Tur-
bulence," Physics of Fluids B, Vol. 1,1989, pp. 2330-2339.
Passot T., Politano H., Pouquet A. and Sulem P. L., "Sub-
grid Scale Modelling in Two-Dimensional MHD Turbu-
lence," Theoretical and Computational Fluid Dynamics,
Vol. 1, 1990, pp. 47-60.
Priest E.R., "The Magnetohydrodynamics of Current
Sheet," Report Progress Physics, Vol. 48, 1985, pp. 955-
1090.
Axisymmetric Hydromagnetic Dynamo
M. A. Goldshtik* and V. N. Shternt
Siberian Branch of the Academy of Sciences,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

630090 Novosibirsk, Siberia

Abstract
A mechanism of axisymmetric dynamo is reported. There are no
contradictions with Cowling's and Braginsky's antidynamo theorems because
their conditions are not satisfied here. Velocity and induction are inversely
proportional to the distance from the origin, and magnetic lines and streamlines
are not closed. The dynamo is found in jetlike flows of viscous incompressible
electrically conducting fluid, particularly the Squire jet and a heat-convection
flow, which may serve as the simplest models of cosmic jets observed near
young stars and galaxy cores. The bifurcation of magnetic field is pitchfork. It
is supercritical for nonswirling jets and subcritical in high swirling flows.
Besides the lamina case, a simplified model of the turbulent dynamo is also
considered. It is found that the self-generation of magnetic field leads to
relamination of the flow. In a turbulent heat-convection flow near a star,
bifurcation of self-swirling regimes is found, but this vortex dynamo effect is
suppressed after the self-generation of the magnetic field.
Introduction
Magnetic fields of planets and stars are almost axisymmetric. Therefore,
one may think that the problem of their generation due to the hydromagnetic
dynamo has to be studied in the axisymmetric formulation. But Cowling1
proved that the axisymmetric dynamo is impossible. His theorem condition
concerns the existence of closed magnetic lines. Cowling emphasized that his
proof is not applied to cases with unclosed magnetic lines.2 Later, Braginsky3
proved another version of the antidynamo theorem using the requirement that
magnetic induction decays at infinity as the third power of the inverse distance or
rapidly.
Here we study a few problems in which both of the conditions are not
satisfied and the axisymmetric dynamo takes place. In all of the cases there is
conical similarity. This means that the velocity field and self-excited magnetic
induction are inversely proportional to the distance from the origin. The Navier-

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Lie.


All rights reserved.
*Head, Modeling Laboratory, Institute of Thermodynamics,
tSenior Researcher, Modeling Laboratory, Institute of Thermodynamics.
87
88 M. A. GOLDSHTICKANDV. N. SHTERN

Stokes equations and the MHD equations admit such a class of exact self-similar
solutions. This class contains the well-known Landau5 and Squire6 analytic
solutions for jetlike flows and some MHD problems.7
The conical solutions have a number of features that seem to be
paradoxical.8 Some of them, such as a singularity appearance at finite Reynolds
numbers and a variety of rather unusual bifurcations, are demonstrated in this
paper. Three flow patterns are considered in which the magnetic and vortex
dynamo are found.
The first example is the plane flow induced by a point sink of fluid.
Bifurcation of induction occurs when the magnetic Reynolds number achieves 1.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

In the MHD regime the flow rate is locked despite the pressure gradient
increases.
This problem is solved analytically and serves as the limit case in the
following more complex examples.
The second and third problems are self-similar models of large-scale flows
near stars. Recently, strong jetlike flows have been observed in space near
young stars and galaxy cores.9'10 Their hydrodynamic model based on the Squire
solution has been proposed in Refs. 11 and 12. In such flows, favorable
velocity fields are formed for magnetic induction self-excitation.13 One more
mechanism of the large-scale motion driving is the thermogravitational
instability in a near-star medium. The conical self-similar solutions have been
found for this convection problem.8 Here it is shown that self-rotation
excitation (vortex dynamo) and the magnetic dynamo can be realized in this flow.

Sink Flow
Here we consider a plane problem in which, in the cylindrical coordinate
system (r, q>, z), velocity and induction vectors only have the components vr and
Vq,, with Bz being nonzero and not depending on z. In this case the
electromagnetic force is potential and may be joined with the pressure gradient in
the Navier-Stokes equations. We seek solutions of the MHD equations in the
following form:

vr = vr1U; vq) = vr1V; Bz = vr1H; p = poo-l/2 v 2 r 2 P

where U, V, H, and P are constants. Then it follows that

P = U2 + V2 + H2; (l+BtU)H = 0 (1)

Bt = v/vm, and v and vm are kinematic and magnetic viscosities. The motion is
considered to be driven by a pressure gradient, so that P is given and V is fixed
but U and H have to be found. There are two solutions of system 1: the pure
hydrodynamic (HD) and magnetohydrodynamic (MHD).

HD: H = 0; U = ±(P-V 2 ) 1/2

MHD: U = -Bt-1; H = ±(P-P*)1/2; P* = V2 + Bf 2


AXISYMMETRIC DYNAMO 89
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 1 Bifurcation of the MHD regime in a plane sink flow.


Flow rate U and induction H vs. pressure gradient P.
V = 0, Bt = 2-1/2.

The negative U corresponds to the sink, and the positive U corresponds to


the source. Solutions HD and MHD merge at P = P*. When P increases to
achieve P*, a supercritical pitchfork bifurcation takes place (see H vs P in
Fig. 1), resulting in the appearance of two new self-similar solutions. Because
the bifurcation is supercritical, we suppose that new MHD regimes are stable and
the initial HD solution becomes unstable. This metamorphosis may be
interpreted as the dynamo effect. At P < P* potential energy relating to the
pressure gradient is transformed into kinetic energy of the fluid motion. But at P
> P* the flow rate is locked and remains constant as P increases. Now the
energy of the pressure is transformed into the energy of the magnetic field.
The physical mechanism of this effect seems to be rather clear. If some
initial perturbation of axial magnetic field appears, then the convergent motion
of the conducting medium induces vortex currents, and their magnetic field
amplifies the initial induction. If the magnetic Reynolds number Rm = -BtU<l,
then this effect is suppressed by diffusion and dissipation. But at Rm > 1 the
steady MHD regime develops in which a growth of pressure gradient leads to an
increase of magnetic induction without any increase of the flow rate. We
emphasize that this effect takes place in the special self-similar class of exact
solutions. It is shown in the following that in this conical class the bifurcation
of the magnetic field is typical when a convergent motion of conducting fluid
occurs.

Squire Jet

Conically Self-Similar Class


Here we consider velocity and magnetic induction fields having in the
spherical coordinate system (R, 0, cp) the following representation:

v =- v0 = -(r sin9)-1vy(x); v<p = (rsin0)-1vr(x)

Br = - ^); B^ = B(rsin0)-1L(x)
90 M.A. GOLDSHTICKANDV. N. SHTERN

where x = cos 0, the prime denotes differentiation with respect to x, and B is a


constant to be determined. Substituting into the MHD equations and carrying
out some simple manipulations, we obtain the following system of equations:
(1-x2)/ + 2xy - y2/2 = F-SO2/2 (2)

(l-x2)0" = Bt (yO'-y'O) (3)

(l-x2)r" = yr-S<WL f (4)


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(l-x2)Lu = Bt[yL'-0r + 2(ytL-O'O + 2x(yL-OO/(l-x2)] (5)


•- i i '
(l-x2)F" = 2rrf - 2SLL! (6)

where S is a dimensionless criterion characterizing a value of the self-generated


induction. At the plane x = 0 the vortex-sink motion should be given:
y'(0) = Re;

together with the impermeability and symmetry conditions,

y(0) = <D'(0) = L(0) = 0

Regularity requirements for the velocity and induction at the symmetry axis
yield

and it follows from Eq. (2) that F(l) = 0.


Linear Dynamo Problem
First we consider the simplest case, F0 = 0. Then the problem admits
F = L = O = 0, and we have the Squire6 solution:

y = Re(l-x)(x cotfo InCl+x)]-!^}'1; % = 1/2 (2RC-1)1/2 (7)


The linear dynamo problem concerns finding the values of Bt or Re at which
Eq. (3) together with the conditions O(l) = O'(0) = 0 have a nontrivial solution.
Let Re be fixed and Bt be varied. Because the problem is linear and uniform, the
condition O'(l) = -1 may be added. Then at Bt = 0 the solution of the initial-
value problem, provided by integration from x=l to x=0, is <D=l-x, i.e.,
<£'(())= -1. In contrast, with Bt —> <*> we have the "frozen" field:

O = -y(x)/y'(l); *'(0) = -y'(0)/y'(D


The derivative of solution (7) has opposite signs at x = 0 and x = 1; thus,
<I>1(0) is positive. Because O'(0) depends on Bt continuously, a value of Bt
exists at which O'(0) = 0. Numerical calculations confirm that such a value
exists and is unique (curve 1 in Fig. 2). The critical magnetic Reynolds number
AXISYMMETRIC DYNAMO 91
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 2 Regime map for the Squire jet of conducting fluid.


Magnetic Reynolds number vs. Batchelor number. H, hydro-
dynamic solution; M, MHD solution; L, laminar; T, turbulent;
S, stable; U, unstable. In region I only HLS exists: in region
II MLS and HLU exist; in region III MLS and HUT exist; in
region IV MTS and HTU exist; and in region V HTU exists.

R m = BtRe varies in the rather small range from 1.74 at Re = Re* = 7.67,
Bt* = 0.226 (point K) to 3.5 with Bt -> <*> (dashed line in Fig. 2).

Nonlinear Dynamo
To solve the nonlinear problem we must integrate Eqs. (2) and (3) using
F = (1-x)2 [Re+SO2(0)/2]. It is convenient to introduce the condition O(0) = 1;
then S = AIRe2, where the Alfven number Al is a kinetic/magnetic energy ratio
at the plane x=0. Calculations have shown (Fig. 3) that, when Re increases at a
fixed Bt, the supercritical pitchfork bifurcation occurs at a Re value according to
curve 1 of Fig. 2. It relates to a MHD regime appearance with arbitrary sign of
the axial projection of induction. The transformation of kinetic energy to
magnetic energy leads to suppression of the axial jet. If in the absence of the
self-induced magnetic field the axial velocity turns to infinity at Re = Re* 12
(ray 2, Fig. 2), then because of the self-excitation of induction, the axial velocity
begins to decrease. At Rm » 1, using the technique of asymptotic expansion
matching, the uniform approximation of the MHD solution has been found:

y = Bf 1 [ 1 - x - exp(-Rmx)]

(8)
= 1 - x - Rm-1 [1 - exp(-Rmx)]
92 M. A. GOLDSHTICK AND V. N. SHTERN

This means that potential distributions of the velocity and induction occur
far from the plane x = 0 and the current layer near the plane. In the limit we also
have Al = Bt. The physical mechanism of the dynamo is the same as the one
reported in the preceding section. For the divergent flow at Re < 0 in which the
motion direction is opposite, the dynamo effect does not happen.

Turbulent Dynamo
The found bifurcation of magnetic field takes place at Bt > Bt* = 0.226. At
smaller Bt the the initial hydrodynamic solution loses its existence at
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Re = Re* before this bifurcation occurs. For liquid metals and star plasma, Bt
has an order near 10~6; thus, the case of small Bt is important for applications.
When the near-axis jet is strong enough, it can become turbulent. Ignoring real
transient processes, here we shall use very simplified models to describe the
transition to turbulence and turbulent regimes. In submerged jets a region of
turbulent motion is usually bounded by a rather narrow cone. Because of
entrainment the turbulent jet serves as a sink for ambient fluid. This is why we
shall study the outer flow in a sink at the symmetry axis modeling the turbulent
axial jet: y(l) = q * 0. The singularity appearing in the laminar solution at
Re = Re* causes that q becomes 4. According to Schlichting,14 this is the limit
entrainment ability for the laminar jet. For the turbulent regime we suppose
that

y(l) = q > 4

Then from Eq. (2) it follows that F(l) = 2q-q2/2. A requirement of the axial
momentum flux boundedness yields another condition:

In the common case the circulation may not be zero at the axis for the outer
solution, but <£ and L have to be zero, and

)1^ + o(l-x); Y = qBt/2 < 1

L = L^l-x) - (2/q)r(l) ^(1-x)1-? + o(l-x)

Here o(l-x) denotes a function which tends to 0 faster than 1-x as x -^ 1. It


follows from Eq. (4) that F(l) = 0 ( ) The boundary conditions at the plane x = 0
remain the same.
Numerical solution of this problem gives a q value of 4 at ray 2, Fig. 2 (in
the case T = 0), grows together with Re and asymptotically q = 1.73 Re3/4. But
at curve 4, Fig. 2, the MHD solution appears due to the supercritical pitchfork
bifurcation. In the MHD regime q decreases and achieves a value of 4 at curve 3,
which is the left boundary of the laminar MHD solution existence region in Fig.
AXISYMMETRIC DYNAMO 93

2. Curves 4 and 3 have the following asymptotes when Bt -> 0:


Rm = 0.674 Bf1/3; Rm = 2.52 Bf1/2

According our model we have the following scenario when Re increases at a


fixed Bt < Bt*. At small Re the flow is laminar and purely hydrodynamic.
When we pass curve 2, the flow becomes turbulent, remaining purely
hydrodynamic. At curve 4 the magnetic induction is self-excited and the
generated MHD regime is turbulent. Then at curve 3 turbulence in the axial jet
is suppressed and the MHD regime becomes laminar and remains laminar with
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Re —> oo.

Swirling Jet
Swirling is a stabilizing factor for turbulence development and for the
magnetic dynamo. At a fixed Bt the critical Reynolds number, at which the
MHD bifurcation occurs, increases together with F0. If F0 « Re, then at
Re —> oo we have the same asymptotic relations (8), added to

L = r0Re-1[exp(-Rmx)-l*(Q]

where F*, 1*, the solutions of a relevant boundary-layer problem,14 are shown
in Fig. 4. The aximuthal magnetic field is uniformly small in the flow region.
The rotational motion is localized not only in the near-plane region but also in

_ -j- zr^-J- _-J- _ —I 0


o
Fig. 3 Alfven number Al and axial velocity Rei = -y'(l) vs.
Reynolds number at Bt - 0.25. Typical streamlines (solid lines)
and magnetic line (dashed lines) are shown in the inner sketch.
94 M. A. GOLDSHTICK AND V. N. SHTERN

0,5 -
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 4 Near-axis distributions of circulation F* and aximuthal


induction 1* for a swirling MHD flow at Re » 1, F0 « Re.

the near-axis zone. This means that in the MHD regime the strong collimated
and swirling jets are formed like jets observed in space near young stars and
galaxy cores.
Also, the swirling leads to another interesting effect. It has been found that
at a high enough swirling the supercritical bifurcation of the MHD regime
metamorphoses into the subcritical one. As a result, the hysteresis transition
takes place between the hydrodynamic and MHD regimes in high swirling flow
(see Fig. 5). More detailed results on the magnetic dynamo in the Squire jetlike
flow and its generalization on the swirling motion have been reported in Ref. 13.

Convection Near a Star


Convection Onset
Another example of the axisymmetric hydromagnetic dynamo relates to the
problem of thermogravitational convection near a star. The pure convective
problem has been studied earlier.8 A star is regarded as a point source of heat and
gravity. Density p, gravity acceleration g and temperature T have the following
representation.
p/poo = 1 - P(T - Too); g = or'2; T = Too + yd(x)rl
where the constants a and y characterize the star mass and heat flux,
respectively; p is a coefficient of heat expansion; and the subscript <» corresponds
to ambient medium characteristics. Using the representation, the heat-transfer
AXISYMMETRIC DYNAMO 95
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 5 Metamorphosis of the supercritical bifurcation into a


subcritical one with circulation T 0 increase. Numbers denote
r 0 = 1, 5, 10.

equation may be reduced to

(1 - x2)#' = Pryd (9)

and Eq. (6) is modified to

(l-x2)FM = 2ITf - 2SLL1 + Rayfl (10)

where Pr is the usual Prandtl number and the Rayleigh number Ra = Prapy/v2.
Equations (9) and (10) must be added by equations (2-5) and the regularity
requirements:
y (±1) = F(±l) = *(±l) = L(±l) = F(±l) = 0
1
Normalizing condition J $(1 - Pry') dx = 2 gives the relation between
-1
heat flux Q, thermal conductivity X, and j: Q = :
96 M. A. GOLDSHTICK AND V. N. SHTERN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 6 Amplitude of near-star convection A vs. Rayleigh


number. Pr = 2. Typical streamlines at A > 0 and A < 0 are
shown in the inner sketches.

There is the trivial solution y = r = L = O = F = 0, # = 1


corresponding to a hydrostatic equilibrium and heat-conduction regime. Because
of the unstable stratification, a convective motion appears when Ra approaches
the critical value Ra*. The convection onset problem is solved analytically, and

Ra*(n) = n(n+l)(n+2)(n+3)

where n+1 is the number of convective cells. The minimum critical value is
Ra* = 24 at n=l, and the eigenfunction is y = Ax(l-x2). A weak nonlinear
analysis yields the relation between convection amplitude A and the Rayleigh
number:
Ra = 24 - 4A(3+2Pr)/7+0(A2)

We see that the bifurcation is two-sided. This means that the hysteresis
phenomenon and bistability occur (Fig. 6). With Pr —» 0 the parameters of the
minimum point of function Ra(A) tend to Ra -> 11.2, A -> °o.

Laminar Dynamo
The linear magnetic dynamo problem concerns finding a relation between Bt
and Ra at which a nontrivial solution of Eq. (3) exists. This formulation is
similar to one in the subsection on the linear dynamo problem, but now we have
the Prandtl number as an additional free parameter. The results calculated for Pr
= 1 are shown in Fig. 7. The upper branch of curve Bt*(Ra) corresponds to the
dashed part of curve A(Ra) in Fig. 6 and Bt* -> °o with Ra —» 24. The lower
AXISYMMETRIC DYNAMO 97
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

O
10

Fig. 7 Linear dynamo at Pr = 1. Critical Batchelor number Bt*


and axial velocity Re x vs. Rayleigh number.

branch Bt*(Ra) corresponds to the upper solid part of curve A (Ra). In the
downward convective regimes [see the lower sketch and lower solid part of curve
A(Re) in Fig. 6], the bifurcation of the MHD solution is absent.
The asymptotic regime at Ra -> °° for the upward convection is found
analytically. Near the symmetry axis a strong bipolar jet is formed that carries
out the whole heat flux. In the external region y=4x and $ = 0 at Pr > 1/2.
This outer solution corresponds to the uniform sink of fluid placing at the axis,
which has been considered in the section on sink flow (here U= -4). Equation (3)
also has the analytical solution $ = (1-x2)172, Bt* = 1/4, which coincides with
the MHD solution in the section on sink flow.
Returning to the dependence Bt*(Ra), we conclude that it is a monotonically
decreasing function, and 1/4 < Bt* < Bt^ at Ra^ < Ra < °o. The values Bt^
and Ra^ depend on Pr. At Pr = 1 we found Bt^ = 5.56, Ra^ = 23.4 (point 0
in Fig. 7).
To solve the nonlinear magnetic dynamo problem we must integrate Eqs.
(1), (2), (8), and (9). Numerical results for Pr = Bt = 1 are shown in Fig. 8. As
in the Squire flow the generation of magnetic field leads to a weakening of the
near-axis jet (compare curves 1 and 2 in Fig. 8). The potential energy of
buoyancy transforms mainly in magnetic energy, and kinetic energy does not
increase; i.e., the locking effect takes place like the effect reported in the section
on sink flow.

Turbulent Magnetic Dynamo


At a small enough Pr and Bt the convective motion that develops as a result
of a finite-amplitude instability of the hydrostatic equilibrium [the upper branch
98 M. A. GOLDSHTICK AND V. N. SHTERN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 8 Nonlinear dynamo at Pr = Bt = 1. Axial Alfven number


A l j and axial velocity Rej in hydrodynamic (1) and MHD (2)
solutions vs. Rayleigh number.

of A(Ra) in Fig. 6] may include a turbulent near-axis jet. Heat transfer near a
star is provided mainly by radiation and depends on ambient medium motion
rather weakly. Such a situation may be modeled with the aid of a small effective
Pr value. Here we consider the limit case Pr = 0. Then $ = 1. To calculate a
stable large-amplitude regime, we use a method of continuous variation of
parameters, starting with the bifurcation point Ra = 24, A = 0, and then
increasing A. At finite amplitudes it is convenient to introduce the new
parameter Re0 = y'(0) instead of A. When Re0 increases, the axial jet
strengthens, and at Re0 = 4.62, Ra = 11.2 (point T in Fig. 9), the axial velocity
turns to infinity.
To model a supposed turbulent regime we use the same narrow jet
approximation as that in the subsection on the turbulent dynamo. Numerical
results for such a turbulent flow correspond to curve 2 in Fig. 9. In contrast to
the laminar solution (curve 1), now Re0 increases together with Ra. The
character of dependence Reo (Ra) seems to point out that the laminar solution is
unstable and the turbulent one is stable. In the turbulent regime q increases from
q = 4 at point T.
As stated earlier, the linear magnetic dynamo problem concerns finding a
Bt* at which Eq. (3) has a nontrivial solution. A characteristic, that is more
conservative than Bt*, is the magnetic Reynolds number Rm0 = Bt*Reo. This
number varies from 2.23 at Re0 = 0 to 1.1 at point T in Fig. 9 and remains
practically constant at R m0 = 1.1 at curve 2.
AXISYMMETRIC DYNAMO 99
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 9 Sequence of bifurcations at Pr = 0. Equatorial velocity


R e 0 v s . Rayleigh number. Curves 2 and 3 correspond to
hydrodynamic and MHD solutions, respectively, in the model of
an infinitely narrow axial turbulent jet. The flow and induction
patterns in the sketch correspond to the laminar MHD regime,
curve 4. Bt = 0.2.

In the nonlinear case we must solve Eqs. (2), (3) and (10) together. Results
for Bt = 0.2 are shown by curve 3 in Fig. 9. After the MHD regime bifurcation
at point M (Re0 = 5.5, Ra = 13), q decreases and turns to 4 at point L (Re0 =
6.8, Ra = 78.4), where the relamination happens. To the right of the L the
laminar MHD regime exists at arbitrary large Ra. If we would not apply our
model of the turbulent regimes, then this laminar solution is separated from
other laminar solutions.
We have supposed for the turbulent solutions that swirling is absent:
F s 0. But if we ignore the regulatory requirements at the axis, then in the pure
hydrodynamic case, Eq. (4) permits the solution T = const * 0. This constant
cannot be found for the model with an infinitely narrow turbulent core. To study
the possibility of the self-swirling generation, we must consider a more detailed
model of the turbulent convective flow.

Turbulent Vortex Dynamo


Here we use a very simplified model of turbulent motion with the aid of the
concept of eddy viscosity having a step-function distribution.11 Based on the
Schlicthting data,14 we suppose that a turbulent region is bounded by the cone
x > xt = 0.976. Inside the cone viscosity is effective (vj, and outside the cone
viscosity is molecular (J), and TV = vt /v » 1. The turbulent Reynolds number
rv r (l)/v t should be constant and equal 460.5 (Ref. 14). If -y'(l) is less than
460.5, then the flow is considered to be laminar. In the opposite case the
100 M. A. GOLDSHTICK AND V. N. SHTERN

parameter TV is chosen so that -y'(l)/Tv = 460.5. The magnetic viscosity vm is


not modified here in the turbulent regime. At x = xt, conditions of continuity
for velocity and momentum flux are satisfied. The parameter TV is found in the
solution process as was q in the preceding section.
The results of our calculations are shown in Fig. 10. Curves 1 and 4 are the
same as the curves in Fig. 9. They correspond to unstable and stable laminar
solutions accordingly. At point T1 (Re0 = 4.57, Ra = 11.9), the turbulent
regime appears without any rotation and magnetic field. It corresponds to curve
2' in Fig. 10. Since this solution seems to be stable, the part of curve T is
shown by a solid line. At point S (Re0= 5.65, Ra = 13.8, TV = 1.37), a
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

supercritical pitchfork bifurcation occurs, and two new regimes appear that have
opposite swirling. These solutions correspond to curve 5. When Ra increases,
the axial velocity remains near constant, but the rotational motion energy grows.
The magnetic dynamo problem has been calculated for Bt = 0.2. At point
M" (Reo = 6.43, Ra = 27, TV = 1.36), a bifurcation of the MHD solutions takes
place, and all components of the induction vector are nonzero. However, this
bifurcation is subcritical; thus, the MHD solutions seem to be unstable (which
is why curve 6 and curve 5 at Ra > 27 are shown by dashed lines).
If we ignore the bifurcation at point S and move along curve 2', then at
point M' (Re0 = 6.03, Ra = 14.5, TV = 1.48) a bifurcation of a nonswirling
MHD solution occurs that corresponds to curve 3'. A vicinity of point M' is
enlarged in the right lower part of Fig. 10. New bifurcation occurs at point S f
(Re0 = 6, Ra = 14.6, TV = 1.48) where curves 31 and 6 intersect.
To study stability features of the solutions a special investigation is needed.
Our conjectures on solution stability (which are shown by solid and dashed lines)

/G -

«'/-r^
v
5O Tta 5QO

Fig. 10 Vortex and magnetic dynamo in the model with eddy


viscosity in a near-axis region. Bt = 0.2. For more details see
the subsection on the turbulent vortex dynamo.
AXISYMMETRIC DYNAMO 101

are based on the bifurcation types. In some range of Ra three locally stable
regimes may coexist corresponding to curves 3' and 5 and the equilibrium
(Re0 = 0). But if we take into account different directions of rotation and
induction, then the stable solution number is five. At point L' (Re0 = 8, Ra =
244) curve 3f intersects curve 4. This means that the self-induced magnetic field
suppresses turbulence and at Ra > 224 the MHD regime becomes laminar.
Conclusion
We have shown that the axisymmetric hydromagnetic dynamo is possible.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

It is a result that is in contrast to the well-known antidynamo theorems of


Cowling and Braginsky. But there is actually not any contradiction because the
theorem conditions are not satisfied in our examples. Two features are
significant for the exposed dynamo effect. The first one is the conical self-
similarity. It occurs because magnetic lines are not closed and induction decay at
infinity is weak. The second condition is a convergent character of flow. This
leads to cumulation of kinetic and magnetic energy near the symmetry axis. As
a result, a strong near-axis jet is formed and a number of rather unusual
bifurcations take place, leading to generation of new fields: rotation and
induction. The self-excitation of rotational motion is needed in the additional
condition, i.e., increasing viscosity to the axis. This increase may be interpreted
as a result of turbulence. But the self-generation of magnetic induction is also
possible in the laminar regime.
The physical reason of the magnetic dynamo is a positive feedback. If we
have an initial axial induction disturbance, then the convergent motion of
conducting medium induces vortex electric currents and their magnetic field
amplifies the initial induction. On the other hand, the conical symmetry is a
serious shortage for exposed axisymmetric hydromagnetic dynamo. Indeed, such
a dynamo is found in the rather wide class of exact solutions of the MHD
equations. But real problems do not exactly possess this self-similarity, and this
may change the result.

References
1
Cowling, T. G., "The Magnetic Field of Sunspots," Monthly Notices of
the Royal Astronomical Society, Vol. 94, 1934, pp. 39-48.
2 Cowling, T. G., Magnetohvdrodynamics. Interscience, New York, 1957.
3 Braginsky, S. I. "Self-Excitation of Magnetic Field During the Motion of
a Highly Conducting Fluid," Jornal of Experimental and Theoretical Physics.
(Sov.).. Vol. 48, 1964, pp. 1084-1098.
4 Moffat, H. K., Magnetic Field Generation in Electrically Conducting
Fluids. Cambridge Univ. Press, 1978, Chap. 6.
5 Landau, L. D., "On an Exact Solution of the Navier-Stokes Equations."
Soviet Physics Doklady, Vol. 43, No. 7, 1944, pp. 299-301.
6 Squire, B., Some Viscous Flow Problems. 1. Jet Emerging from a Hole
in a Plane Wall." Philosophical Magazine. Vol. 43, No. 343, 1952, pp. 942-
945.
102 M. A. GOLDSHTICK AND V. N. SHTERN

7 Bojarevich, V., Freiberg, J. A., Shiliva, E. I. and Shcherbinin, E. V.,


Electrically Induced Vortical Flows. Kluwer, Dordrecht, 1989.
8 Goldshtik, M. A., Shtern, V. N., and Yavorsky, N. L, Viscous Flows
with Paradoxical Features, Nauka, Novosibirsk, USSR, 1989 (in Russian).
9 Lada, C. J., "Cold Outflows, Energetic Winds, and Enigmatic Jets Around
Young Stellar Objects," Annual Review of Astronomy and Astrophysics. Vol.
13, 1985, pp. 267-317.
10 Konigl, A., "Stellar and Galactic Jets: Theoretical Issues," Canadian
Journal of Phvsics. Vol. 64, No. 3, 1986, pp. 362-368.
11 Goldshtik, M. A. and Shtern, V. N., "Conical Flows of Fluids with
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Variable Viscosity," Royal Society of London Series A. Vol. A419, 1988, pp.
91-106.
12 Goldshtik, M. A. and Shtern, V. N., "On a Mechanism of Astrophysical
Jets," Soviet Physics-Doklady, Vol. 304, No. 5, 1989, pp. 1069-1072.
13 Goldshtik, M. A. and Shtern, V. N. "Self-Similar Hydromagnetic
Dynamo," Jornal of Experimental and Theoretical Physics. (Sov.)« Vol. 96, No.
5, 1989, pp. 1728-1743.
14 Schlichting, H., Grenzschicht Theorie. Braun, Karlsruhe, Germany,
1965.
Bifurcations of Self-Oscillating and Almost Periodical
Regimes in an Azimuthal MHD Jet
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

K. Sergeev* and V. N. Shternt


Siberian Branch of the Academy of Sciences, 630090 Novosibirsk, Siberia

Abstract
Results concerning the analysis of nonlinear stability and subsequent
bifurcations of an azimuthal MHD jet are presented. The analysis is performed
in the framework of a low-dimensional dynamic system for the amplitudes of the
most dangerous perturbations and distortions of the initial average flow, which
are generated by oscillations. This system is obtained by the Galerkin method.
The structures of the inner and other vortex streets obtained by linear theory as
well as the functions obtained for additions to the initial average motion as a
result of solving the Reynolds equations are used as the basic modes of the
Galerkin approach. The number of vortices in the streets is chosen to be equal
to and corresponding to the most dangerous perturbations. As a result, the
evolutionary system of ordinary nonlinear differential equations of the eighth
order is obtained. In the framework of this system it is found that a supercritical
bifurcation of a self-excited oscillatory regime, corresponding to a number of
vortices rotating with some angular velocity, occurs. When the Reynolds
number exceeds the first critical value by approximately 1.75 times, the almost
periodical regime arises. The structure of such a regime corresponds to a vortex
"breathing." With the further increase of the Reynolds number in the frame of
the model considered, the flow pattern does not change qualitatively.

Introduction
The Taylor flow between rotating cylinders or spheres and the Rayleigh-
Benard convection are the most popular topics for the investigation of transition
dynamics from a laminar motion to a turbulent motion in bounded regions. In
this paper results are presented for the numerical investigation of the flow
occurring in the experimental apparatus suggested by Levin,1 which will be

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Graduate Student, Modeling Laboratory, Institute of Thermodynamics.
fSenior Researcher, Modeling Laboratory, Institute of Thermodynamics.
103
104 K. SERGEEVANDV. N. SHTERN

called an azimuthal magnetohydrodynamic jet (MHD jet). The MHD jet is


shown schematically in Fig. 1 for large magnetic fields. In Fig. 1,1 denotes the
horizontal nonconducting plates; 2 denotes the two coaxial electrodes; h is the
distance between plates; I? is the flux density of uniform magnetic field applied
in a direction perpendicular to plates (it has a z component only); I is the electric
current; and r, 9 z are the cylindrical coordinates. An outer cylindrical wall is
not pictured, since there is no fluid motion in the vicinity of the wall and
consequently the latter cannot influence the flow. The MHD jet is an attractive
topic for studying the transition to turbulence for the following reasons.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1) The transition to turbulence is delayed by the presence of the magnetic


field. This occurs because of the appearance of an additional stabilizing factor
such a the joule dissipation, which increases the growth of the magnetic field in
three-dimensional disturbances to hinder their enlarging. A wide range of
transition parameters is a large stimulus for the investigation of the transition
picture in the magnetic field. The MHD jet is a closed flow, except for the
presence of disturbances brought from without.
2) It is of a practical interest to investigate the laminar-turbulent transition in
the MHD jet, the latter being an often used element of flows in various MHD
devices.

Fig. 1 Scheme of the MHD jet.


JET BIFURCATIONS 105

3) A two-dimensional turbulence is realized in the MHD jet so that one can be


simulated in laboratory conditions. The investigation of this turbulence is
topical because of its applied significance for geophysical and astrophysical
objects. For example, a picture arising after the loss of the MHD jet stability is
similar to the street of cyclones and anticyclones observed above Antarctica.
Thus, this global atmospheric phenomenon can be modeled by the MHD jet.
4) In a number of cases the results of the MHD jet investigation can be
applied to such wide class of flows as usual jets. By virtue of closing of the
MHD jet the phenomenon of unparallelism is eliminated. This is an important
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

problem in the theoretical analysis of jets.

The experimental investigation of a laminar regime of the MHD jet has


been made by Klyukin et al.2 It was found that, when the magnetic field is
increased, the flow becomes homogeneous in the direction of the magnetic field,
with the exception of the thin Hartmann layers near walls which are normal to
the magnetic field. Practically all of the flow is concentrated in the region over
and between the electrodes. The free shear layers are formed at the inner and
outer jet boundaries. The measurements have shown that the secondary radial
flows are smaller at least by a factor of 10 of the main flow.
The dependence of the critical Reynolds number on the Hartmann number
has been obtained by Klyukin et al.,3 and the structure of the disturbances in the
supercritical regime has also been investigated by those authors. The values of
the two-point correlation coefficient of the azimuthal and radial components of
the velocity disturbances are positive, and the former are close to 1 in the greater
part of the MHD jet height. This indicates the high homogeneity of the flow
along the magnetic field. The visualization of a free liquid surface with the help
of Lehnert's method4 shows that in the motion region the two closed vortex
streets appear when the stability boundary is crossed. Their axes are directed
along the magnetic field.
Klyukin and Kolesnikov5 and Kolesnikov and Polyakov6 have investigated
experimentally the transition dynamics from the laminar regime of a MHD jet to
a turbulent one. They found that, with the inrcease in the Reynolds number
while the value of the magnetic field is held constant, either a bifurcation to
another oscillating regime is observed (with the frequency decrease) with
subsequent quasiperiodic regimes appearing, or the two-frequency regime
immediately appeared, with subsequent growth of a number of frequencies.
Turbulence has been observed after four bifurcations. It develops with the
conservation of the high anisotropy degree of the disturbances in the direction of
the magnetic field.
Klyukin and Levin7 have shown that the MHD jet stability depends
strongly on its extent along the magnetic field. Their experiments show the
stability depends on the modified Hartmann number based on the extention
length. Klyukin and Levin have mentioned that the Hartmann boundary layers
strongly influence the flow stability because the main dissipation of disturbance
106 K. SERGEEVAND V. N. SHTERN

energy occurs in the layers. Levin and Shtern8 have taken this influence into
account by averaging the equations of stability along the z coordinate.
The Hartmann velocity profile has been used for both the base flow and
disturbances. Then the averaging with respect to z transforms the term
containing the second derivative of velocity into a product of the velocity and a
factor. The factor is the modified Hartmann number.
An analysis of the linear stability of the azimuthal MHD jet9 has been
shown that it has two types of the disturbances. The first corresponds to an
inner vortex street, and the second corresponds to an outer one. The later is more
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

dangerous. The distribution of the intensities of the velocity disturbances


coincides qualitatively with the experimental data.6 The critical Reynolds
numbers have been found to be about 2.3 times smaller than experimental ones.
This is probably explained by that the dependence on z of the MHD jet velocity
may be different from Hartmann's one in the near-electrode regions. The
difference influences essentially the critical Reynolds number value because,
namely, in these regions the generation of the disturbances takes place.

Formulation
We denote inner and outer edges of the inner electrode TI and r2,
respectively, and the inner and outer edges of the outer electrode r3 and r4
respectively. Let b0 = r4-r! be a scale of the length. In accordance with the
parameters of the experimental apparatus,2'3 we use ^=3, r2=3.2, r3=3.8 and
r4=4.
The investigation has been done in the noninductive approximation. In
this case the magnetic Reynolds number is Rem=p,0|iaV0b0 « 1. Here ju^is a
permeability of vacuum, |i is the magnetic permeability, a is the conductivity,
and V0 is the scale of the velocity.
At Rem « 1 the hydrodynamic part of the MHD equations is separated
from the electrodynamic one. Thus, the MHD jet is described by the Navier-
Stokes equations with the electromagnetic force given in the righthand side:

and by the continuity equation:

div V = 0
where
^ is a velocity field, t is time, p is pressure Re = VJ) Jv is the Reynolds
number, v is the kinematic viscosity ,Ha=Bzb0Va/(pv) is the Hartmann number,
p is a density, and j is an electric current density.
The external magnetic field is considered to be large enough. Then the
laminar flow does not depend on z everywhere, and only thin Hartmann's
boundary layers are formed near horizontal plates. The large-scale velocity
disturbances seem to have the same feature. Stephenson's10 results attest to
this.
JET BIFURCATIONS 107

Thus, the velocity field has the following form:

V(r,(p,z;t) = \fc,(p;t) {l-exp[-Ha(h/2-lzl)]}


(2)
- h / 2 < z < h / 2 ; Ha(h/2)

For the laminar flow the velocity has only the cp component, with the
radial flows neglected, and does not depend on cp:
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

V<p(r,z) = U(r) { 1 - exp[-Ha(h/2-lzl)]}

The laminar velocity profile U has been obtained analytically by the authors.9 It
corresponds qualitatively to the experimental data.2 The velocity of the disturbed
flow (in the process of the transition to turbulence) has r and cp components. In
the presence of the strong magnetic field the disturbance vorticity becomes
parallel to the magnetic field.
The nonlinear equations for the disturbances8 are two dimensional.
Therefore, a stream function \\f can be introduced:
v
Vr
_ . I d_i Vm _ 3JSL (3)
~ r 3cp' V ( P ~ 3r ' V>
where
V(r,<p,z;t) = V(r,(p;0{ l-exp[-Ha(h/2)-lzl)]}

according to Eq. (2). Then Eq. (1) is satisfied automatically.


A viscous term in the nonlinear stability equations is transformed after
averaging with respect to z. It is linear thus, we can write an equation for the
stream function describing the MHD jet:

_ _>
J (curl jx B) z dz (4)
h/2
where

and M = 2Ha/h is the modified Hartmann number.

The stream function is examined in the following form:

V(r,cp;t) = \|/0(r) + \|/g(r;t) + v'(r,<P;t) (5)


108 K. SERGEEVAND V. N. SHTERN

where \\f0 is the stream function of the laminar flow with the velocity profile U,
yg is an axisymmetric part of the stream function of the pulsations appearing
due to the Reynolds stresses, and \|/' is the nonaxisymmetric part of the stream
function of the pulsations.
Substituting Eq. (5) into Eq. (4) and subtracting the laminar flow equation
in view of Eq.,(3), we obtain
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

!«*»
= (ArM)A2Ug)

and the Reynolds equation for Ug is

= (ArM)A2Ug) (7)

where the bars denote averaging with respect to 9.

Subtracting Eq. (7) from Eq. (6X we obtain an equation for y':
i [(U+Ug ) (Al¥> f* § [(A2(U+Ug))

(8)

The space structure of the disturbances with respect to r and q> in a first
approximation is defined by the results of the linear stability theory. Time
evolution of the disturbances is described by complex amplitude functions Ak(t).
In view of this one can approximate \yf as

cp' = A^Ov^r) exp (im^) + A2(t)v2W ^XP (im2q>)+c.c. (9)

where the subscripts 1 and 2 denote the values corresponding to the outer and
inner vortex street, c.c. denotes the complex conjugate terms, and x^ and \j/2
are the eigenf unctions found in the linear theory. Here we choose mj = m2 = m
to indicate the azimuthal wave number at the most dangerous disturbance.
JET BIFURCATIONS 109

Evolutionary Equations Systems


Substituting Eq. (9) into the stationary form of Eq. (7) and executing the
operation of averaging, we obtain
Ld
rdr 1
1 d, 1 TTT M

where the asterisk denotes the complex conjugation, and the following
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

designations are introduced:

g
-
dr2 r dr r2
-»Un = j O T , ~ 9

; F02 = v 2 N 2 + v2N2; F03 = v^N 2 + v2N*;

~; Uk =
^ ; k=1'2

We integrate Eq. (10) with respect to r. The integration constant is 0


because of the uniform conditions at r = 0. The averaged and pulsating velocities
are 0 at this point. Then, representing Ug in the following form:

Ug = A01(t) Ugl(r) + Ao2(t) Ug2(r) + Ao3(t)Ug3(r) + A0*3 U0*3 (11)

from Eq. (10) we obtain the following equation system:

LU g j -MU g j = F0j, j = l,3

Each of these equations is a boundary problem with the following


conditions:

The problem is solved by a nonuniform differential factorization method.11 The


resulting functions Ugl and Ug2 are real, and Ug3 is complex. According to this
the amplitudes A0i and A02 are real, and the amplitude Ao3 is complex.

In view of Eq. (10) the Reynolds equation [Eq. (7)] becomes

- 4- FoiAiA +F02A2A + F 0 3 AA 2 + F * A i A = LU
" U
110 K. SERGEEVANDV. N. SHTERN

Applying the Galerkin method to the last equation with the basic functions
Ugi, Ug2, and Ug3 from expansion (11), and to Eq. (8) with basic modes
xJlTj exp(im19) and $2 exp(im29) from expansion (9), one can obtain an
evolutionary equation system. Having made the following redesignations (Ai =
xl + ix2, A2 = x3 -i- ix4, AOI = x5, Ao2 = x6, A03 = x7 + ixg), also by means of
introducing the amplitude for the laminar velocity profile x0 = 1, the system can
be written as
dx 1
•^ = ^ aijXj - bijkXjXk, ij,k = 0.8 (12)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where a^ and b^ are the coefficients determined by the Galerkin method.

Equation (12) represents a finite-dimensional model of the Navier-Stokes


equations for the case of small Re - Re* values. In this model the vortex streets
interact by means of the average velocity profile. The dynamic system obtained
is of the hydrodynamic type.12 In particular, the condition

bijk + bjki + bjdj = 0


for a tensor b^ is executed.
For convenience the following transformation is used:

exp(-icot), A2 = A2(t) exp (-icot)

which refers to a system of coordinates rotating with an angular velocity co/m.


Then the evolutionary system has two parameters: co and Re. The Reynolds
number is chosen as a free parameter. Let us induce the representation A! = I AI I
exp [i(Yi + <p0)], A2 = IA2I exp [i(y2 + cp0)]. The arbitrary constant phase cp0 can
be excluded further by application of the variables IAj I, IA 2 I , and y = Yi - Y2-
In this case the order of the evolutionary system is reduced by 1 and another
equation is obtained for co.

Supercritical Regimes
A numerical analysis of the MHD jet bifurcations has been performed at
Ha = 421, h = 7, and m = 9. The bifurcation diagram is shown in Fig. 2, where
e is the intensity of the velocity pulsations that is the ratio modulus of the
maximal velocity pulsation value to the maximal average velocity value. At
Re* = 252 the laminar MHD jet becomes unstable, and the stable stationary
regime occurs. This is shown by the solid curves in Fig. 2.
An illustration of the MHD jet stream functions corresponding to
stationary regime at Re = 440 is presented in Fig. 3. For this case the
coordinates of the attracting fixed point in the phase space are AOI = 0.1396, Ao2
= 0.8796, Ao3r= -0.06864, AOSI = 0.3436, lAil = 0.01781, IA2I = 0.04471, and
Y = -1.768. Since the flow is periodic along q> with a period equal to 27i/m, an
JET BIFURCATIONS 111

2 OH

15-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

10-

5-
• 1 I
1
1

':
0- i
J.,... .(MM,,,,,,,.. *^
200 400 600 800 1 0 Re
00

Fig. 2 Bifurcation diagram for intensity of azimuthal (1) and radial


(2) velocity pulsations.

2 0 2.5 3.0 3.5 4.0 4.5 5.0


Fig. 3 Streamlines in self-oscillating regime.
112 K. SERGEEVANDV. N. SHTERN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

t = 7/2 t = 37/4

Fig. 4 Streamlines in almost periodical regime.

illustration of only one period is shown. The vortex couple can clearly be seen.
The intensity of the outer street vortex is considerably larger.
Except for the stationary solution, which branched at Re=252 and was
related to the outer vortex street,9 the system has another stationary solution
bifurcation at Re=313.5 relating to the inner vortex street. This stationary
regime appears unstable and remains unstable with an increase in Re. In Fig. 2
it corresponds to the broken lines.
At Re = 442.13 the stationary regime loses its stability, and a periodical
regime occurs. 'This bifurcation takes place at Re/Re* = 1.75. This conforms
to the experimental data. At this Re value the coordinates of the fixed point are
AOI = 0.1426, Ao2 = 0.8848, A03r = -0.07006, Ao3i = 0.3483, IAJ = 0.01796,
IA2I = 0.04473, and y = -1.769. In Fig. 2 the stationary regime losing the
stability corresponds to the dashed lines. The curves corresponding to the stable
JET BIFURCATIONS 113

1.04-1

1.03-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1.02-

1.01-

1.00-
200 400 600 800 1000 Re

Fig. 5 Dependence of frequency on Reynolds number.

periodical regime are shown by the dotted lines. Discovering periodical solution
of the evolutionary system and investigating its stability have been performed by
use of the Poincare map.13
In Fig. 4 the streamlines are shown at Re=500, with T = 12.061 being the
period. The coordinates of a point on^he periodic orb# are Aol = 0.3428, A^2 =
1.104, A03r = 0.433, Ao3i = 0.643, IAJ = 0.0605, IA2I = 0.0424, and y = -
1.98. In this figure one can see as the vortices are "breathing," changing its
intensity and sizes during the period. Also, the vortices rotate with different
angular velocities.
The dependence of frequency on the Reynolds number is shown in Fig. 5.
Here the stable stationary regime is shown by the solid curve, and the dotted
curve corresponds to the stable periodical regime. _
The average velocity profile of the MHD jet U = U+<Ug> at the various
Reynolds numbers is plotted in Fig. 6, with the quantity within average bars
indicating averaging with respect to time. With the Reynolds number increase
the average velocity profile is disturbed drastically in the areas of the location of
the vortices.
The study of the stability of the almost periodical regime has been made up
to Re/Re* = 20, and a new regime has not been found. However, at Re = 1000
the average profile of the MHD jet (see Fig. 6) is strongly disturbed, and the
experiment6 shows that two additional bifurcations are observed, after which at
114 K. SERGEEV AND V. N. SHTERN

U/
1.0-

0.8-

0,6-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0.4-

0.2-

0.0
2.5 3.5 4.5 r
Fig. 6 Average velocity profiles for Re = 440 (———),
Re = 500 (—— —), and Re = 1000 (—•—•—).
The laminer profile (••—••) is shown for comparison.

Re/Re* = 3.5 the MHD jet becomes completely turbulent. Our model is too
simplified to describe this.
References
*Levin, V.B., "Free Rotating Layer of Conducting Fluid in Axial Magnetic
Field," Magnetohydrodynamics, Vol. 16, No. 1, 1980, pp. 86-92.
2
Klyukin, A.A., Kolesnikov, Y.B., and Levin, V.B., "Experimental Study
of Free Rotating Layer in Axial Magnetic Field. Part 1. Stable Flow,"
Magnetohydrodynamics, Vol. 16, No. 1, 1980, pp. 93-98.
3
Klyukin, A.A., Kolesnikov, Y.B., and Levin, V.B., "Experimental Study
of Free Rotating Layer in Axial Magnetic Field. Part 2. Stability Limits and
Perturbation Structure," Magnetohydrodynamics, Vol. 16, No. 1, 1980, pp.
140-143.
4
Lehnert, B., "An Instability of Laminar Flow of Mercury Caused by an
External Magnetic Field," Proceedings of the Royal Society of London, Series
A, Vol. 233, Dec. 1955, pp. 299-302.
5
Klyukin, A.A. and Kolesnikov Y.B., "Experimental Study of a Rotating
MHD Flow Stability," Matnetohydrodynamics, Vol. 16, No. 2, 1980, pp. 140-
142.
JET BIFURCATIONS 115

6
Kolesnikov, Y.B., and Polyakov, N.N., "Experimental Study of an
Axisymmetrical Shear Rotating Row in a Axial Uniform Magnetic Field. 1.
Averaged Flow and Intensity of Velocity Fluctuations,"
Magnetohydrodynamics, Vol. 19, No. 3, 1983, pp. 83-89.
7
Klyukin, A.A. and Levin, V.B., "Stability of Free Submerged Rotating
Layer of Conducting Fluid in Axial Magnetic Field," Fluid Dynamics, Vol. 19,
No. 5, 1984, pp. 166-173.
8
Levin, V.B., and Shtern, V.N., "Stability of Jet MHD Flow Between
Insulating Plates in Transversal Magnetic Field," Magnetohydrodynamics, Vol.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

21, No. 2, 1985, pp. 23-28.


9
Sergeev, K.A. and Shtern, V.N., "Stability of Azimuthal Jet Flow of
Conducting Fluid in Axial Magnetic Field," Magnetohydrodynamics, Vol. 23,
No. 3, 1987, pp. 39-46.
10
Stephenson, C.J., "Magnetohydrodynamic Flow Between Rotating
Coaxial Disks," Journal of Fluid Mechanics, Vol. 38, Sept. 1969, pp. 335-352.
H
Goldshtik, M.A. and Shtern, V.N., Hydrodynamic Stability and
Turbulence, Nauka, Novosibirsk, USSR, 1977, Chap. 1.
12
Gledzer, E.B., Dolzhansky, F.V., and Obukhov, A.M., "Systems of
Hydrodynamic Type and Their Application," Nauka, Moscow, 1981, Chap. 1.
13
Anishenko, V.S., "Stochastic Oscillations in Radiophysical Systems,"
Saratov University Publishing, Saratov, USSR, 1985, Chap. 4.
Bifurcations in MHD Flow Generated
by Electric Current Discharge
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

A. A. Petrunin* and V. N. Shternt


Siberian Branch of the Academy of Sciences, 630090 Novosibirsk, Siberia

Abstract
We consider a steady conical flow of viscous incompressible conducting
fluid driven by an electric current diverging from a point source placing on a
plane. Earlier it was found that the axial velocity becomes infinite at a finite
critical current value. Here it is shown that a bifurcation of a new solution with
a nonzero poloidal electromagnetic field and azimuthal rotation occurs. This
happens if the electrical conductivity of medium is high enough. The revealed
phenomenon of the axisymmetric MHD dynamo does not contradict Cowling's
"antidynamo" theorem because its conditions are not fulfilled. At a low
conductivity the paradox is solved by a simple model that takes into account the
possibility that the axial jet may become turbulent. In this case a self-swirling
of the jet flow due to a bifurcation is found. There is a possibility that such a
phenomenon does indeed take place in some real MHD processes. As the current
increases first the self-swirling develops and then the poloidal magnetic field
appears. The growth of the poloidal magnetic field and azimuthal rotation leads
to a relamination of the turbulent jet flow.

Introduction
Jetlike flows of conducting fluids are widespread in some high-current
industrial processes. Electrical arcs and electroslag welding are typical examples
where there are such flows. A convenient model for studying them is the flow
in the semi-infinite region of the conducting viscous incompressible fluid
induced by an electric current discharge emerging from a point on a plane.
Zhigulev* pointed out that the spherically symmetric current causes the motion of
fluid because of the rotational Lorentz force. The flow converges near the plane

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Graduate Student, Modeling Laboratory, Institute of Thermodynamics.
fSenior Researcher, Modeling Laboratory, Institute of Thermodynamics.
116
MHD FLOW BIFURCATIONS 117

and forms an axial jet outflow. This is described by conically symmetric


solutions of MHD equations and has been studied in Refs. 2-6. Sozou5 found an
effect similar to one studied earlier by Goldshtik7 for the Navier-Stokes equations
of the same self-similar class. When the current exceeds a certain critical value,
a regular solution ceases to exist. On the subcritical situation the axial velocity
begins to increase rapidly and turns to infinity at the current critical value.
Further attempts to overcome this paradox have not met with success.
Narain and Uberoi8 investigated the case of an electric current discharge
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

from a point source into a conical region. They have an increase in the critical
value S* (S is direct proportional to square of the electrical current value) with a
decrease in the cone angle. The effect of velocity on the electromagnetic field
also causes the current critical value to increase.9 Sozou and Pickering10
considered the electrovortex flow in a hemispherical bowl induced by a
symmetric discharge of an electric current from a point source at the center of the
fluid surface. Their computations indicate that, as the current grows, the axial
velocity increases drastically. This leads to a high-velocity gradient, and the
numerical algorithm fails. Bojarevich and Millere11 and Atthey12 encountered an
analogous difficulty when they sought a numerical solution for flow between
finite electrodes.
The development of an azimuthal rotation resulting from an external axial
magnetic field has been studied. In Refs. 13 and 14 it was shown that the
azimuthal rotation can slow down the meridional flow or even reverse it.
Millere et al.15 reported the energy transfer from meridional flow to the
azimuthal rotation. Furthermore, when the outer poloidal field tends to zero at
S = S* = const, the rotation remains nonzero16. The authors have interpreted
this to mean that converging flow becomes unstable with respect to rotation
disturbances.
Here the analysis of self-similar problem by bifurcation methods shows
that the singularity is preceded by a bifurcation of a new MHD regime with a
nonzero poloidal magnetic field. It is rather similar to an axisymmetric dynamo,
which has been found in a pure hydrodynamic jet flow.17 A new regime already
exists at arbitrarily large current values; hence, the paradoxical effect is
overcome. The account of a turbulence generation is another way of overcoming
the paradox. A simple turbulent model is considered and a variety of bifurcation
phenomena concerning the self-swirling and magnetic dynamo are revealed.

Formulation of the Problem


A steady axisymmetric flow of viscous incompressible electrically
conducting fluid in a half space is considered. A given MHD flow is driven by
an electric current I diverging from a point source 0 placed on the rigid plane or
on the free nondeformed fluid surface. The flow converges near the plane and
forms an axial jet spreading out to infinity (Figs. 1 and 2).
The corresponding solution of the MHD equations is sought in the conical
self-similar class with the velocity field and electromagnetic induction following
118 A. A. PETRUNIN AND V. N. SHTERN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 1 Sketch of the MHD flow. I, supplied electrical current; 0, point


electrode; ( )rigid wall (——-) current lines: (___) streamlines.

the representation in the spherical polar coordinates (r, 0, 9):

vr = -vr ^'(x); v0 = -(r sin0)"1vy(x); v<p = (rsin6)"1vr(x)


(1)
Br = -Br^O'OO; Be = -B(r sin0)-1O(x); B9 = B(rsin0)-1L(x);

where B= |j.0 I/(27Cr), x = cos(0), v, and JLIO are the coefficients of kinematic
viscosity and magnetic permeability, respectively.
Substituting Eq. (1) into the MHD equations and carrying out simple
manipulations, we obtain the system of ordinary differential equations:

(l-x2)yf + 2xy -y2/2 = F(x) - (2)


2x (3)
(l-x2)L" = Bt {yL'-OP + 2(ytL-O'F) + f^ (yL-OF)} (4)
2 1
(l-x )F" = yF - SOL (5)
(l-x2)F! f = 2FP - 2SLL' (6)

where F(x) is an ancillary function, Bt = v/vm is the Batchelor number, vm is the


magnetic viscosity, S = MjP/^nfyv2), and p is the fluid density.
At x = 0 and x = 1, the boundary conditions are formulated:

y(0) = y"(0) = 0 (7)


F'(0) = 0 (8)
These conditions are the impermeability and slip conditions on the free surface.
The no-slip and impermeability conditions at the rigid wall yield

y(0) = y'(0) = 0 (9)


F(0) = 0 (10)
MHD FLOW BIFURCATIONS 119
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 2 Streamline of the meridional motion (left) and poloidal magnetic


fieldline (right). S = 50000, Bt = 0.5; free surface.

The azimuthal magnetic field L satisfies the normalizing condition at x = 0:

L(0)=l (11)

We assume that a medium in the region x < 0 is nonconducting. Then it


follows from Eq. (3) that O" (x) = 0 at x < 0, and in view of regularity at x = -1
one has O(x) = C(l+x). The requirement of the induction continuity at the plane
yields the following boundary condition:

0(0) = O'(0) (12)

It follows from the regularity of velocity and induction at x = 1 that

= 0(1) = 0,

Then from Eq. (2) with allowance for Eq. (13) one can infer that

F(l) = 0 (14)

Differentiating Eq. (2) and taking into account that (l-x2)y" = 0 at x = 1, we


have

F(l) = 0 (15)

Equations (2-6) and boundary conditions (7-15) admit solution with O(x) =
F(x) = 0. But here one seeks a solution with nonzero F(x) and O(x).
120 A. A. PETRUNIN AND V. N. SHTERN

Bifurcation of Poloidal Magnetic Field in a Laminar Regime


One may think that for high conducting media the paradox of loss solution
existence is absent due to the effect of flow on the electromagnetic field. But it
has been shown in Ref. 9 for the case of the rigid plane that this effect merely
leads to an increase in the current critical value. Line 1,1' in Fig. 3 indicates the
boundary of the solution existence region. Above it this regular solution is
absent.
A similar result is obtained for the problem with the free surface. In this
case it is necessary to solve the system of Eqs. (2), (4) and (6) where F = 0,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

d> = 0) with boundary conditions (7), (11) and (13-15). Curve 1,1' in Fig. 4
corresponds to the critical value S*, at which the axial velocity becomes infinite.
For example, S* = 47 at Bt = 0 and S* = 259 at Bt = 10.
Bifurcation of a new solution may come before the loss of solution
existence, which allows the equations at S > S* to be solved. Here we show
that this bifurcation actually occurs. A solution with nonzero poloidal
electromagnetic field appears.
A condition of poloidal field generation is the nontrivial solvability of Eq.
(3) together with Eq. (12). Because of the linearity of Eq. (3), its solution has
an arbitrary multiplier. We choose the normalization O(0) = 1. Then at Bt = 0
the solution is F(x) = 1-x. Hence, 3>'(0) = -1, and expression O'(0) - <l>(0) = -1.
When Bt -> «>, the electromagnetic field is "frozen", and

500

$-&t

2.50

Q 0,5

Fig. 3 Regime map for the rigid wall problem. L and T denote the existence
regions of the laminar and turbulent regimes, respectively, with <& = 0;
L<5 and T<X> denote the regions with <X> * 0.
MHD FLOW BIFURCATIONS 121

The regular solution y(x) is such that, at S > 0, y'(0) > 0 and y'(l) < 0. Then
O'(0) - O(0) > 0 with Bt —> oo. Because of the continuous dependence on
parameter Bt, such a value of Bt has to exist at which 3>'(0) - 3>(0) = 0, [i.e., Eq.
(12) is fulfilled]. This value does exist and is single.
When O(x) is small, then in the linear approach one can neglect the
influence of the self-generated poloidal field and induced rotation on the
meridional flow. Therefore, the problem is reduced to finding a Bt value at
which a nontrivial solution of Eq. (3) exists. Function y(x) is sought from the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

following autonomous system:

(l-x2)y' + 2xy-y 2 /2 = ;
2x
(l-x2)L" = Bt{yL' + 2y'L + ^: (16)

(l-x2)F'" = -2SLL'

with conditions (7), (9), and (11-15).


Our numerical calculations show that the minimum Bt value at which the
self-generation of poloidal field occurs is Btk = 0.35 for both the rigid plane and
the free surface. Curves 2 in Figs. 3 and 4 correspond to the bifurcation of the
new regime. Its existence region is signed by LO. In region L the regular
solution with O(x) = F(x) = 0 is unique.
As S increases, the bifurcation value S-Bt goes from 30 at point K to
S-Bt =55 with Bt -> °o (this limit is shown by line 3 in Fig. 4) for the free
surface and goes from S-Bt = 80 to S-Bt = 215 (line 3 in Fig. 3) for the wall.

Fig. 4 Regime map for the free surface problem. Notations are the same as
those in Fig. 3.
122 A. A. PETRUNIN AND V. N. SHTERN

The new regime with non zero poloidal field and rotation exists at an arbitrarily
large S, in contrast to the initial regime. Examples of distributions for y(x),
L(x), T(x) and F(x) are shown in Fig. 5 (rigid wall). The streamline and
magnetic fieldline at S = 5000 and Bt = 0.5 (free surface) are shown in Fig. 2.

Boundary of the Secondary Regime Existence Region


Region L3> is bounded by curve 4 from the left at Bt < Bt^. On this curve
and on curve 1 a singularity appears. But there are some distinctions. When we
approach curve 1 in region L, the axial jet is amplified infinitely, but <D(x) =
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

F(x) = 0 and current density L'(l) remains bounded. When we approach curve 3
in region LO, the axial jet is amplified infinitely, but first, the jet remains
rotated. This is valid for both types of boundary conditions. Secondly, at the
axis, a strong magnetic jet and accordingly a current layer are formed, but for the
free surface, the magnetic jet intensity tends to zero. At small Bt there is a range
of S where the regular solution of the problem under consideration does not
exist.

Model of a Narrow Turbulent Jet


Now consider in more detail a situation in which Bt < Btk. This case is
interesting because typical Bt numbers are by the order of magnitude 10'6 = 10"7
for real conducting fluids, such as liquid metals. Ab Bt < Bt^ with an increase in
S the loss of the solution existence occurs before the bifurcation of the poloidal
field and rotation occurs. This means that a laminar jet model cannot be used. It
is known that jets become turbulent at low Reynolds numbers. In axisymmetric

2-10

O i

Fig. 5 Distributions of stream function y, circulation T, azimuthal induction


L, and magnetic function O at S - 2050, Bt - 0.5 (LO region in Fig. 3).
MHD FLOW BIFURCATIONS 1 23

submerged jets the turbulent core is known to be bounded by a rather narrow


cone. We use, as the first assumption, the fact that the cone angle equals zero,
that a turbulent jet acts a a line mass sink of constant flow rate per unit length
27T(pq, where q=y(l). Thus, the laminar regime is considered to correspond to the
main term of an outer asymptotic expansion and the turbulent jet is modeled by
this singularity at the axis. It follows from Eq. (2) that q=4 at curves 1 and 4 in
Figs. 3 and 4. A more detailed analysis for a general case has been reported in
Ref. 19.
Regarding the higher entrainment of turbulent jets in comparison with
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

laminar jets, we believe that


y(l) = q > 4 (17)

A value of q is found to satisfy the boundary problem. If q becomes less than 4,


one has to return to laminar formulation of the problem. In the limiting case a
full momentum flux per unit area nze = IT^ • x - ITee • sin(0) must be finite.
The components nee, n^ and n^ have the following form:
2
nee = - 2< 2 F •F"<!-x2> + r2 + s<*2 -

2rx + 2<E>L}

Using Eq. (18), we get in the limit x -» 1

lim [2(F - x F) - F"(l-x2) + T2 + S(02 - L2)] = 0 (19)

From Eqs. (2-6) at y(l)=q*0 the analytical representations of functions


O(x), y(x), L(x), T(x) and F(x) have the form near x=l:

y = q-yi(l-x)+y2(l-x)2-2v+o(l-x)

(20)

F = F0-F,(l-x)+ (l-x)2+F3(l-x)3-v+ o(l-x)


1 24 A. A. PETRUNIN AND V. N. SHTERN

where y = qBt/2 and o(l-x) denotes a function which tends to zero faster than
1-x as x —> 1.

(21)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

F3=(2ror1-2L^)/{2(l-27)(3-27)}

It follows from Eq. (20) that at x=l,

r(l) = r 0 ,F(l) = 0,L(l) = 0

and as x —» 1,
lim [(l-x2)F"] = 0 (22)

This is valid if 7 < 1 (calculations confirm this). Then F0 can be obtained from
Eq. (2):

Using Eq. (19), we have

Fi = F0 + rV2

Differentiating Eq. (2), we get

y1 = (2q-F1)/q

First we consider the problem with the rigid plane. Equations (2-6) are
integrated from xl = 1-8, 8 « 1, where the representation (20) is used, to plane
x = 0. In a general case parameters, q, F0, F2, L!, and F! are chosen so that the
boundary conditions (7-10) are satisfied. Parameters S and Bt remain free.
Evidently, the solution exists with <X>(x) = F(x) = 0 and q > 4. This solution
appears at curve 1,1' and is unique to region T, but it is not unique to region L<E>
(Fig. 3). When S is raised (Bt=const), the value of q asymptotically increases;
i.e., the turbulent jet is amplified.
Now we find a bifurcation of the poloidal magnetic field. For this one
needs to integrate Eqs. (5) (as far as <X> and F are small) and choose parameter S
so that a nontrivial solution of Eq. (3) exists. The bifurcation value S-Bt vs. Bt
corresponds to curve 5 in Fig. 3. The asymptotic relations S(Bt) and q(Bt) are
S=6.15-Bf2 and q=1.46-Bf1. The 7 value grows in magnitude along curve 5
from 0.71 at point K to 0.73 with Bt -^ 0.
MHD FLOW BIFURCATIONS 125
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 6 Viscosity ratio e against the current value. Free surface; Bt = 0.3;
numbers of points S correspond to the curve numbers in Fig. 8.

The generation of the poloidal field gives rise to the azimuthal force which
causes the fluid to rotate. As S increases the poloidal field and rotation amplify,
but q begins to decrease, and at some S, q = 4. This means that relamination
occurs.
Thus, one may describe a typical sequence of the change in regimes with
the growth of S at Bt = const < Btk. When S is small, the flow is laminar, and
O and T are absent (region L in Fig. 3). At some value of S the jet becomes
turbulent (curve 1). Further elevation of S amplifiers the turbulent jet and
accordingly the value of q (region T). On curve 5 the bifurcation of O and T
occurs, but the flow remains turbulent (region TO). At a higher value of S the
turbulence of the axial jet is suppressed (curve 4) and the laminar regime with
O(x) = 0, F(x) = 0 develops (region LO).
Consider the problem with the free surface. Now the sequence of the
change in regimes is not the same, as in the previous case. In region T Eqs. (2-
6) with conditions (8) and (21) admit solution with F(x)=ro=const*0. The
parameter F0 is free, as are S and Bt, and q depends on them. But not all of the
solutions in the parameter space (S, Bt, F0) are the limiting ones, when the
angle of the turbulent cone tends to zero; only some of these are limiting. In
order to seek the limiting solutions one has to consider a model of the turbulent
jet in which angle is finite.
126 A. A. PETRUNIN AND V. N. SHTERN

Model of Fluid with Eddy Viscosity


An investigation of the conical flows of fluid with variable viscosity
reveals an unexpected phenomenon: a bifurcation of a self-swirl of a jet flows.20
One may expect that the self-swirling also happens in our problem.
We choose the simple turbulent model with a step function for viscosity
distribution (as in Ref. 20):

v = YI at 0 < x < xt (region 1)


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

v = VT at xt < x < 1 (region 2)

where xt is the boundary of the turbulent region. Near the plane in region 1 the
flow is laminar and vi is the molecular kinematic viscosity, whereas in region 2
the flow is turbulent and vt = E-V! (e > 1) is the eddy viscosity. At the boundary
of the turbulent region, x = xt, the continuity conditions have to be fulfilled by
the velocity of the momentum flux tensor and of the magnetic induction:

y2(xt) = eyi(xt) , y^xj = ey'^xj, T2(xt) =

, 0'2(Xt) = O'KxJ, L2(Xt) = L^xJ (23)

and S2 = E'Si

We suppose that boundary xt coincides with the maximum locus of y(x), i.e.,
y'(xt) = 0.
According to Schlichting,21 we believe that for the turbulent non-swirling
jet the following condition is valid:

- = - y ' 2 ( l ) = 460.5 (24)

This means that the jet is laminar as long as y, (1) < -460.5. After that the
increase of S leads e to grow in accordance to Eq. (24).
The sequence of the events remains the same in the problem with the rigid
wall, but the numerical values vary. For example, the MHD bifurcation in the
laminar regime now starts from point K' (Bt = 0.44 and S-Bt = 1 10).
Furthermore, we deal with the free surface problem. As S grows, the value
of E increases in accordance to Eq. (24), moving along curve 1 in Fig. 6. At e =
1.085 at point 85, bifurcation of a self-swirling regime occurs (see Figs. 6 and
7). Since this bifurcation is supercritical, the swirling regime seems to be stable
and the initial nonswirling regime becomes unstable. The value of e at which
bifurcation occurs depends on Bt to a small extent. The bifurcation values of e,
L'(l), F"(l), and S are chosen so that the boundary conditions (7) (11) and (16) at
x = 0 and the boundary xt are fulfilled and a nontrivial solution of Eq. (5) with
MHD FLOW BIFURCATIONS 127

10

5
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

•Sjj 13?
250 500

Fig. 7 Dependence of circulation and meridional induction at the free surface on


current value. Bt = 0.3.

r'(0) = 0 exists. The bifurcation value of S vs. Bt corresponds to curve 5 in


Fig. 8. In order to calculate the self-swirling regime, it is necessary to know the
dependence of eddy viscosity on S. This relation is unknown, but we suppose it
is the same as that for the nonswirling flow: i.e., Eq. (24) holds true. The
dependence e(S-Bt) is shown in Fig. 6 (curve 2). When S increases, T grows,
but e decreases (Figs. 6 and 7).
Furthermore, the bifurcation of the poloidal field occurs (curve 6 in Fig. 8
and point S 6 in Figs. 6 and 7) and the magnetic field have quite specific signs,
depending on the rotation direction. F and G increase with the growth of S.
This causes q to decrease and the turbulence to be suppressed. At some S, e = 1
(point S4, Fig. 6). This means that relamination occurs and at higher S the
laminar regime with 3>(x) * 0, F(x) * 0 exists (LO region, Fig. 8).
Another scenario is observed when we go in the direction of S decreasing
from the region L3>. In this case O changes its sign as it crosses curve 6 (point
S 6 in Fig. 7). As S decreases, O passes the maximum and turns into zero
together with F (point S7 in Fig. 7 and curve 7 in Fig. 8). Thus, in the region
between curve 7 and curve 6 in Fig. 8 the solution with F(x) ^ 0 and <X>(x) = 0 is
not unique because there are the MHD solution with nonzero <E> and the initial
turbulent regime with F = r = 0 (curves 2, 4 and 1 in Fig. 6). At Bt = 0.3 the
MHD solution exists between S7 and S6 in Fig. 7. In the limiting transition to
the infinitely narrow jet, the asymptotic behavior of the numerical solution
shows that 3> -> 0, but T remains finite. Curve 7 in Fig. 4 corresponds to the
limiting positions of points S7 (see Fig. 7).
128 A. A. PETRUNIN AND V. N. SHTERN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 8 The same as in Fig. 4, but using the Schlichting model for the
turbulent axial jet. TF is the region of turbulent regime with rotation,
but without induction.
Conclusion
It is useful to give a qualitative physical interpretation of how the
convergent motion of a conducting fluid may generate an azimuthal electric
current j9, which is necessary for the poloidal electromagnetic field to be
sustained. Consider a toroidal-shaped fluid conductor placed near the plane
axisymmetrically, and suppose there is a small disturbance of the poloidal field.
Converging meridional flow carries the conductor to the axis and compresses it.
The magnetic fieldlines are intersected by the conductor. As a result, an
azimuthal current is generated. Induced current leads to excitation of the poloidal
electromagnetic field, which has the same direction as the initial disturbance.
But the amplification may be prevented by diffusion and dissipation processes.
At small magnetic Reynolds number Rm these processes predominate and the
dynamo is absent. But when Rm exceeds its critical value, the generation
predominates and the initial disturbance of the poloidal induction beginss to
amplify. This growth stops when a significant part of the kinetic energy of flow
is transformed into the magnetic one and therefore the induction generation is
attenuated. This generation mechanism does not contradict Cowling's theorem,22
because the magnetic fieldlines are unclosed, and Braginsky's results 23 because
the poloidal induction decays at infinity being inversely proportional the first
power of the distance. Because of higher gradients of the induction in the
Braginsky case, the dissipative effects predominate at all R m and the dynamo
effect is absent.
A similar phenomenon concerns the appearance of the self-swirling at
small Bt. This effect happens at a rather small current value, and one may hope
MHD FLOW BIFURCATIONS 129

that this effect can be observed experimentally. In this relation the result of an
experiment described in Ref. 16 may have a new interpretation. At some critical
current value, the azimuthal rotation has been observed in a hemispherical
container filled with mercury. Authors believe that this rotation occurs due to
the Earth's magnetic field, though our result points out that at these current
values the near-axis jet may become turbulent and the rotation may appear as a
result of the bifurcation.
References
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

iZhigulev, V.N., "On Ejection Effect Due to an Electrical Current


Discharge," Doklady Akademii Nauk SSSR, Vol. 130, No 2, 1960, pp. 280-
283.
2
Lundquist, S., "On the Hydromagnetic Viscous Flow Generated by a
Diverging Electric Current," Arkiv foer Fysik, Vol. 40, No.5, 1969, pp. 85-95.
3
Shercliff, J.A., "Fluid Motion Due to an Electric Current Source," Journal
of Fluid Mechanics, Vol. 40, No. 2, 1970, pp. 241-250.
4
Shcherbinin, E.V., "On a Kind of Exact Solutions in the
Magnetohydrodynamics," Magnitnaya Gidrodinamika, 1969, Vol. 5, No. 4, pp.
46-58.
5
Sozou, C., "On Fluid Motion Induced by Electric Current Source," Journal
of Fluid Mechanics, Vol. 49, No. 1, 1971, 25-32.
6
Bojarevich, V., Freiberg, J. A., Shiliva, E. I. and Shcherbinin, E. V.,
Electrically Induced Vortical Flows. Kluwer, Dordrecht, 1989.
7
Goldshtik, M.A., "A Paradoxical Solution of the Navier-Stokes
Equations," Prikladnaya Matematika Mekhanika, Vol. 24, No 4, 1960, pp. 610-
621.
8
Narain, J.P. and Uberoi, M.S., "Magnetohydrodynamics of Conical
Flows," Physics of Fluids, Vol. 17, No. 12, 1971, pp. 2687-2692.
9
Sozou, C. and English, H. "Fluid Motions Induced by an Electric Current
Discharge," Proceedings of the Royal Society of London Series A, Vol. A329,
1972, pp. 71-81.
10
Sozou, C. and Picketing, W.M., "Magnetohydrodynamic Flow due to the
Discharge of an Electric Current in a Hemispherical Container," Journal of
Fluid Mechanics, Vol. 73, No 4,. 1976, 641-650.
H
Bojarevich, V.V. and Millere, R., "Amplification of Rotation of a
Meridional Electrovortex Flow in a Hemisphere," Magnitnaya Gidrodinamika,
No. 1982, pp. 51-56.
12
Atthey, D.R., "A Mathematical Model for Fluid Flow in a Weld Pool at
High Currents," Journal of Fluid Mechanics, Vol. 98, No. 4, 1980, pp. 787-
801.
13
Bojarevich, V.V., Sharamkin, N.I., and Shcherbinin, E.V., "Effect of
Longitudinal Magnetic Field on the Medium Motion by Arc and Electrical
Welding," Magnitaya Gidrodinamika, No. 1977, pp. 115-120.
14
Crine, R.E. and Weatherill, N.P., "Fluid Flow in a Hemispherical
Container Induced by a Distributed Source of Current and Superimposed Uniform
Magnetic Field," Journal of Fluid Mechanics, Vol. 99, No 1, 1980, pp. 11-12.
130 A. A. PETRUNIN AND V. N. SHTERN

15
Millere, R., Sharamkin, N.I. and Shcherbinin, E.V., "Effect of
Longitudinal Magnetic Field on Electrovortex Flow in the Cylindrical Volume,"
Magnitnaya Gidrodinamika, No. 1, 1980, pp. 81-85.
16
Bojarevich, V.V., and Shcherbinin, E.V., "Azimuthal Rotation in the
Axisymmetric Meridional Row due to an Electric Current Source," Journal of
Fluid Mechanics, Vol. 126, No 3 1983, pp. 413-430.
17
Goldshtik, M. A. and Shtern, V. N. "Self-Similar Hydromagnetic
Dynamo," Jornal of Experimental and Theoretical Physics. (Sov.). Vol. 96, No.
5, 1989, pp. 1728-1743.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

18
Goldshtik, M.A. and Shtern, V.N., "Conical Flows of Fluid with Variable
Viscosity," Proceedings of the Royal Society of London, Series A* Vol. A419,
1988, pp. 91-106.
19
Schneider,W., "Flow Induced by Jets and Plumes," Journal of Fluid
Mechanics, Vol. 108, 1981, No 1, pp. 55-65.
20
Goldshtik, M.A., Shtern, V.N., and Yavorsky, N.I., Viscous Flows with
Paradoxical Features, Nauka, Novosibirsk, USSR, 1989 (in Russian).
21
Schlichting, H., Grenzschicht Theorie. Braun, Karlsruhe, Germany, 1965.
22
Cowling, T.G. Magnetohydrodynamics, Interscience, New York, 1957.
23
Braginsky, S. I. "Self-Excitation of Magnetic Field During the Motion of
a Highly Conducting Fluid," Jornal of Experimental and Theoretical Physics.
X. Vol. 48, No 6, 1964, pp. 1084-1098.
Homogeneous MHD Turbulence at Weak
Magnetic Reynolds Numbers:
Approach to Angular-Dependent Spectra
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

C. Cambon*
Ecole Centrale de Lyon, Ecully, France

Abstract

To quantify the two-dimensional features in MHD homogeneous


turbulence at weak magnetic Reynolds number, the complete spectral
tensor of double velocity correlations is approached. For
axisymmetric (around the external magnetic field B Q ) turbulence, two
spectral two-dimensional eigenmodes <J**a(kj_,k^,t), a = 1,2, have to
be considered. The governing equations are closed by an EDQNM model,
which takes into account also linear operators in a straightforward
way. Two kinds of anisotropy are scrutinized. The first one is the
directional dependence, or the angular dependence in wave space of
the whole spectral density of kinetic energy e = 1/2 ($1 1 + <&2 ) .
Such anisotropy, which is mainly induced by the conical structure of
the Joule dissipation term, leads to concentrate e in the transverse
wave plane (k 1 BQ or k^. = 0). The second one, called polarization,
is the emergence of a net difference Z = 1/2 (4^2 - 4>11 ). It is
found that the polarization is only induced by nonlinear
interactions; the EDQNM model used in this paper leads to <f>1 > <f£ in
the wave-vector range, which has a significant contribution to the
energy. By using a splitting in terms of e and Z, both the trends
are then characterized by_classical "deviatoric" quantities
u|> - Uj_ , (d^ - (0^ , u2^ L^ - 2 u^Lj^ , including velocity, vorticity
and integral length scales. The splitting into a directional and a
polarization part displays the relevant indicators of two
dimensionality, and provides new information for improving the
current single-point closure models. A direct validation of the
EDQNM model is also proposed by using the experimental results of
Caperan and Alemany.

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Research Scientist, CNRS, Laboratoire de M£canique des Fluides et d'Acoustique.
131
132 C. CAMBON

Introduction
The effect of an external magnetic field, among other body
forces, is capable of causing the transition from three to
two-dimensional structures in a turbulent conducting fluid.1 The
corresponding anisotropy consists primarily of a conical structure
of the spectral region that contains energy. This can be seen from
the Rapid Distortion Theory and has been experimentally revealed by
detailed measures at two points. In this paper we propose to
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

provide a detailed examination of the previous analytical approaches


using a closure model for nonlinear interactions. The same strategy
has been successfully applied to the case of turbulence subjected to
solid-body rotation.4
First, one retains the complete information given by the
spectral tensor of double correlations. In the axisymmetric
configuration (with respect to the axis n bearing the external
magnetic field BQ ), two spectral modes, <J>11 and ^2, are considered.
They depend on both the modulus k and the orientation 9k of the wave
vector k with respect to the axis n (or, equivalently, on
k^ and kjj . The <£ corresponds to the spectral energy contained in
a transversal (In) and nondivergent velocity component. Similarly,
the following spectral quantities are defined:

(1)

where n^ = cosS^ = k.n/k = k^/k is chosen as the angular parameter,


and e is associated with the total kinetic energy. A first kind of
anisotropy (referred to as the directional part) is linked to the
angular dependence of e, whereas the polarization is given by the
deviator Z, at fixed k.
Starting from a three-dimensional isotropic turbulence
[where e=E(k) /4irk2, z=0] , a "true" two-dimensional isotropic state
could be reached if the two following trends are identified:
1) e concentrates on the transverse wave plane M^=0 (conical
distribution of the energy); and
2) <f> becomes dominant with respect to 4^ in the wave plane
Mfc^O, so that Z reaches a negative value.

Accordingly, <$ (k, ^=0, t) would contain the whole energy if the
true two-dimensional state had been achieved. Regarding the
two-dimensional features in physical space, the true two-dimensional
state requires that the velocity field has only two components
(u^=0) and that the vorticity field has only one component (0)^=0).
In several flow patterns, wrongly considered as twodimensional, only
one of these tendencies is achieved. For example, a strong decrease
HOMOGENEOUS MHD TURBULENCE 133

of the axial variability d/dxj, leads to enhance the one-dimensional


character of the vorticity field in both rotated and MHD turbulent
flows, whereas the decrease of the vertical velocity u^> is not
obtained (an opposite trend u^. > u]^ is more often found).
On the contrary, the collapse of the axial (or vertical in this
case) velocity component (u^ -» OJ in stably stratified turbulence
can be obtained without a significant decay of CJ^. The concentration
of turbulent motion in thin horizontal layers, vertically
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

decorrelated, could even enhance the vertical variability (and thus


increase OJj_).Note that the influence of the axial variability on the
one-component structure of the vorticity field can be seen from the
following relation:

n x 0) =* 0), = V ux. - -— (2)

Hence, the two trends u^ -* 0 and du/dxf* 0 have to be found


simultaneously, in order to ensure a two-dimensional state. Coming
back to spectral space, it is clear that the trend 1 corresponds to
a decrease in the axial variability, according to
du/dxj,= 0 <* k^u = 0 where the subscript denotes a
three-dimensional Fourier transform. Trend 2 always leads to
enhancing 0)|> with respect to co^, considering that <P contains the
whole axial vorticity. A peculiar result is that trends 1 and 2 can
have a cumulative effect, when the one-component structure of the
vorticity is considered, and the opposite effect regarding the
two-components structure of the velocity. As in Ref. 4 , these
features are supported by a detailed relationship. In the next
section three deviatoric quantities, relevant in axisymmetric
turbulence, are split into two parts in order to quantify the
different (and often contradictory) trends 1 and 2:

4- 4= (4-4)' + (4-4)
—— —— ,—— —i e >—— —v Z
(4 - wi - (<4 - "ij + (<4 - ^J
u| L^ - 2 u^ Lj_ = (u| LS - 2 u* L J

Such an analysis prepares improvements of cur_rent_one-point closure


models, in which the deviatoric part (here u|j. - u^) of the Reynolds
stress tensor is used as the unique anisotropy indicator. In the
section on the characterization of linear and nonlinear trends, the
equations that govern the two relevant spectra <f>1 1 and <J^2 (or,
similarly, e and Z in Eq.(l)) are discussed. In the limit of low
134 C. CAMBON

magnetic Reynolds numbers, it is found that:

f— -i- 2 vk2 + 2M cos2ekj e(k, cos9k, t) = Te(k, cos9k, t)


(4)
f— + 2 vk2 + 2M cos29k] Z(k, cos9k, t) * TZ (k, cos9k, t)

in which i> is the kinematic viscosity, M = oBg /p is the intensity of


the joule dissipation effect (in a fluid having a conductivity a and
a density p, and undergoing an external magnetic field BQ ) . MH|
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

gives the magnitude of the "conical" joule dissipation effect and


corresponds to a linear operator.
On the other hand, Te and TZ represent transfer-like terms,
mediated by nonlinear interactions. They involve third-order
correlations. When Te and TZ are discarded (according to the rapid
distortion theory), a trend of type 1 is created, but the
increasingly conical structure is the same for 4>1 1 and <J^2, so that
Z remains null (if the calculation is started with three-dimensional
isotropic initial data). The directional anisotropy induced by
linear terms was quantified by Moreau5 , by means of an angle 3
(— —S c
(which could be closely connected with fu^, - Uj] in our analysis).
Nevertheless, the influence of the polarization (appearence of Z) is
completely ignored in the current models. Only a nonlinear spectral
approach allows this effect to be captured. Hence, TC and TZ are
computed in terms of e and Z by implementing (nearly without change)
the anisotropic EDQNM model used for studying turbulence in a
rotating fluid.4 Trends of type 1 and 2 are obtained from first
numerical results of the EDQNM model for Eqs. (4) and discussed in
connection with the indicators (Eqs. (3)).
The section on experimental validation prepares a detailed
comparison with the experimental results obtained by Caperan and
Alemany. 3 The comparison particularly concerns the one-dimensional
spectrum E^(k^,t), which clearly shows the appearence of a k"3 zone
in the inert ial range. A detailed relationship between the different
spectra is given, as well as a simple method for generating initial
data (in the model Eq. (4) a method consistent with the experiment
is also proposed) .
Moreover the experimental derivation of a two-dimensional
spectrum U33 (kj_,k^,t) = <J^2 (k,cos9,t) sin29 gives an opportunity to
validate the EDQNM model at its finest level of description.
The experiment we consider here is performed in a deep tank full
of mercury, and the turbulence is created by the fall of a grid. The
times of decay are huge and the Reynolds number are very high
compared with the ones obtained in a wind tunnel. Thus, the EDQNM
theory is well adapted. The use of a logarithmic step for the
modulus k allows the computational time with respect to a DNS to be
reduced dramatically. Moreover, the statistical symmetries are
HOMOGENEOUS MHD TURBULENCE 135

better ensured and the angular dependence is preserved. Such a


dependency is lost in DNS data processing in which spherical
integration is carried out by way of statistical averaging.

Two-Dimensional Indicators in Physical and Spectral Space


The structure of double velocity correlations at two points is
highly simplified by working in Fourier space and by using a local
frame (e1, e2, k/k) attached to the wave vector k, in accordance
with Craya ° and Herring. In this frame, the "incompressible"
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(verifying u.k = 0) spectrum of velocity u and the associated


covariance matrix (u- Ujy have only a few nonzero components, so
that
'^\
<P 4,11 4> o e-ReZ DmZ 0
~2 4,12 $22 o SB 3mZ e+ReZ 0 (5)
<P
0 0 0 0 0 0

Moreover choosing «1 = (k x n)/lk x nl and e2 = (k x e 1 )/Ik x e1 I,


^ ^2.
<p is attached to the axial vorticity, and <p contains the whole
axial velocity; thus the four components <f**P of the spectral tensor
U , connected with yp <p \ have a simple physical interpretation.
In order to display invariant terms and to simplify the expressions
in the fixed frame of reference (in which "physical" single-point
correlations are expressed), the two quantities e and Z are used.4
In the plane (e1, e2), such a set of variables corresponds to a
splitting into trace and deviator. The e is the density of kinetic
energy attached to k and the complex deviator
(Z = ReZ + I DmZ ; I2 = -1) gives the nonisotropic tensorial
structure of the spectral tensor at fixed k. (It can be found that
e+ Izl, e- iz! , and ArgZ determine the nonzero principal
components and the associated principal axes of U). In the fixed
frame of reference it is found:

U f j ( k , t ) = e(k,t) + R e ( Z ( k , t ) N, ( k ) N j ( k ) ]

with M ( k ) = e 2 ( k ) - I e 1 ( k ) ; p .} (k) . . - k, j /k 2 = ee ee + e?
e e? .
The strict isotropy requires that Z = 0 and that e has an uniform
spherical distribution. By introducing the classical energy spectrum
E(k,t) , which is the integral of e over a spherical shell of radius I
k, a splitting of U into three parts is easily obtained:

,t)
u(k,t) = Re {ZN ® N}
4lTk2
(6)
three-dimensional directional tensorial
isotropic part anisotropy anisotropy
(angular dependency of e) (at fixed k)
136 C. CAMBON

-. -ISO ~e -Z
U * U + U + U

2 f -e -Z
Note that e - E/4irk has zero integral, and I U and U have zero
trace. Thus, two kinds of anisotropy could be accounted for in any
integral expression including U (Reynolds stress tensor, "rapid"
part of the pressure-strain correlations, vorticity correlations,
—e
integral length scales, etc.). The first one, U , attached to
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

e - E/4irk2 , reflects the "dimensionality," and the second one, --Z


U ,
attached to Z, reflects the "componentality" of the spectral tensor.
(These two terms are from Reynolds who also proposed to study
separately similar features in single-point closures). Regarding the
classic terminology, relevant for transverse dispersive waves (in
-* ~z
wave space) U corresponds to the directional dependence and U does
the polarization. Such a terminology will be used from now on.
By integrating over the wave space, a splitting corresponding to
Eg. (6) is exactly derived for the Reynolds stress tensor:

Uj uj (t) - / J / U, j (k,t) d3k


(7)

q2(t) —— +q 2 (t) afj.(t) + q2(t)

where q2/2 is the turbulent kinetic energy. The classical


"deviatoric" (with zero trace) tensor,

(t) - u iUj /q 2 - Sj j/3

appears to have two contributions (ae and az ) , which reflect the


directional and the polarization dependence, respectively. Note that
a?j is also found to be simply connected with the structure tensor
used by Kida and Hunt9 and Reynolds8 :

k )C
i J

Relations (5), (6), and (7) are valid for any anisotropic
homogeneous and incompressible turbulent flow. (Possibly a helicity
term could appear in Eqs.(5) but has no contribution in Eq. (7) and
thus it has been ignored for the sake of brevity) .
In the case of axisymmetry (around the unit vector n) including
reflectional symmetry, the extradiagonal mode <}>12= 3mZ vanishes.
Such a configuration is relevant here, when the choice of n is
linked with the external magnetic vector BQ , and the simplified
relationship introduced in the first section is valid. Regarding any
HOMOGENEOUS MHD TURBULENCE 137

single-point correlation tensor R}- • , we are concerned with.only two


independent scalar terms R . - and R-•n.n-, or, similarly,

R
ii - V
R_L = ———————

Thus, the relations (6) and (7) lead to


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

4 - ui = j <*2 a ?j n i n j = (4 - ui) + (4 - ui) <8>


(bo
(4-4)' - JOw
X

/ - ~T\ Z 3 ft» 2

with n^ = cos9 = k^/k. A slightly different form of Eqs. (8) and (9)
is obtained by reintroducing the angle p (Moreau's proposal5), which
characterize the "directional", or the "conical" structure of the
spectral region that contains energy:

— (1 - 3 cos^p) (11)
4

(
with
J J J cos29ke(k,t) d3k
cos2(3 - ' ' ^——-————————— (13)
!! !•(*>*>
The spectral formalism allows a quite similar treatment of the
vorticity equations (CO = Ikxu):

—.e ,— —;Z
(14)

The spectral derivation of the two contributions (in terms of e and


Z) to this vorticity term is simply obtained from Eqs. (9) and (10)
by only changing e into k2e and Z into -k2Z. If a separated
contribution coming from e and Z is not _accounted for, the
deviatoric part of the Reynolds stress tensor u2^ - u^ appears to be
a very poor and insignificant indicator of anisotropy. Hence, a last
indicator was proposed by Cambon and Jacquin , and was extensively
studied in the experimental approach of rotated turbulence
138 C. CAMBON

by Jacquin et al. 1 0 :

- 2 u2 L = uL - 2

- 2 u2!^ = ' T T 2 J 0 Z ( k , O f t ) k dk

Where L^> and L± denote the integral length scales that characterize
the correlation at two points, separated along the axial direction,
between axial and transverse velocity components, respectively. For
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

example,

? ft»
W = Jo^V<x)V<x + rn) )dr
It i« clear from Eqs. (15) that this last indicator emphasizes the
polarization effect in the transverse wave plane 10^ » cos9k = 0.
The relationship of Eq. (8) to Eqs. (15) allows a complete
discussion of the anisotropic tendencies, including the influence of
both a directional and a polarization dependence on base of
physically relevant indicators.

Characterization of Linear and Nonlinear Trends


In the following three-dimensional isotropy, "true"
two-dimensional isotropy, "linear" MHD behavior, and nonlinear
behavior, respectively, are presented. The results are summarized in
Table 1.

Three-Dimensional Isotropy
Since e = E/4irk2 and Z s 0, all of the anisotropy indicators are
separately and rigorously null. From this definition, it is also
found that cos2(3 = 1/3.

True Two-Dimensional Isotropy


In true two-dimensional isotropy, e and Z are concentrated in
the transverse wave plane, in accordance with

; Z= - - S ( V ) or <t> = e and
2irk ' 2irk '

Accordingly,

cos 2 p = 0 and u2^ = 0;

(4 - ui)C (4 - "i
- 3c * 2 / 4;
HOMOGENEOUS MHD TURBULENCE 139

Table 1 Values of the dimensionless indicators, which reflect


directional dependence (superscript e) and polarization (superscript
Z) in physical space N°, interaction parameter.

Axisymmetric
MHD case = (1-3 cos2(3)/4 («)'/? (4-^i)/^ u$L^-2u2L1

3D I sot ropy
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(initial data) 0 0 0 0

linear -» 1/4 0 -> 1/4 0


evolution
moderate positive value reaches a positive value strong
value of N° attenuated negative value (attenuated) negative
0 < < 1/4 value
-1/2
1/4 -3/4
"Pure 2D" or (cos(3 as 0) or \\A, - Oj -00
\ ' /

and

- 2 u - 2 (17)

Linear MHO Behavior (Started with Three-Dimensional Isotropic


Initial Data)
By considering a very large value of an "interaction parameter"
(see Ref. 3),

4
the terms TC and TZ can be ignored in Eqs. (4). Thus, the joule
dissipation term 2M cos29k leads to a directional dependence of the
Two-Dimensional type, but the polarization (Z) remains null.
For very large nondimensional times Mt -» oo, E and <f vanish, but
the nondimensional indicators reach the following limits:

I - «i —;
4
140 C. CAMBON

From Eq. (12) it can also be derived that

4
= ==== 2 (18)
u
i "i
Moreover, u^L^, - 2 UJ_LJ_ remains null, together with Z.

Nonlinear MHD behavior


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

For moderate values of N° (from 10"1 to 10), the nonisotropic


transfer-like terms Te and TZ (computed by EDQNM; see the following
section) become relevant. The trends in e and Z are qualitatively
the same as those for initially isotropic turbulence subjected to
pure rotation (see Ref. 4).
Regarding e, Te tends to reduce the angular dependence of e,
especially for large wave numbers, but the conical structure induced
by the linear joule dissipation term holds again. Such a balance
between linear effects and transfer terms mediated by nonlinear
interactions is currently observed in turbulent flows. Accordingly,

u w
(4
———
- i) (4 -
'—— and ———————
q2 OJ2

again reach a positive value, but are reduced when compared to their
two-dimensional (or linear asymptotic) value 1/4.
Regarding Z, TZ is created with a negative value in the range of
wave vectors containing energy (small values of k and cos9k = k^/k).
The appearance of a polarization induced only by nonlinear
interactions is the most important result of the EDQNM model. This
effect is primarily reflected by the emergence ofu|. L^> - 2 u^ Lj_
with a strong negative value, as in the true two-dimensional case. A
simultaneous effect is the emergence of u - Uj with a negative
yalue_.__Thus, the two contributions with opposite signs of e and Z to
u^ - u^ tend to reduce the "linear" increase of this indicator, but
the polarization is not strong enough (with_respect to the
directional dependence) to change the sign (u|, < Uj] as in the
"true" two-dimensional case. The qualitative linear and nonlinear
effects are summarized in the Figs. 1 and 2. Figure 1 shows the
directional dependence by plotting the spectra 4irk^e(k,cos9k ,t) for
different values of cos9k and shows the polarization by plotting a
segment (the heavy line) of length k2Z. In agreement with the
sketches used in rotating turbulence,4 the segment is horizontal for
Z < 0 (<£11 dominant, in accordance with a two-dimensional tendency)
and vertical for Z>Q (<&* dominant, anti-two-dimensional tendency).
Figure 2 shows the associated behaviors of u^/u^ and 2
HOMOGENEOUS MHD TURBULENCE 141

icr
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0.1 10 0.1 10 (cm")

equator

"Linear" (R.D.T) EDQNM2

Fig. 1 Basic spectra e(k,cos6,t) and k2Z(k,cos0,t) for different


values of Jc (horizontal axis) and cosG * k^./k. Both linear and
EDQNM2 cases are calculated with the same three-dimensional
isotropic initial data. The value of M corresponds to an initial
interaction parameter N°=2, and the dimensionless time is Mt » 3.

Towards an experimental validation


The experiment briefly described at the end of the first section
can be regarded as classical homogeneous grid generated turbulence,
subjected to a magnetic field in the streamwise direction. Thus, the
streamwise distance x from the grid plays the same role as the time,
according to

x "0
m — -—
Im m

where m is the mesh size of the grid and UM its uniform speed.
Assuming that the turbulence is almost isotropic in the initial test
section xQ/m, the following expression is first adjusted to the
measured experimental one-dimensional spectrum:

— L^ exp-
4
2 u% —.
5/6
(19)
L 2
<V o> ]
142 C. CAMBON

LINEAR LINEAR
2.00 EDQNH2 EDQNH2

1.80

1.60

1.4)

1.20

1.00

Fig. 2 Evolution of the ratios u|y u^ and (with respect to


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Mt) derived from the basic spectra e and Z in the same conditions as
those in Fig. 1.

by choosing optimal values for LQ and t] (Holland and Key, private


communication). E(k) is then derived analytically from this
expression by using standard formulas for isotropic homogeneous
turbulence. The initial data in our model (Egs. (4)) are therefore

e(k,cos8k,0) » E(k)/4trk2; Z(k,cos9k,0) = 0

Note that a straightforward derivation of from e and Z (in


the axisymmetric case) is given by the following:

4-rr = e + Z (20)

Caperan (private communication) found that an expression more


general than Eq. (19) fit very well with the experimental data for
all x/m:

exp [-<VD 4 / 3 ]

*i
(21)
l+ <Vi> "
i a / c -i5/6

In which LQ and r\ depend on "time" x/m Obviously, both the


k"" 5 / 3 and k" power laws range are thus accounted for. The role of
an anisotropic transfer term Te + TZ (associated to e + Z = <£^2)
appears to be crucial for predicting the emergence of a k"3 range,
in accordance with the straightforward Eq. (20). Hence, a very brief
explanation about the EDQNM2 closure is now given. The symbolic form
of the closure of triple correlations in terms of double ones is

<uuu> = J 0 EO 0 ED 0 E O < u u > < u u > (22)

where O^0 is the product of the matrix O, which generates the linear
response (as in ROT) by an eddy damping factor. In the case of
solid-body rotation, the linear response is easily calculated by
HOMOGENEOUS MHD TURBULENCE 143

using convenient eigenmodas and associated eigenvalues of 6 (see


Table 2):

) * exp ± I 20 — t (<p ± I <p (23)

Thus, the terms Tft and TZ derived from Eq. (22) have the following
form:

e 3
8'kpq
kDa S (€k,€'p,€"q)d p
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Te(k,t) (24)
€=±1 ( *// V// <*/A
1 + 2in 8.k pnq €—— + €'—— + €"——
€'=±1 l k p qj
€"=±1

in which Q^pq is the classical (isotropic) relaxation time of triple


correlations induced by fourth-order cumulants and viscous effects
~1 ^2
(k + p +q » 0). The expression <p ± I<p gives convenient eigenmodes
for MHD, so that the EDQNM2 expression for Te is derived from Eq.
(24) by changing only

V P^ ^
— + €' — + €" — into M — + — + —
k p qj
The detailed expression for Se in Eq. (24) in terms of e and Z can
be obtained from the author upon request.

Table 2 Convenient interaction parameters, eigenmodes, and


eigenvalues for solving the linear problem (RDT) and for
implementing the EDQNM2 model in three similar cases

Interaction Associated
parameter Eigen modes eigen values
'"•-i ""^ i
cp = u.e
MHD N° = M —— M cos29
u' ^ ^ 2
cp = u.e
Solid body -1 -2
rotation V -2^ <p ± I q> ± 21 f* cos6

Stably stratified -1 .....0


cp ........
Turbulence -2 pg ~
1
(Boussinesq F" - Nw —— <P ± 1 —— T ± I Nw sin6
approx . ) u' N
w

R , Rossby number; F, Froude number; Q, solid-body rotation rate;


N , gravity waves frequency; T, fluctuating temperature field.
144 C. CAMBON

Concluding Remark*
Two kinds of anisotropy dependence, directional and
polarization, are investigated in this paper. The knowledge of both
these features appears to be crucial to accurately determine the
emergence of a two-dimensional structure. The two tendencies are
first identified in two-dimensional spectral space, in which "exact"
linear behavior is very easy to calculate, but the influence on the
most relevant single-point correlations is displayed through a
detailed relationship (see the section on anisotropy indicators).
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

This work can be seen as a prerequisite to improve single-point


closure models in the case of MHD, but also "rotated" and "stably
stratified" 1 1 turbulence, especially regarding two-dimensional
trends. The main contribution to the knowledge of the structure of
MHD homogeneous turbulence is the prediction of a "polarization"
effect induced only by nonlinear interactions. This effect
contributes to reinforce both the one-component structure of the
vorticity and the two-component structure of the velocity.
Nevertheless, becaus_e_of the strong linear directional effect, which
tend* to increase u^/u^, the two-component structure of the velocity
field (collapse of u^) is not obtained.
It is difficult to extract direct information about polarization
from published results obtained experimentally or by DNS.
Nevertheless, one can point out that the rise of a polarization in
the transverse wave plane (in which the energy tends to be
concentrated),

<J*2(k^O,k±,t) - ^(k^-Ojkj^t) - Z^Ht-O,*:) < 0

is in agreement with a straightforward (nonisotropic) approach based


on the renormalization group. Garnier et al.12 found that, in the
equation for <tf*^ (see Eq. (5)), v should be renormalized under a
tensorial form (va(3) (k±), so that (v22) * v^ > (v 1 1 ) = v± . In
the near future, the rise of a k"3 power law range in the
one-dimensional spectrum will be scrutinized in connection with
nonlinear directional and (especially) polarization effects (role of
Te and TZ in Eq. (4)). A detailed comparison between the two-point
closure model and the experiment (see the preceding section) is in
progress.
Acknowledgment
The author is grateful to P. Caperan, L. Sallen, L. Van Haren,
and O. Vitali (Metraflu company) for helpful discussions and
assistance.

References
1
Schumann, U., "Numerical Simulation of the Transition from Three
to Two Dimensional Turbulence Under a Uniform Magnetic Field,"
Journal of Fluid Mechanics. Vol. 74, 1976, pp. 31.
HOMOGENEOUS MHD TURBULENCE 145

2
Moffatt, H. K., "On the Suppression of Turbulence by a Uniform
Magnetic Field," Journal of Fluid Mechanics, Vol. 28, 1977, pp. 571.
3
Caperan, Ph.and Alemany, A., "Turbulence homogene MHD i faible
nombre de Reynolds magnetique," Journal de mScanioue theorioue et
appliouee. Vol. 4, N° 2, 1985, pp. 175-200.
4
Cambon, C. and Jacquin, L., "Spectral Approach to Non-Isotropic
Turbulence Subjected to Rotation," Journal of Fluid Mechanics. Vol.
202, 1989,pp. 295-317.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

5
Moreau, R., "On Magnetohydrodynamic Turbulence," Proceedings
Symposium on Turbulence of Fluid and Plasma, Polytechnic Institute
of Brooklyn, Brooklyn, NY, 1968, pp. 359-372.
6
Craya, A. "Contribution i 1'analyse de la turbulence associee %
des vitesses moyennes," These dans Publications Scientifiques et
Techniques, Ministere de 1'Air, Paris, France. 1958.
7
Herring, J. R., "Approach of Axisymmetric Turbulence to
Isotropy," Physics of Fluids. Vol. 17, 1974, pp. 859-872.
8
Reynolds, W. C., "Effect of Rotation on Homogeneous Turbulence,"
10th Australian Fluid Mechanics Conference, Melbourne, Dec. 1989.
9
Kida, S.f and Hunt, J. C. R., "Interaction Between Turbulence of
Different Scales over Short Times," 201, 1989, Journal of Fluid
Mechanics. Vol. 201, 1989, pp.411-445.
10
Jacquin, L., Leuchter, O., Cambon, C., and Mathieu, J.,
"Homogeneous Turbulence in the Presence of Rotation," Journal of
Fluid Mechanics. Vol. 220, 1990, pp. 1-52.
11
Cambon, C., "Spectral Approach to Axisymmetric Turbulence in a
Stratified Fluid,"Advances in Turbulence. Vol. 2; Edited by H. H.
Fernholz and H. E. Fielder, Springer Verlag, Berlin, 1989.
12
Gamier, M., Alemany, A., Sulem, P. L. and Pouquet, A.
"Influence of an External Magnetic Field on Large Scale Low Magnetic
Reynolds Number MHD Turbulence," Journal de MScanioue. Vol. 20, N°
2, 1981, pp.233-251.
Inverse Cascades Generated by Alpha Dynamo
and Anisotropic Kinetic Alpha Effect
P. L. Sulem* and B. Galantit
Observatoire de la Cote d'Azur, Nice, France
and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

A. D. Gilbert^
Cambridge University, Cambridge CBS 9EW, England

Abstract
Like the anisotropic kinetic alpha effect, alpha dynamo can drive a (nonlin-
ear) inverse cascade when an extended range of scales are linearly unstable.
These cascades are related to the existence of unstable steady solutions of
the equations governing the large-scale dynamics. Furthermore, the inter-
action of the two instabilities can lead to periodic or chaotic dynamos.

The alpha effect refers to the situation where a small-scale helical flow,
acting on a large-scale mean magnetic field, generates a mean electromotive
force (emf) that is a function of this field itself.1 When the small-scale flow
is helical,2'3 or otherwise lacks parity invariance,4 the effect may lead to
dynamo action with exponential growth of a seed large-scale magnetic field.
The evolution of dynamo instabilities in extended systems, where mag-
netic fields over a wide range of scales are linearly unstable, was addressed
within turbulence closure schemes, and an inverse cascade of magnetic en-
ergy and helicity to large scales was found.5 This process was corroborated
by direct numerical simulations of the MHD equations6 and illustrated by
phenomenological models.7 An implication of this result is that the scale
of a magnetic field observed in an astrophysical object may be much larger
than that of the linearly most unstable magnetic modes.
More recently, we have shown that an inverse cascade of magnetic field
is not confined to turbulent MHD and can occur in much simpler situations,
such as in a low-Reynolds-number flow driven by a given body force f(x, t)
(Refs. 8 and 9). In this case, equations for the evolution and saturation of
dynamo instabilities can be derived systematically using a multiple-scale
expansion. A simple example of such a force is
f(x,*) = [/sin(Aj 0 aj),/cos(Aj 0 y),/sm(A;ox) -f /cos(Aj 0 2/)] (1)

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Laboratoire Giovanni Domenico Cassini; also at School of Mathematical Sci-
ences, Tel-Aviv University, Israel.
t Laboratoire Giovanni Domenico Cassini.
J Department of Applied Mathematics and Theoretical Physics.
146
INVERSE CASCADES 147

which can be viewed as the sum of two Beltrami waves. At low Reynolds
number, and in the absence of large-scale magnetic or velocity fields, this
steady force generates a simple helical cellular flow, introduced in Ref. 10
to illustrate alpha effect dynamo action. In the linear kinematic dynamo
theory, the most unstable large-scale magnetic field modes depend solely
on the z coordinate. Here, we examine only such one- dimensional fields,
U = U(M) and B = B(z,t), with J73 = B$ = 0. In this case the saturation
of the instability results from the modification of the alpha effect and mean
stresses because of a large-scale field or flow. Furthermore, the body force
given in Eq. (1) does not induce any large-scale instability for the large-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

scale velocity; any initial large-scale flow decays by viscosity, and we may
take U = 0. Assuming a unit Prandtl number for simplicity, we are led to
the following dynamo equations for the large-scale magnetic field:

= -adz [B*/(Bl + I) 2 ] + d*zBl (2a)


2 2
+ I) ] + d zB2. (26)

Here we have made a convenient change of variables: Z = z/L, T = rjt/L2,


B = B/&077, where the length L is a characteristic scale of the domain
in which the MHD evolution is taking place, and TJ denotes the magnetic,
diffusivity. Equations (2) contain a unique parameter a = R^Lkg, which
controls the number of linearly unstable magnetic field modes contained
within the domain, being the ratio of the size of the domain to the scale of
neutrally stable magnetic field modes. Here the magnetic Reynolds number
Rm — u/rjko is constructed with the amplitude u = //f/fco °f the helical
cellular flow generated in the absence of large-scale fields.
We have simulated Eqs. (2) in a domain that is 2-rr periodic in Z,
with a == 20, starting with a weak magnetic field whose integral over the
entire period is zero, which is conserved during the evolution. At early
times the behavior is described by the linearization of Eqs. (2), which are
simply equations for alpha effect kinematic dynamo action in the basic flow
studied in Ref. 10. Magnetic modes whose wave numbers lie in the range
0 < \k\ < 20 are unstable. We observe that, at early times, the magnetic
field is dominated by the growth of the modes with \k\ — 10, which are the
most unstable of the range. At larger amplitudes of the magnetic field the
non linear terms in the equations become important, and energy is trans-
ferred to larger and larger scales in an inverse cascade. This dynamics is
visible in Fig. la, which displays the energy spectrum Ek of the large-scale
magnetic field at various times. The times T = 0.04 and 0.08 correspond
to the linear phase. In physical space, the inverse cascade corresponds to a
reduction of the number of peaks in the magnetic field components by pro-
cesses of merger and destruction. Eventually, the magnetic field saturates
in a stationary configuration dominated by the largest available scale. For
this latter solution the Fourier modes of even wave number vanish, a prop-
erty reflecting a symmetry B(Z -f TT) = — B(Z) of the solution in physical
space. The details of the inverse cascade are seen in Fig. Ib. The energy Ek
of the large-scale magnetic modes is plotted against time for k = 1, ..., 10.
Log-log coordinates have been used to make the first steps of the cascade
more visible. We observe the existence of increasingly long plateaus for
the successive dominant modes. The existence of these metastable states
suggests that during the inverse cascade the system is attracted by a se-
quence of steady solutions with the form of hyperbolic fixed points, stable
to perturbations of smaller scales and unstable to perturbations of larger
scales.
148 P. L SULEMETAL
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

50 100 150 200 250


T

Fig. 1 : Inverse cascade of magnetic field for the body force (1) and a = 20: a)
magnetic energy spectrum at various times; b) evolution of the magnetic
energy in Fourier modes \k\ = 1,..., 10 (log-In coordinates).

The existence of a family of steady cellular solutions for Eqs. (2), of


spatial period 2w/n with n < a, holds for a large class of body forces; it
results from the phase-space structure and scaling properties of the steady
solutions. These are shown to obey a two-dimensional autonomous Hamil-
tonian system (with formal "time" variable r = aZ) that possesses closed
orbits, corresponding to steady space-periodic solutions of Eqs. (2). Evi-
dence that solutions of the evolution problem approach the previously given
steady cellular states together with the stability properties of the latter are
discussed in Ref. 9.
A transport effect analogous to the alpha effect also exists in ordinary
hydrodynamics. When a large-scale velocity field is superposed on a small-
scale flow lacking Galilean invariance, the mean Reynolds stress can be a
function of the large-scale flow. This is referred to as the anisotropic kinetic
alpha (AKA) effect, and may lead to the growth of the large-scale flow when
the original small-scale flow is anisotropic and lacks parity invariance.11 In
the case of small-scale flows of low Reynolds number, the equations for the
large-scale mean fields can be derived in closed form using multiple-scale
analysis.11 When velocity fields are unstable over a large range of scales, a
kinetic inverse cascade analogous to that described earlier, also develops.12
A simple example of a body force that, when stirring a conducting
fluid, can simultaneously lead to an AKA and a dynamo instability, is

= / cos(Aj 0 y 4- /2 =

/3 = (/i + /a) (36)

Using the same units as those given earlier with, in addition, U =


the equations governing the large-scale velocity and magnetic field (as-
INVERSE CASCADES 149

a)
^T =15.00

v
' t ' = 0.65''' -v''\'X-^ : X\ ..*'
--•''
-i<"'_ r> oc ^x
A» " / / V\ , '.'«,••••£.
'.'',!..>
"f = 0.25
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

'°0ro A: iog r o 7*

Fig. 2 : Evolution of the magnetic energy E$* and kinetic energy modes E% for
the body force (2): a) periodic regime for a = 30.5; b) chaotic regime for
a = 62.5. |*| = 1 (-), |*| = 2 (- - -), |*| = 3 (-.-.-), |*| = 4 (.....) iand
|*| = 5 (-...-).

sumed to depend only on the z coordinate) are

£2

(46)

- (U2 + I) 2 - 1
(4c)

B\ - fa -I)2-I
_ !)2 ' ~*v< '

The parameter a, controls the number of linearly unstable magnetic and


velocity field modes. An inverse cascade of velocity and magnetic field is
observed when an extended range of scales are unstable. Although in the
absence of large-scale flow the magnetic field always stabilizes to a steady
state, a variety of long-time behaviors are obtained for the solutions of Eqs.
(4) when a is increased. As an illustration, Fig. 2a shows a periodic regime
achieved with a = 30.5, which displays sudden transitions between time
intervals of quiescent evolution and bursts of violent oscillations. The two
behaviors may also be present in regimes where the solution appears chaotic
(Fig. 2b). In this paper, we restrict our attention to large-scale fields
depending on a unique Cartesian coordinate. For general three-dimensional
solutions, direct coupling between the large scales may be efficient and
lead to other saturation mechanisms. In this context, an inverse cascade
of velocity field was obtained in the case of pure AKA instability.13 The
extension to MHD flows is now under investigation.

References
1
Parker, E.N. "Hydromagnetic Dynamo Models," Astrophysical Jour-
nal, Vol. 122, 1955, pp. 293-314.
150 P. LSULEM ETAL.

2
Steenbeck, M., Krause, F. and Radler, K.-H. "A Calculation of the
Mean Electromotrically Conducting Fluid in Turbulent Motion, under the
Influence of Coriolis Forces," Zeitschrift fuer Naturforschung, Vol. 21a,
1966, pp. 369-376.
3
Moffatt, H.K. Magnetic Field Generation in Electrically Conducting
Fluids, Cambridge Univ. Press, Cambridge, Entgland, 1978.
4
Gilbert, A.D. Frisch, U. and Pouquet, A. "Helicity is Unnecessary for
Alpha Effect Dynamos, but it Helps," Geophysical and Astrophysical Fluid
Dynamics Vol. 42, 1988, pp. 151-161.
5
Pouquet, A., Frisch, U. and Leorat. J. "Strong MHD Turbulence and
the Non-Linear Dynamo Effect," Journal of Fluid Mechanics, Vol. 77, 1976,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

pp. 321-354.
6
Meneguzzi, M., Frisch, U. and Pouquet, A. "Helical and Non-Helical
Turbulent Dynamos," Physical Revew Letters, Vol. 47, 1981, pp. 1060-1064.
7
Kraichnan, R.H. "Consistency of the Alpha-Effect Turbulent Dy-
namo," Physical Review Letters, Vol. 42, 1979, pp. 1677-1680.
8
Gilbert, A.D. and Sulem, P.L. "On Inverse Cascades in Alpha Effect
dynamos," Geophysical and Astrophysical Fluid Dynamics, Vol. 51, 1990
pp. 243-261.
9
Galanti, B., Sulem, P.L. and Gilbert, A.D. "Large-Scale Instabilities
and Inverse Cascades in MHD Flows", Physica, Vol. 47D, 1991, pp. 416-426.
10
Roberts, G.O. "Dynamo Action of Fluid Motions with Two-
Dimensional Periodicity," Philosophical Transactions of the Royal Society
of London, Series A, Vol 271, 1972, pp. 411-454.
n
Frisch, U., She, Z.-S. and Sulem. P.-L. "Large-Scale Flow Driven by
the Anisotropic Kinetic Alpha Effect,'* Physica, Vol. 28D, 1987, pp. 382-
392 12
* Sulem. P.-L., She, Z.-S.. Scholl, H. and Frisch, U. "Generation
of Large-Scale Structures in Three-Dimensional Flows Lacking Parity-
Invariance," Journal of Fluid Mechanics, Vol. 205, (1989) pp.341-358.
13
Galanti, B. and Sulem, P/-L., "Inverse Cascades in Three-
Dimensional Anisotropic Flows Lacking rarity Invariance", Physics of Flu-
ids, Vol. A 3, 1991, pp. 1778-1784.
14
Galanti, B. and Sulem, P.-L., "Large-Scale Instabilities and Inverse
Cascades in Ordinary and MnD Flows at Low Reynolds Number", Larqe-
Scale Structure in Non-Linear Physics, J.-D. Fourmer and P.-L. Sulem eas.,
Lecture Notes in Physics, in press.
Renormalization Group Analysis of MHD Turbulence
with Low Magnetic Reynolds Number
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

S. Sukoriansky,* I. Staroselsky,* B. Galperin,|


S. Roy,$ andS. A. Orszag§
Princeton University, Princeton, New Jersey 08544

Abstract
The renormalization group (RNG) method is applied to study the two-
dimensionalization of three-dimensional (3-D) Kolmogorov turbulence as a
result of influence of the body force that selects some direction. On small
scales we obtain a system of RNG equations for scale-dependent viscosity and
the renormalized Joule dissipation time H"1. Solving these equations we see the
crossover from 3-D to quasi-two-dimensional (2-D) flow: when the wave
number k approaches k,oclf-1/2H3/2 (~e~ is the viscous energy dissipation rate),
the viscosity starts to deviate essentially from its isotropic behavior. We
extend our analysis beyond the crossover region reformulating the problem in
terms of the equation for vorticity. RNG analysis of large-scale properties of
this equation shows that the scale-dependent renormalized viscosity v(k) grows
when k decreases. The renormalization of the value of Joule dissipation time in
this range is absent. We obtain a fixed point of the RNG procedure
corresponding to the self-similar regime with the energy spectrum E(k) °c k'3.
The results are in good agreement with the experimental data.

Analysis
The renormalization group analysis of turbulence has been established
recently as a powerful tool to study a wide variety of turbulent flows.1 This
theory has generated a number of scaling law constants in good agreement with

Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Research Staff Member, Department of Applied and Computational Mathematics.
t Associate Professor, Department of Applied and Computational Mathematics;
currently, Marine Science Department, University of South Florida.
t Doctoral Student, Department of Applied and Computational Mathematics.
§ Professor, Department of Applied and Computational Mathematics.
151
152 S. SUKORIANSKY ET AL.

experiments. In addition, it produces generalized k-e and Reynolds stress


models and subgrid eddy viscosities for large-scale simulations.
This study concentrates on an important practical problem of MHD
turbulence. It is well known that the presence of the magnetic field
dramatically changes the properties of MHD turbulence, making it essentially
anisotropic. On large scales the flow gains the properties of two-dimensional
turbulence. Here we briefly report the results of the RNG approach to both 3-D
and quasi-2-D ranges of MHD turbulence. The results are compared with the
data obtained in experiments on MHD flows with low magnetic Reynolds
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

number.2
For small values of magnetic Reynolds number the flow is governed by
the equation3:

2 2 2
dtt v + (vV)v = -——
p
+ vU
n V v-H n d V~ v + f
U Z

(1)
Vv = 0

Here H^ aB2/p, where a is the electrical conductivity of the fluid, p the fluid
density, and B the magnitude of the magnetic field. The field is assumed to be
oriented along the z axis. The characteristic time of Joule dissipation is
proportional to the value of H"1 and is independent of the spatial scale. The
importance of the H term in Eq. (1) grows with the increase of the flow scale
when the characteristic time of nonlinear interaction increases. For sufficiently
small scales the flow is essentially 3-D. To describe this case we can directly
apply the correspondence principle of Ref. 1 and take the space-time Fourier
transformation of the correlation function of the random stirring force f in
Eq. (1) in the form

< f .(k, Q)f .(k', Q') >= 2DQ k~y R.(k)(27c)d+16(k + k' )8(Q + Q') (2)

k k
ii
Here Pi/k) = 8y --jV is the projection operator, and d=3 the dimension of
space. In the nonmagnetic case the forcing parameter y=d corresponds to fully
developed isotropic turbulence. Using the RNG procedure of Ref. 1, we can
derive the system of differential recursion relations for the scale-dependent
renormalized values of v(k) and H(k):

2 e 1
dv = _D_Q ————————————.^———————————————
_ (3-3a(h) + h-2ha(h)-hh a(h))(6-E-(e-2)h)k~ ~
u
dk 64TC2h2(l + h) v2
RNG ANALYSIS OF MHD TURBULENCE 153

(3b)

Here h(k) = H/(v(k)k2) is the scale-dependent interaction parameter, with


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

a(h) = h'1/2arctan(h1/2). Equations (3) are obtained at lowest order in the


parameter £^4+y-d.
The system of Eq. (3) was solved numerically. Simultaneously we
calculated the renormalized viscosity vis(k) corresponding to the pure isotropic
3-D turbulence.1 Some of the results are shown in Figs. 1 and 2. We found
that the ratio vis/v is the function of the interaction parameter h(k) only
(Fig. 1). The values of v(k) are almost equal to the corresponding values of
vis(k) for h < 1, i.e., for sufficiently large wave numbers. At h(k) ~ 1, which
corresponds to k=k*oc~e~-1/2H3/2 (T~ is the viscous energy dissipation rate), we
see a strong decrease of the viscosity v(k) in comparison with the
corresponding isotropic value vis(k). We regard this qualitative change in the
behavior of vis(k) as a manifestation of the two-dimensionalization of the

—I——I——I
= 1.0 E-4

^
-.2
^

-.4

-.6

-2 0
log (h)
Fig. 1 The ratio of the MHD turbulent eddy viscosity to the one corresponding
to the ordinary hydrodynamk 3-D turbulence vs the scale-dependent interaction
parameter.
154 S. SUKORIANSKY ET AL.

flow. The effective viscosity v(k) is responsible for the flux of energy from
large scales towards the eddies of size ^k"1 and less. When the field is
sufficiently strong (h(k) « 1), this energy flux toward small scales becomes
restricted by the body-force induced effects.
Beyond the crossover region we perform our analysis using the equation
for the z component of the vorticity field co = (curl v)z:

(4)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where the correlation function of the force <|> is readily obtained from Eq. (2):

>=

This equation can be derived from the equations

which were obtained in the study by Gamier et al.4. In equations (4), (5),
v1(x,y,z) is the projection of the velocity field onto the plane normal to the
magnetic field, VJL the gradient in this plane, a,b = x,y, and eab the unit
antisymmetric tensor. In Eqs. (4) and (5) only the terms of the leading order in
k/k« are kept.
Scale elimination procedure for Eq. (4) leads to the set of RNG equations
for scale-dependent values of v(k) and H(k) (here and in what follows we omit
the index _L):
dv ^ AD^-8
=
- l 2 3 2 (6a)

= 0 (6b)
dk

Here the coefficient A = l/(256n) is calculated in the lowest order in the


effective coupling parameter Ar = 512ji(e-l)/5 in conformity with Ref. 4.
In this approximation the solution for viscosity is given by

2/5
RNG ANALYSIS OF MHD TURBULENCE 155

The system of Eqs. (6) can be rewritten in terms of the dimensionless


variables: the nonlinear coupling constant X2 = D1te+1/(H1/2v(k)5/2) and the
interaction parameter h = H/(v(k)k2).
/
— = (e-lA 2 -(5/2)AA, 4 (8a)

dlnh
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

dln(k~V (8b)

The system of Eqs. (8) has the fixed point corresponding to e = 6, where
both the coupling parameter X2 = 512n(e-l)/5 and the interaction parameter are
independent of the limiting value of the spatial scale of the flow k'1.
With the use of Eq. (7) we can evaluate the correlation function of
vorticity and express the kinetic energy spectrum in terms of this correlation
function:
l/5 9-4e
D / 5H- 2 / 5 5

hn = 0.001"

1.5 h n = 0.01
O
ffi

h n = 0.1
so
O

.5 — 0 = 1.0 _I

h n = 10
0
-4 0
log (h)

Fig. 2 The renormalized Joule time H,-1 vs the scak-dependent interaction


parameter. On large scales the renormalized Joule times reaches a constant
value that diffen from the initial one.
156 S. SUKORIANSKYETAL.

Values of Ha/Re x10 3


E,(k) (I) ~ 0
(m 3 s- 2 ) (II) — 3.0
(III) — 9.2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

10

10- 10
10 3 10'
k(rrr 1 )
Fig. 3 Turbulent energy spectra for different magnetic field values.2

Fig. 4 Comparison of the formula (12) with the experiment of Ref. 2 (see also
Fig. 3). E(k) means the values of the energy spectra measured in experiment and
HO is the measured inverse Joule dissipation tune. The solid line corresponds to
the theoretical curve, and the stars correspond to experimental values. It is
important that the renormalization of the effective Joule dissipation time
should be taken into account.
RNG ANALYSIS OF MHD TURBULENCE 157

For the fixed point corresponding to £=6, we find that


E(k) = 65.5D^/5H~2/5k~3, k « k, (10)
There are also some simple dimensional considerations that lead to
Eq. (10). The dimensionality of the value Dj is [DJ = /*-er3. When k « k*,
our problem does not have any characteristic length scale. On the other hand,
we have the parameter H which has the dimension of r1. So we should set e=6
in Eq. (9) in order for [DJ to be spatial-scale independent.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

This t3 spectrum might have nothing to do with the enstrophy cascade


range spectrum. Moreau5 and Alemany et al.6 explain this spectrum as a result
of equilibrium between the angular energy transfer and Joule dissipation.
To specify the parameter of the large-scale force Dl we have to adopt that
at the crossover point k* the small-k asymptotics of the energy spectrum in Eq.
(10) continuously fit the large-k asymptotics, which is almost
Kolmogorovian:

= 0.87e2/3k~5/3, k » k * (n>

Combining relations (10) and (11) with the relation h(k*)=h*, where h* is
the crossover value of the dimensionless magnetic field, we find that
D! = 0.026h/5/2H3

and the expression (10) takes the final form


E(k) = 3.51v2H2k-3 (12)

The formula (12) was used to compare the predictions of this theory with the
experimental data of Ref. 2 (see Fig. 3). The comparison of the theoretical and
the experimental data presented in Fig. 4 demonstrates a very good fit.
Some comments are necessary.
1. It should be realized that our consideration resulting in Eq. (12) is valid
only if k* is smaller than the wave number corresponding to the
Kolmogorov dissipation scale.
2. We derived our expression for the stirring force <|) in the vorticity Eq. (4)
from an isotropic expression for stirring force in the momentum equation,
but the assumption of the force isotropy is not essential. We could as well
take any stirring force with the correlation function which is continuous in
the angular space. It stems from the fact that the main contributions to all
of the renormalized quantities are coming from the region kz/k1«l.)
Acknowledgments
This work was supported by the Air Force Office of Scientific Research under
Contract AFOSR-90-0124, and Defence Advanced Research Projects Agency
under Contract N00014-86-K-0759.
158 S. SUKORIANSKY ET AL.

References
*Yakhot, V., and Orszag, S.A., "Renormalization Group Analysis of Three-
Dimensional Turbulence," Journal of Scientific Computing, Vol. 1, No. 1, 1986,
pp. 3 - 53.
^Sukoriansky, S., and Branover, H., "Turbulence Peculiarities Caused by
Interference of Magnetic Fields with the Energy Transfer Phenomena," Progress in
Astronautics and Aeronautics, Current trends in Turbulence Research, Edited by H.
Branover, M. Mond, and Y. Unger, Vol. 112, AIAA, 1988, pp. 87 - 99.
^Roberts, P.H., An Introduction to Magnetohydrodynamics. Lingsman, 1967.
^Garnier, M., Alemany, A., Sulem, P.L. and Pouquet, A., "Influence of an
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

External Magnetic Field on Large Scale Low Magnetic Reynolds Number MHD
Turbulence. "Journal de Mecanique, vol.20, No.2, 1981, pp. 233 - 251.
*Moreau, R., "On magnetohydrodynamic turbulence", Proceedings of
Symposium on Turbulence of Fluid and Plasma, Polytechnic Inst. of Brooklyn, New
York, N.Y., 1968, p.359.
"Alemany, A., Moreau, R., Sulem, P.L., and Frish, U., "Influence of External
Magnetic Field on Homogeneous MHD turbulence", Journal de Mecanique, Vol. 18,
No. 2, 1979, pp. 277 -313.
Renormalization Group Approach to
Two-Dimensional Turbulence and the
e-Expansion for the Vorticity Equation
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

I. Staroselsky* and S. Sukorianskyt


Princeton University, Princeton, New Jersey 08544

Abstract
Renormalization group (RNG) method is applied to hydrodynamic
turbulence and turbulent transport of passive scalar in two dimensions. The
method uses vorticity as the main dynamic variable. We obtained the solutions
corresponding to the energy and enstrophy cascades. The energy transfer equation
is derived in lowest order of E-expansion and transport coefficients are calculated.

Analysis
Recently developed renormalization group (RNG) theory of three-
dimensional (3-D) turbulence1 allows one to describe quantitatively a wide
variety of flows. This theory is based on the system of incompressible Navier-
Stokes equations and introduces random stirring forces to maintain developed
turbulence. Scale-dependent effective viscosities are computed at the lowest order
of the e-expansion and then the parameters of artificial stirring forces are related
to the parameters of inertial range such as energy dissipation rate. The transport
coefficients and the values of numerical constants calculated in Ref. 1 for 3-D
turbulence are in good agreement with the available experimental data.
Understanding of turbulence in two dimensions is complicated due to
existence of two nontrivial integrals of motion: the energy and the square-mean
vorticity (the enstrophy). The enstrophy conservation prevents cascade of energy
from large scales to small scales because such cascade would increase
enstrophy.2-4 However, the cascade of enstrophy from large scales to small scales
is possible. It results in dissipation of large-scale vorticity at small scales. The
very concept of eddy viscosity becomes more subtle since large and positive eddy

Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Research Staff Member, Department of Applied and Computational Mathematics.
tLecturer, Department of Applied and Computational Mathematics; currently, Mech-
anical Engineering Department, Ben-Gurion University of the Negev, Beer-Sheva,
Israel.
159
1 60 I. STAROSELSKY AND S. SUKORIANSKY

viscosity, which is essential to maintain dissipation of the large-scale vorticity,


would result in robust dissipation of energy on large scales. It was pointed out
by R. Kraichnan5 that two-scale-dependent eddy viscosity has to be introduced to
account properly for both energy and enstrophy transfer in two-dimensional (2-D)
turbulence. In this paper we demonstrate a possible way to calculate transport
coefficients in 2-D turbulence using RNG approach and the Kraichnan's ideas.
In two dimensions the Navier-Stokes and continuity equations for
incompressible fluid can be reduced to the single equation for the pseudoscalar
field co:
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where eab is the unit antisymmetric tensor and the presence of the inverse
Laplasian operator V'2 indicates the nonlocality of Eq. (1), which is in fact
integro-differential. In order to develop the perturbation procedure it is convenient
to present Eq. (1) in the space-time Fourier transformed form:

-
2
= G°(k)eaDK ka J -—JqJqr
D to(q)co(k-q) + G ( k ) f (2)

A A
where k = (k,Q), G°(k) = [ift+Vok2]-1. We introduced the random stirring force f
to the right hand side of Eq. (2), following the correspondence principle of
Ref. 1. This force maintains fully developed nondecaying turbulence for
homogeneous unbounded systems. The force is assumed to be Gaussian with the
correlation function

<f(k,Q)f(kf,Q')> = 2D0k-y+2(27r)3S(k+k')S(a+Q') (3)

where y and D0 are parameters to be specified later. Eq. (2) is defined on the
interval 0 < k < A0.
We can eliminate the fast modes co>(k) with wave vectors satisfying
A0 - 8A0 < k <A0 from the equation of motion (2) for the slow modes co<(k)
with wave vectors from the interval 0 < k < A 0 - 8A0. Following the scale
elimination procedure,1'6 the decomposition of the field co into fast and slow
parts is substituted into the equation of motion (2) and the average is taken over
the fast modes. The correction 8v0 to the bare viscosity v0 at this step of scale
elimination is given by

k-4j> _dq_ (k 2 q 2_ (kq) 2 )(q -2_ |q _ k) -2 )G( £_q)|G(q)l 2 D Q q-y +2

where the symbol J > indicates integration over the band being removed.
RNG APPROACH TO 2D TURBULENCE 161

The integral (4) is calculated in the limit k -» 0, Q, -> 0. This leads to the
differential equation for the function v(k) which is the renormalized value of V0:

dv = A2(e)DQ
d k ~ vV4"1 (5)

where 6 = 2 + y, A2 = (2-E)/(327i). At the lowest order of the e-expansion


A2 = l/(16rc). The solution of Eq. (5) is
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

v(A)=vii—«-^ - (6)

where Ac is a new value of cutoff. The effective constant of nonlinear coupling


X2 = D0A~ce/v3(Ac) is small when e -> 0.
The kinetic energy spectrum E(k) can be expressed in terms of the
correlation function of vorticity U(D(k,Q)=<coa)>(k,Q):

With the use of the expression (6) we obtain from Eq. (7) in the limit of small
A"c''

kl 2£/3
^^o'l^-l '

The energy spectra corresponding to the inertial ranges of 2-D turbulence2'4


can be_pbtained from Eq. (8) by choosing y and D0. The first (y=2, e = 4,
D0 oc e ) corresponds to the inverse energy cascade with the constant energy
transfer rate T: E(k) = CKT2/3k'5/3. The second (y=4, e = 6, D0 oc TI)
corresponds to the enstrophy cascade with the constant enstrophy transfer rate t|:

At this point we must emphasize that the functions v(k) and A2 given by
the Eq. (5) become negative at e = 4, 6. This jeopardizes the possibility of
continuation of the RNG-calculated amplitudes into the region of large e in the
2-D case. Feasibility of such continuation remains to be an open question.
However, let us proceed further and derive the equation for second order moments
(the energy evolution equation). This is done analogously to Ref. 7.
We apply perturbation expansion in powers of e1/2 to the fixed-point
equation of motion.7 Our goal is to derive expression for the energy transfer
162 I. STAROSELSKY AND S. SUKORIANSKY

function to lowest nontrivial order in e. Energy transfer is absent at zero order in


e. At this level the statistically steady state solution of the energy equation E°(k)
is given by formula (8).
In the second order of e-expansion, the energy equation reads

= J J T(k,p,q,t)dpdq (9a)
A
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

k) sin(a) (9b)
kq kp J

where vk = v(k)k2, a is an angle opposite k in the triangle k+p+q = 0 and the


integration domain A is defined by the triangular inequalities lk-pkq<k+p. The
function v(k) evaluated at e = 0 is used to compute the relaxation time
Okpq(t)=(l-e(vkfvp+vq)t)/(vk-i-Vp+vq) entering the energy equation.
Eq. (9) resembles very much the ones obtained through EDQNM-type
closures, but within the RNG approach we derive it self-consistently with no
ad hoc assumptions. For small e the equation is exact. Following the
Kraichnan's ideas we can define now the effective eddy viscosity at wave number
k in terms of the energy transfer through the cutoff Ac:

v(klA c ) = -T(klA c )/[2k 2 E(k)] ;

T(kl A c ) = Jf Jf ' T(k,p,q)dpdq, k < Ac

where the integration J ^signifies p and/or q > Ac. We have omitted the time
argument in function T(k,p,q) assuming that the limit t -^ <» is taken in Eq. (9).
The two-scale-dependent effective viscosity (10) is negative when k « Ac and
positive when k -> Ac (see Fig. 1). This accounts for direct transfer of enstrophy
and inverse transfer of energy, in accord with Ref. 6.
The RNG approach can be useful to describe two-dimensional and quasi-two-
dimensional anisotropic flows encountered in MHD, geophysics and engineering.
The RNG-calculated eddy viscosities and energy/enstrophy transfer functions can
be exploited in sub-grid-scale modeling.
After this study was presented at the Beer Sheva Seminar, A.A. Migdal,
S.A. Orszag and V. Yakhot8 proposed a derivation of the correspondence
principle of Ref. 1 in the case of 3-D flow described by the system of
incompressible Navier-Stokes equations. Applying their approach to the
RNG APPROACH TO 2D TURBULENCE 163
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-1
0.4 0.6 0.8 1
k/A c
Fig. 1 Normalized eddy viscosity v(k|Ac)/|v(0|Ac)| in the 2-D inverse energy
cascade range.

vorticity equation (1) we have been able to demonstrate9 that the zero-mean white
Gaussian random force peaked at some wave number k$ reduces to the algebraic
stirring force (3). The correlation function of this stirring is proportional to
k-y+2, where y=4 for k » 1^ and y=2 for k«k0. In other words, the algebraic
forcing (3) is not merely a good model to describe two inertial regimes in 2-D
turbulence but provides quite natural description of dual cascade2"4 in the presence
of energy input localized at certain scale k^. This further motivates the choice of
vorticity equation to analyze 2-D stirred fluids and turbulence.

Acknowledgments
This work was supported by the Air Force Office of Scientific Research under
Contract AFOSR-90-0124, and Defence Advanced Research Project Agency
under Contract N00014-86-K-0759.

References
t, V. and Orszag, S.A., "Renormalization Group Analysis of Three-
Dimensional Turbulence," Journal of Scientific Computing Vol. 1, No. 1, 1986, pp.
3-53.
2
Kraichnan, R.H., "Inertial Ranges in Two-Dimensional Turbulence," Physics
of Fluids, Vol. 10 , 1967, pp. 1417-1423.
164 I. STAROSELSKY AND S. SUKORIANSKY

^Leith, C.E., "Diffusion Approximation for Two-Dimensional Turbulence,"


Physics of Fluids, Vol. 11 , 1968, pp. 671-672.
^Batchelor, G.K., "Computation of the Energy Spectrum in Homogeneous
Two-Dimensional Turbulence," Physics of Fluids Supplement II, Vol. 12, 1969, pp.
233-240.
^ Kraichnan, R.H., "Eddy Viscosity in Two and Three Dimensions," Journal of
the Atmospheric Science, Vol. 33, August 1976, pp. 1521-1536.
"Forster, D., Nelson, D.R., Stephen, M.J., "Large Distance and Long-Time
Properties of a Randomly Stirred Fluid," Physical Review A, Vol. 16, No. 2, 1977,
pp. 732-749.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

7
Dannevik, W.P., Yakhot, V., and Orszag, S.A., "Analytical Theories of
Turbulence and the e-Expansion," Physics of Fluids, Vol. 30, No. 7, 1987, pp. 2021-
2029.
8
Migdal, A., Orszag, S.A. and Yakhot, V., private communication (1990).
^Staroselsky, I., and Sukoriansky, S. to be published.
Liquid-Metal Flows in Sliding Electric Contacts:
Solution for Turbulent Primary
Azimuthal Velocity
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

G. Talmage* and J. S. Walker t


University of Illinois At Champaign-Urbana, Urbana, Illinois 61801
S. H. Brown$'and N. A. Sondergaard§
David Taylor Research Center, Annapolis, Maryland, 21402
and
H. Branovert and S. Sukorianskyf
Ben-Gurion University of the Negev, Beer-Sheva, Israel

Abstract
Understanding the genera! principles of high-performance
sliding electric contacts, such as those using liquid metals, requires
a thorough knowledge of flow profiles in a narrow gap between a
fixed surface and a moving surface, with a free surface beyond each
end of the gap. The radial and axial velocities in the secondary flow
are reduced by a strong axial or radial magnetic field, For a suffi-
ciently strong magnetic field, the azimuthal momentum transport by
the secondary flow can be neglected. This assumption reduces the
problem for a primary azimuthal velocity to a fully developed mag-
netohydrodynamic (MHD) duct flow problem with a moving wall and
two free surfaces. A typical device can have a Reynolds number,
based on the rotor velocity and radial gap width, well within a tur-
bulent flow regime. Until the advent of a theory that modeled tur-
bulence in the presence of a magnetic field, it was not possible to
explore turbulence in such a system. Now that such a theory exists
we can begin to address the questions concerning the interaction
between the MHD effects and the turbulence. The most critical

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved. The opinions and conclusions are solely those of the authors and
do not necessarily reflect the opinions and conclusions of the U.S. Government.
*Graduate Research Assistant, Department of Theoretical and Applied Mechanics.
t Professor, Department of Mechanical Engineering.
^Senior Research Physicist, Electric Machinery Technology.
§Senior Research Chemist, Electric Machinery Technology.
HLecturer, Department of Mechanical Engineering.
165
166 G.TALMAGEETAL

question is whether the turbulence is damped by a magnetic field


to such an extent that its effect on the velocity and electric potential
distributions is insignificant. If the effects of turbulence cannot be
neglected, then the turbulence model becomes an issue. Here the
question is whether the effects of turbulence can be estimated by
an ordinary hydrodynamic (OHD) model or whether the effects of
the magnetic field must be included. Before attempting to answer
these questions directly, we considered both one-dimensional and
two-dimensional OHD flow problems. The issues that arise in the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

OHD problems also arise in the MHD problems, but their resolution
is simplified in the OHD case. The issues at hand included the
definition of an integral length scale and the procedure for calcu-
lating the turbulent viscosity that must be cast in an appropriate
form; otherwise, numerical instabilities can develop. Once these
difficulties were overcome, our one-dimensional results compared
favorably with Reichardt's data provided that the Reynolds number
was sufficiently large. The two-dimensional results, when taken
along the axial plane of symmetry, were slightly less than their
one-dimensional counterparts. We had anticipated this, given that
the side walls act to retard the flow. Our initial one- and two-
dimensional MHD results indicate that turbulence in the presence
of a magnetic field has a profound effect on both the velocity and
electric potential distributions. However, for the parameter range
currently of interest, an OHD model for turbulence is sufficient. This
is not the case for other parameter ranges. We observed that, due
to turbulent mixing, the magnitude of the velocity is reduced and
the profile flattened for fully turbulent flows as compared to laminar
flows. In MHD flows, a reduction in the magnitude of the velocity is
equivalent to a decrease in the electrical resistance. Therefore, as
a consequence of the velocity decrease, the magnitude of the
current density increases for a given voltage difference between the
rotor and the stator.
Introduction
In a typical high current, low voltage D,C. electromechanical
energy converter, a series of solid conducting disks are mounted
on an axle. Each disk or rotor is encased in a stationary conducting
shroud or stator. The system operates in the presence of a magnetic
field. When a torque is applied to the axle, causing the disks to
rotate, the interaction of the disk velocity with the magnetic field
induces an electric field. Sliding electric contacts between the rotor
and stator complete the electric circuit, and electric currents flow,
crossing magnetic field lines. The interaction between the currents
and the magnetic field creates a Lorentz or electromagnetic body
force that opposes the original disk velocity. This example repre-
sents a generator in which a torque is appliea to the axle and currents
result. For a motor, current is supplied to the system and a torque
is generated.
SLIDING ELECTRIC CONTACTS 167

A major problem is the resistance of the electrical connection


between the rotor and the stator. Conventionally, wire brushes are
used to carry electrical current from the rotor to the stator. For the
extremely large electric currents that develop in certain devices, the
voltage drop with a brush represents a large fraction of the voltage
difference available, Therefore, the system is inefficient. A liquid
metal in the narrow gap between the rotor and stator provides a low
resistance connection making the system more efficient. A sche-
matic of a sliding electric contact with a liquid-metal electrical con-
nection is illustrated in Fig. 1. A,s an example, the eutectic mixture
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

of sodium and potassium (NaK) is a liquid at room temperature, has


an extremely low electrical resistance, forms an excellent electrical
contact with the metal disk tip and stator, and produces very little
drag on the disk because of its low viscosity. When a disk begins
to rotate, the liquid metal is carried by viscous drag to fill the gap
around the entire periphery of the disk. As the speed increases,
centrifugal forces produce a relatively uniform distribution of the
liquid around the periphery. Depending on the orientation of the
device, gravity may cause a perturbation in this uniform distribution.
Unfortunately, at some critical speed, the liquid-metal flow may
become unstable. The liquid metal then flows out of the gap between

Shaft Centerline a

R
Inert Inert
Gas Gas
z = -b z=b

Rotor ,
F.S. r y= i F.S.

Liquid
z = -(a+b) , Metal z - (a-fb)

*~ - -*
y

Stator

Fig. 1 Schematic of a liquid-metal in a sliding electric contact.


168 G.TALMAGEETAL

the disk tip and the stator, and the electrical contact is lost. Several
design objectives are 1) to minimize the voltage drop for a given
load current between the stator and rotor, 2) to minimize the viscous
dissipation and Joule heating in the liquid metal, and 3) to avoid the
instabilities that eject the liquid metal and break electrical contact.
To overcome this problem in which electrical contact is lost, we
must have a thorough understanding of the liquid-metal flows and
their instabilities. Since there are strong magnetic fields present,
the flows are MHD flows. Several analytical studies have greatly
increased our understanding of these flows.1 ~5 Yet, these analytical
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

studies assume that the flow is steady and laminar. Transition from
laminar flow to turbulent flow, for liquid-metal flows in sliding electric
contacts, occurs at roughly the middle of the typical range of angular
velocities. To establish a more complete understanding of the
physics, we must investigate the effect of turbulence on these MHD
flows.
Numerous papers, published over the past two decades, treat
turbulence in electrically conducting liquids with a magnetic field
present and are reviewed by Sukoriansky7
ef a/.6 In a companion
paper, Sukoriansky and Branover proposed a method for calcu-
lating flow characteristics in a turbulent MHD flow. The method was
based on the results of Yakhot and Orszag8 who derived the
transport equations for large scales using the renormalization group
method. Sukoriansky and Branover considered two cases both
involving a one-dimensional, steady state flow between two parallel
plates. One case considered a magnetic field transverse to the plate
walls, and the other a magnetic field parallel to the plate walls. In
both instances, it was possible to integrate the one-dimensional
governing equations to obtain a mean velocity profile. Our effort
extends Sukoriansky's work to two-dimensional rectangular regions
with the magnetic field parallel to the top and bottom walls and
perpendicular to the sidewalls. The geometry precludes the pos-
sibility of integrating the governing equations to determine the mean
velocity and necessitates a re-examination of the integral length
scale Definition. The results indicate that the influence of turbulence
on liquid-metal MHD flows in sliding electric contacts is more dra-
matic than the influence of turbulence on the corresponding OHD
flows.
The Governing Equations and Boundary Conditions
A liquid metal can provide a low-resistance electrical contact
across the small radial gap between the outer tip of a rotating copper
disk (rotor) and a fixed current collector (stator). To compute some
typical values of the dimensionless parameters, we consider a
room-temperature, sodium-potassium eutectic mixture in a radial
gap of 0.1 m m at the tip of a rotor with a radius of 30 cm, as shown
in Fig. 1. Since the rotor radius R is 3000 times the radial gap L
SLIDING ELECTRIC CONTACTS 169

between rotor and stator, we treat the liquid-metal flow as unidi-


rectional, fully developed flow. In other words, we only consider the
primary, azimuthal velocity, and we ignore the transport of azimuthal
momentum by the secondary flow, which involves radial and axial
velocities and which is driven by the axial gradient of the centrifugal
force due to the azimuthal motion. We consider the liquid-metal
flow in the presence of an applied axial magnetic field that is uniform
over the small region occupied by the liquid metal.
The dimensionless governing equations are
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(Ib)
Jz = -'f-
O xT
(1C)

and

42-+4± = °
dy dz cid)
As shown in Fig. 1, z is the axial coordinate, whereas y is radially
inward with the origin at the middle of the bottom stator surface.
Both y and z are normalized by the radial gap L so that the rotor
surface is at y = 1. In Eqs. (1), jy and jz are radially inward and
axial electric current densities normalized by cU0 B0, u is the azi-
muthal velocity normalized by £ / 0 l vt is the turbulent viscosity'
normalized by v 0 , and $ is the electric potential function normalized
by £/ 0 #o Land the Hartmann number

M - B0Ly

Here, a, p, and v 0 are the electrical conductivity, density, and


molecular kinematic viscosity of the liquid metal, respectively, while
U o is the azimuthal velocity of the rotor surface and B0 is the
magnetic field strength. For a typical sliding electric contact, M
ranges from 5 to 50, so the viscous effects are generally significant
throughout the flow domain.
As shown in Fig. 1, the rotor surface lies at y = 1 for - 6 < z < 6,
whereas the three stator surfaces lie at y = 0 for
170 G. TALMAGEETAL

-(a-*- 6) < z < (a + 6) and at £ = ±(0 + 6). The axial gaps


between the rotor sides at z = ± b and the stator surfaces at
z = ± ( a ^ - 6 ) are filled with an electrically insulating inert gas. The
location of each free surface between the liquid metal and inert gas
is one of the unknowns in the secondary flow problem, which also
involves the radial and axial velocities, the azimuthal electric current
density, and the pressure. Since we ignore the secondary flow
problem, we must assume free surface locations, taken here as
y = 1, for 6 < | z | < ( a + 6). We use the values b = 10 and a = 3,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

which are typical of sliding electric contacts.


The rotor and stator are typically copper so that these surfaces
can be treated as perfect conductors at different constant electric
potentials. It is convenient to reduce the governing Eqs. (1) to two
unknowns either by 1) introducing the Ohm's law [(1b), (1c)] into
the conservation Eq. (1 d) or 2) introducing an electric current stream
function and eliminating <jx Whereas these two approaches are
mathematically equivalent, they lead to very different rates of con-
vergence for our successive-over-relaxation finite-difference
numerical solution. The numerical solution converges more quickly
with boundary conditions on values rather than on normal deriv-
atives. Since 89% of the boundary is a perfect conductor where 4>
is known, the solution with the electric potential converges much
faster than the solution with the electric current stream function.
Therefore, we replace Eqs. (1) with

( da\ a ( da\\
v r — U — v r{ — = 0 (2a)

and
»Ji . 0 (2b)
a y 2 dzz
The boundary conditions are
, 2 ) = 0 , f o r | z | < ( a + 6) (3a)
z) = 0, for | z | < ( a + 6 ) (3b)
u(y , ± ( a * 6 ) ) = 0, for 0 < y < l (3c)
^ ( y , ± ( a + 6)) = 0, for 0 < y < l (3d)
u ( l , z ) = 1, for \z\<b (3e)
< t ) ( l , 2 ) = 4. 0 , for | z | < b (3f)
- = 0 for 6 < U | < ( a - « - 6 ) (3g)
dy
SLIDING ELECTRIC CONTACTS 1 71

and
; y ( l , z ) = 0, for 6 < | z | < ( a + 6) (3h)

We have set the electric potential of the stators equal to zero with
no loss of generality. Therefore, $ 0 is the voltage difference between
the bottom stator and rotor, normalized by U0B0L. We have
assumed that the viscous shear stresses in the inert-gas flow is
much smaller than that in the liquid-metal flow,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

The Turbulent Viscosity


The turbulent viscosity is given by

vt = [l + t f ( 0 . 0 2 2 / ? e | V u | / v , - 110)] (4)

where H ( x ) is the ramp function

0 x <0
x *>0

R Q = U o L / v 0 is the Reynolds number, and I is the dimensionless


integral length scale.7& This particular form of theturbulent viscosity
results from the assumption that the Reynolds number is sufficiently
large for an inertial region to exist.
To gain physical insight into the turbulent viscosity formulation,
we present a heuristic argument for Eq, (4). Mathematical deriv-
ations of the turbulent viscosity are presented in papers by Yakhot
and Orszag8 and Sukoriansky and Branover.7 For the duration of
this argument all variables_are dimensional. In the inertial region,
the energy transfer rate, e , is constant and equals the viscous
dissipation rate, e diss

constant = e diss = v 0 | V u :

where u* is the dimensional velocity. Since viscous dissipation


occurs at the smallest scales, | Vu* 1 2 can be related to the Taylor
microscales

|VU*|2 - V2/T]2

with T] the length scale and v the velocity scale. In the microscale
region, the eddy turnover time Ttu = r\ / v equals the characteristic
172 G. TALMAGEETAL

viscous dissipation time t diss = r\2/ v 0 . This implies that


r|/v = r ) / v 0 o r that v = v 0 / T) Therefore, e « v Q / TI 4 which
2

indicates that the energy transfer rate is proportional to the cube of


the viscosity and is inversely proportional to the fourth power of a
length scale.
At this point, we eliminate the small scales. Physically, in a real
flow, the smallest residual scale does not loose its energy due to
dissipation through kinematic viscosity, but rather transfers the
energy to smaller scales. If all smaller scales are eliminated such
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

an energy transfer means a loss of energy for the system. In other


words, this energy loss implies dissipation, Therefore, the turbulent
or effective viscosity is not a real viscosity but rather a means for
expressing the average effect of the eliminated small scales on the
residual scales. During the scale elimination procedure, we move
the dissipation cut-off from the Kolmogorov microscale to a new
cut-off at wave number /c. As a result, the molecular viscosity
changes to a new viscosity which is a function of the cut-off wave
number /c: v k«(e / k 4 ) /3 where k is related to the integral length
scale and v k represents the new effective viscosity. We repeat this
scale elimination procedure until the entire inertial range is elimi-
nated, then the effective viscosity equals the turbulent viscosity.
To account for the fact that the viscosity is purely molecular
when k < /c diS9 we introduce the ramp function

1/3
v, = constants -4
- diss .

This is indeed the form of Eq. (4). The actual constants are based
on work performed by Sukoriansky.
Even in the absence of a magnetic field, the definition of the
integral length scale, which appears in Eq. (4), does pose problems.
The integral length scale reflects the maximum eddy size that plays
a role in the turbulent viscosity. It must take into account the effect
of all surfaces with an emphasis on any corner region. Where the
effects of only one wall are important, / is the distance from that wall
to the point of interest. For our flow, which occurs in a very small
gap, the effects of all four walls are significant at all points, so we
sum the effects of the four walls with

I = JL + JL + 1 + 1
I l l * 1 * I** I*
SLIDING ELECTRIC CONTACTS 173

where Zl is the distance from the rotor face, / 2the distance from the
right stator wall, / 3 the distance from the bottom stator, and Z 4 the
distance from the left stator wall. This gives the form

(5)
y(l-y).
It is important to note that this model of the turbulent viscosity
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

treats the free surface as though it were a solid surface, The free
surface has an impact on the turbulence. In the absence of a
magnetic field, a free surface reduces the integral length scale and
redistributes the turbulent eddies. 9 We have chosen to neglect the
true effects of the free surface on the turbulence,
In the presence of a magnetic field, Sommeria and Moreau^O
demonstrate that for an interaction parameter N significantly greater
than 1 , turbulent eddies are elongated in the field direction. Here,
2
M
pU0 Re

Therefore, these eddies may become anisotropic in nature. Since


strongly anisotropic scales do not interact with the mean flow and
only substantially isotropic scales contribute to the turbulent vis-
cosity, the effect of the magnetic field on turbulent eddies must be
taken into account. For an axial magnetic field the maximal
substantially isotropic length scale can be defined as the
m i n ( / 1 , / 2 ) where

and 12 , which represents a constraint on the isotropy of the integral


length, is such that ___
CjN(lz')+l<2 (7)

where C is a universal constant and

const W / ? e 1 / 3 Z 2 / 3
3
(" (|Vu| 2 ) 1 / 3

In our calculations,
174 G.TALMAGEETAL

In current experimental studies of sliding electrical contacts, the


range of M is 5 -15, and the interaction parameters are not sub-
stantially greater than one. Therefore, the additional constraint used
to ensure isotropy, Eq. (7), never comes into play. The coefficient
C V N ( / ) + 1 is approximately one, and the integral length scale
given in Eq. (6) is approximately equal to the OHD integral length
scale, Eq. (5). Therefore, the simpler OHD integral length scale
definition can be used in place of Eq. (6). Of course, this would not
be the case when the interaction parameters are greater than order
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

one quantities. In future larger machines, the Hartmann number will


be as large as 50, and the full MHD integral length scale would be
required.
Numerical Method
Several numerical issues arise in the solution of turbulent liq-
uid-metal flows in sliding electric contacts that do not arise in the
treatment of their laminar counterparts. The governing equations
[Eqs. (2a), (2b), (4) and (5)] form a nonlinear system of equations
whose solution requires an iterative procedure. For a laminar flow,
the turbulent viscosity would be set to one, and the system of
governing equations, which reduce to Eqs, (2a) and (2b), form two
coupled linear second-order partial differential equations that can
be solved without resorting to iteration. A further issue, that does
not appear in laminar flows, involves the calculation of the turbulent
viscosity. The stability of any iterative scheme depends on how Eq.
(4) is used to update the turbulent viscosity. Approximating the
governing equations [Eqs. (2a), (2b), (4) and (5)] also requires
special attention that is not warranted in laminar flows. Experience
indicates that the shear stress must be continuous across the cell
walls of the discretized flow domain. A continuous shear stress
across cell walls dictates that the Taylor series expansion of the flow
variables be about the center of each cell wall, rather than about the
nodal point at the center of the cell. A final issue entails the dis-
cretization of the flow domain. The mesh can determine whether
the iterative procedure will converge. Even when a convergent
solution is obtained, the solution may not be valid if the mesh has
not sufficiently resolved the large velocity gradients that occur near
the rotor corners and in the boundary layers. With laminar flows,
discretization is not an issue. Uniform meshes perform as well as
graded or nonuniform meshes.
Before describing the iterative process, the finite-difference
approximation to the governing equations needs closer scrutiny.
Once the iterative procedure is formulated, we can examine the
numerical problems associated with the turbulent viscosity calcu-
lation. The discretization of the flow domain is independent of the
iterative process and will be discussed separately. However, one
point must be kept in mind when we distinguish between the iterative
process and the flow domain discretization: the finite-difference
SLIDING ELECTRIC CONTACTS 175

approximations of the governing equations we must take into


account the possibility that the mesh may have nonuniform spacings
in either the axial direction, radial direction, or both.
Along the stators and the rotor, values of the velocity and electric
potential are known, and the turbulent viscosity is set to 1. The
velocity and electric potential in the interior of the flow domain and
along the free surfaces remain to be determined. Since the free
surface is treated as a solid boundary, solely for the purpose of
calculating the turbulent viscosity, here, too, the turbulent viscosity
is set to 1. The geometry of the nodal points differs for those nodes
that are within the interior of the flow domain (interior nodes) and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

those along the free surface (free surface nodes): the interior nodes
have four nearest neighbors whereas the free surface nodes have
three. Two separate finite-difference formulations result as a con-
sequence of these geometric differences. To obtain a finite-
difference formulation, the flow variables ut 4>, and v, are expanded
in a Taylor series about the center of a cell wall and the expansions
substituted into Eqs. (2a) and (2b). For interior nodes, Eqs. (2a)
and (2b), in terms of the Taylor series expansions, are integrated
over a full cell (Fig. 2). For free surface nodes, the same equations
are integrated over a half-cell that is illustrated in Fig. 3. The
boundary conditions along the free surface are applied to the
integrals associated with the free surface nodes. The actual finite-
difference formulation for a given free surface node depends on
whether the free surface node has a stator node as a nearest
neighbor, a rotor node, or free surface nodes on either side.
In the first step of the iterative procedure, a Gauss-Seidel
over-relaxation technique is applied to the finite-difference approx-
imation of Eqs. (2a) and (2b) to determine the velocity and electric
potential throughout the flow domain. At a given node, the
finite-difference approximation of Eq. (2a) is solved prior to the
finite-difference approximation of Eq. (2b). In the second step, the
turbulent viscosity is calculated from Eq. (4) based on the integral

i[

hn/2

D
iI
UP*) he/2 ^ ftk+1) n

hs/2
] '

D
0-1,k)
Fig. 2 Typical full cell.
176 G. TALMAGE ET AL.

(l.k-1) 0,k) P+1)


D n a
i

hs/2
r

*^——————^.^—————————|te~
hw/2 he/2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 3 Typical half cell.

length scale given in Eq. (5) and with the new velocity values used
in | Vu |. The first and second steps are repeated until a simple
convergence criterion is achieved.
There are any number of ways to initiate the iterative procedure,
all of which lead to the same final results. However, the choice of
the initial data set influences the number of iterations required for
convergence. A poor choice could necessitate thousands of
additional iterations. If no prior MHD calculations are available, then
results from an OHD calculation at the given Reynolds number are
used as initial data. Otherwise, given a range of Hartmann numbers
with particular values of the Reynolds number and <|)0, cases
involving successive Hartmann numbers use results based on the
previous Hartmann number as initial data.
We now consider how to update the turbulent viscosity in a
manner that preserves the stability of the iterative scheme. From
Eq. (4), v, is 1 when 0.022/?e 2 1 Vu | 2 / 4 < 110. For
2 2 4
0.022/?e | Vu | / > 110, we might solve Eq. (4) as the cubic
equation

v, 3 -0.022£e 2 V u | 2 / 4 v + 1 0 9 = 0

Unfortunately, for 0.022Re 2 | Vu | 2 / 4 slightly greater than 110,


the physically realistic root of this cubic equation is 9.95. In the next
iteration, with this large V M Eqs. (2a) and (2b) give much smaller
local variations of u. In the subsequent correction of v,, | Vu | is
sufficiently smaller so that 0.0 22 Re2| Vu | 2 / 4 < 110,andv, = 1
SLIDING ELECTRIC CONTACTS 177

again. Thus, the scheme simply alternates between v,= 1 and


v t > 9.95 in the successive iterations with corresponding oscilla-
tions in a at many grid points. The scheme never converges. The
problem associated with this approach is that | Vu | at a point
decreases when the viscosity increases from 1, so that | V u (should
not be considered as a fixed value in Eq. (4). However, the local
shear stress changes only moderately as both v, and | Va (change.
The local shear stress can be characterized by the value of v, | Vu |.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

If we replace the v, on the right side of equation (4) by v f / v l ,


where vt^ is the local value of v, at the previous iteration which
was used to calculate the current value of | Vu |andv, ^represents
the turbulent viscosity at the current iteration, then Eq.n(4) is quartic:

v,4nrw + 1 0 9 v ,n»«w - 0 , 0 2 2 / ? e 2 | Vu | 2 / 4 v t 2_
1 l
IH
= 0 (8)J
V

This equation also has only one physically realistic root, but there
are no discontinuities about the point at which the turbulent viscosity
is activated. Here, too, v, is 1 when
2 2 4
0.022 Re | V u | / v f o i d < 1 10. However, Eq. (8), rather than the
cubic equation, is used to determine v, when
2 2 4 2
0.022Re I Vu I / v old > 1 10. In our calculations, we use an
analytic solution to Eq. (8) to determine the turbulent viscosity.
The discretization of the flow domain is the last issue arising in
the numerical solution of liquid-metal flows in sliding electric con-
tacts that we will treat. Any mesh used to discretize the flow domain
must provide a smooth transition between the laminar boundary
layer and the core region in addition to resolving the large velocity
gradients near the rotor corners and inside the boundary layers. An
extremely fine uniform mesh would satisfy these conditions but is
inefficient. One alternative to a uniform mesh is to use a more refined
mesh in the boundary layer and a coarser mesh in the core region.
Unacceptable truncation errors can be generated at the interface
between the more refined mesh and the coarser mesh that then
propagate throughout the flow domain. A second, more satisfactory
approach is to use a graded mesh that ensures a smooth gradation
over the entire domain.
Graded meshes based on trigonometric functions have proven
successful. Along the radial coordinate, a cosine function creates
a mesh that is highly graded through the boundary layers:

(1 -cosG)
y = —o—
178 G. TALMAGE ET AL

where0e[0, it] is uniformly discretized. In the axial direction, where


a highly graded mesh about the rotor corners is desirable due to
the local nigh velocity gradients, one possible discretization is

for 0<6< (9a)


sin ( j t / p )

and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

6sin(2ji/p-6)
z = 26- ——— , , .
sin(n/p)

-<z<2n/p-sin"l\ ——- s i n ( n / p ) I (9b)


P \ b J
The parameter pcontrols the gradation of the mesh about the rotor
tip corner. If pis too small, A 2 may become too large in the core
region and the free surface region, creating an unfavorable cell
aspect ratio, A y / A z. Additionally, a large A z in the free surface
region indicates that the boundary layers along the side stator will
not be resolved. As pincreases, the mesh becomes more uniform.
A more uniform A z generally will not resolve the boundary layers
along the side stators. Experience has shown that a pof 3 produces
reasonable results. This particular discretization in ogives a rela-
tively coarse mesh near the centerline, z = 0 , where the axial gra-
dients are small. The mesh becomes progressively more refined as
z increases to 6. Finally, as z increases to (a + 6) the mesh
becomes moderately refined. With a moderately refined mesh, the
boundary layer is resolved provided that the number of node points
is sufficiently large. The flow domain extends from z = -(a + 6)to
z = (a + 6 ) . Equations (9) discretize only half this domain,
0<;:<(a + 6). Therefore, to discretize the entire flow domain, the
nodes in the positive ^-direction are reflected about the centerline.
The actual length of the rotor face is calculated such that the rotor
corner lies halfway between the right-most free surface node and
the left-most rotor node (or, equivalently, the right-most rotor node
and the left-most free surface node).
More recently, we have explored meshes based on exponential
functions with highly successful results. In light of the law of the
wall, an exponentially graded mesh may be the most natural mesh
to use. However, only the work performed on trigonometric meshes
will be presented here.
SLIDING ELECTRIC CONTACTS 179

Numerical Results and Discussion


To more fully appreciate the MHD/turbulence interaction we will
begin with a comparison of a laminar OHD flow to a turbulent OHD
flow. The OHD channel is identical to the MHD channel that is shown
in Fig. 1. The velocity contours for Re = 1 and Re = 68,000 are
presented in Figs. 4 and 5, respectively. Because of symmetry, only
the right half of the flow domain appears. The radial gap is 26 times
smaller than the length of the bottom stator. To resolve the velocity
contours, the radial axis has been stretched by a factor of 10.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

A Reynolds number of 1 is representative of laminar flow. In the


core region of Fig. 4 the velocity profiles are typical of Couette flow
where the velocity is proportional to y. As a consequence of our
nondimensionalization, in the core region, the primary azimuthal
velocity is a linear function of y a (y , z) = y. The velocity gradient
in the radial direction is uniform, and there is no variation of the
velocity in the axial direction until the rotor corner is approached.
In the free surface/bottom stator region a balance occurs between
the viscous effects of the Couette flow and the retardation due to
the stators. The stress-free surface plays a passive role.

u = 0.96
——v

u - 0.06

Fig. 4 Two-dimensional OHD case: velocity contour for Re = 1.


180 G. TALMAGE ET AL.

A Reynolds number of 68,000 lies well within the turbulent flow


regime. The velocity profile in the core region is no longer typical
of Cpuette flow (Fig, 5). The central core region has a flatter velocity
profile. Although there is still no axial variation of the velocity in the
core region, the radial velocity gradient is not uniform. Steep velocity
gradients occur along the rotor face and the bottom stator. The
fluid velocity in the free surface/bottom stator region is no longer as
stagnant as it was in the laminar flow. A transfer of momentum has
occurred, with higher velocity fluid parcels having had their velocities
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

decreased and slower velocity fluid parcels having had their velo-
cities increased.
An 1 1 x 53 mesh proved adequate for the laminar flow. Here,
the first number, 11, denotes the number of radial nodes, and the
second number, 53, the number of axial nodes. For turbulent flow,
we chose a 21 x 53 mesh. A 21 x 53 mesh does not fully resolve
the boundary layers but does permit a relatively smooth transition
between vt = 1 and v, > 1. A comparison of our results along the
centerline, z = 0, to Reichardt's results indicates a slight reduction

u - 0.78

u = 0.30

Fig. 5 Two-dimensional OHD case: velocity contour Re


= 68,000.
SLIDING ELECTRIC CONTACTS 181

in turbulence mixing.11 By increasing the number of nodes in the


radial direction we were able to improve the comparison, but such
a large mesh proved infeasible for the turbulent MHD flows, When
the mesh is not sufficiently refined, the derivatives of the velocity are
not accurate. As a result, the estimate of the turbulent viscosity is
poor, and with a poor estimate of the turbulent viscosity, the finite-
difference approximation to the velocity cannot be good. In addition,
a small aspect ratio A y / A z produces a large truncation error. With
turbulent flows, \da/dy\ and \dvt/dy\ are large, and their
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

product larger still. This introduces an additional truncation error in


the Taylor series formulation. A more refined mesh reduces these
errors.
In summary, for OHD flows, turbulence flattens the velocity
profile in the core region and steepens the velocity gradients along
the walls through a momentum transfer mechanism. The effect of
turbulent mixing extends into the free surface/bottom stator region.
For the OHD problem one parameter controls the flow, the
Reynolds number. For liquid-metal flows in sliding electric contacts
with an axial magnetic field present, there are three parameters: the
Reynolds number, which controls the level of turbulence; the
Hartmann number, which controls the strength of the magnetic field;
and <)>0, the electric potential difference maintained across the rotor
gap, which controls the strength of the static electric field. We will
concern ourselves with only the first parameter by considering the
cases Re = 1 and 68,000, with M = 5.0and <|>0 = -5.0,
Again, a Reynolds number of 1 represents laminar flow. The
velocity, electric potential, and current density stream function
contours for Re = 1.0 are presented in Figs. 6 - 8, respectively.
Although there is still symmetry about the centerline,z = 0,the entire

u = 1.2

u = 4.8

u = 0.6
Fig. 6 Two-dimensional MHD case: velocity contour for Re
- 1.0, M = 5.0, and $0 = -5.0.
182 G. TALMAGE ET AL.

--4.8
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

=- 0.6
Fig. 7 Two-dimensional MHD case: electric potential con-
tour for Re = 1.0, M =5.0, and j>o = -5.0.

flow domain is shown, and the y axis is stretched by a factor of 5


rather than 10. The velocity contours in Ficj. 6 do not resemble the
laminar OHD velocity contours given in Fig. 4. For laminar MHD
flows, the velocity in the core region is parabolic in y while remaining
independent of z. The peak velocity is nearly five times the rotor tip
speed. The magnitude of the velocity is an order one quantity in
the free surface/bottom stator region. For <|>0 = -5.0, the current
flows from the bottom stator to the rotor. This current interacts with
the axial magnetic field to create an azimuthal electromagnetic body
force that augments the viscous drag due to the Couette flow. The
electromagnetic body force introduces a symmetry about y = 0.5.
With a sufficiently large Hartmann number, the core flow is domi-
nated by the symmetric electromagnetically driven flow, and the
Couette flow becomes a small perturbation. In a one-dimensional
context these statements can be more readily observed. All the
assumptions made for the two-dimensional MHD flow hold for the
one-dimensional MHD flow: the flow is unidirectional, fully devel-
oped in the presence of an axial magnetic field. However, here,
| 6 | -> «>. The one-dimensional governing equations based on the
Navier-Stokes equations, Ohm'slaw, and the conservation of charge
equation are

= 0
dy

y'y(y) - -
and
0
dy
SLIDING ELECTRIC CONTACTS 183

The boundary conditions are given by u (0) = 0, <t>(0) = 0atthe


bottom stator and u (1) = 1, <|> (0) = <|> 0 at the rotor. The solution
to this system of equations is

3(2<{>0+1)
"(y) - y 1+12/M2
(1-y
1
<t>(y) = y2 y2 2y
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1
2
2 1+ 1 2 / M V 3 2 M2
and

M2/12)

The first term in both the velocity and the electric potential solutions
represents the Couette flow contribution. The second term is
associated with the electromagnetic body force in the velocity
solution and with the induced electric field in the electric potential
solution. The current density is constant. In the example, M = 5.0
and(})0 = -5.0

"(y) = 18.24(l-y)J
2
y
<Ky) = ~ 18.24 L-d--
and
1.46

Fig. 8 Two-dimensional MHD case: current density stream


function for Re = 1.0, M = 5.0, and <j>o = -5.0.
184 G. TALMAGE ET AL.

Even with a relatively small Hartmann number, the flow is dominated


by the electromagnetic body forces. The one-dimensional results
also show that, in the core region of the two-dimensional flow, the
velocity contours in Fig. 5 are essentially one-dimensional: they
display a parabolic dependence on y and the peak velocity is
skewed towards y > 0.53. The flow that occurs in the free surfa-
ce/bottom stator region results from a balance between the viscous
effects of the Couette flow between the rotor and bottom stator, the
electromagnetic body force that drives the fluid in the azimuthal
direction, and the retardation due to the stators.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

The electric potential contours are shown in Fig. 7. The core


region has a one-dimensional nature; the electric potential has a
cubic dependence on y and is independent of z . As | z \ increases
and approaches the rotor corner, the two-dimensional effects
become important. Large electric potential gradients develop near
the free surface/rotor interface. The large gradients are due to the
electric potential difference at the interface. At the side stators, the
electric potential is zero. In the free surface/bottom stator region
the electric potential begins to adjust to this boundary condition by
decreasing in magnitude and developing a ^-dependence. Along
the free surface, where yy is 0, d$/dy and the velocity balance in
accordance with Ohm's law.
The current density stream lines are shown in Fig. 8. In the core
region, j'z = 0 and j y flows from the bottom stator to the rotor. As
previously noted, with y y > 0 throughout the flow domain, the
electromagnetic body force is in the azimuthal direction and aug-

,u = 0.84

-V
u = 0.12 u = 0.42

Fig. 9 Two-dimensional MHD case: velocity contour for Re


= 68,000, M = 5.0, and <f0 = -5.0.
SLIDING ELECTRIC CONTACTS 185

ments the Couette flow. Near the free surface/rotor interface, the
current density lines fringe due to the discontinuity in the electric
potential between the free surface and the rotor.
Increasing the Reynolds number to 68,000 places the flow in a
turbulent regime. Figures 9-11 show the velocity, electric potential,
and current density stream function contours, respectively, for
Re = 68, 000, M = 5.0, and <|>0 = -5.0. The most striking feature
of Fig. 9 is the magnitude of the velocity. It now is an order one
quantity in the core region. The velocity profile is not parabolic, but
instead appears similar to a shear flow profile that has been flattened
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

in the central core region. Near the rotor and bottom stator, intense
velocity gradients develop. The magnitude of the velocity in the free
surface/bottom stator region is much smaller for turbulent flow than
for laminar flow. The reduction in the velocity magnitudes, the
flattening of the velocity profile, and the steepening of the velocity
gradients near the rotor and bottom stator are a result of turbulent
mixing.
Although the electric potential contours in Fig. 10 appear similar
to those in Fig. 7, there are several differences. In Fig. 10, intense
electric potential gradients occur near the free surface/rotor inter-
face. The contour lines are more compressed in the free surfa-
ce/bottom stator region, and as a consequence, the region with zero
electric potential is larger. In the core region, when y < 0.53, for a
given y the magnitude of the electric potential for the turbulent flow
is smaller than the magnitude of the electric potential for the laminar
flow. The situation is reversed for y > 0.53. Furthermore, | d§/ by \
has a smaller value for turbulent flow than for laminar flow, except
possibly near the rotor and bottom stator.
The magnitude of the current density is much larger for turbulent
flow than for laminar flow, as shown in Figs. 8 and 11. We do not

Fig. 10 Two-dimensional MHD case: electric potential con-


tour for RQ s 68,000,/If = 5.0, and *o = -5.0.
186 G.TALMAGEETAL

observe the fringing of the current density stream lines that is


associated with the electric potential discontinuity at the free sur-
face/rotor interface because the magnitude of jy is so large (Fig.
11). Here, too, the current flows from the bottom stator to the rotor.
For a given <|>0l since the current densities are much larger for tur-
bulent flow than for the laminar flow, the effective resistance of the
gap is dramatically reduced by the transition to turbulence. For a
laminar flow, the current density j yproduces a body force that drives
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

a large velocity a. The induced electric field a x z due to this velocity


opposes j y . Therefore, a very large $ 0 may be needed to overcome
the opposition of the induced electric field in order to produce a
given current. Relative to a solid conductor with the same electrical
conductivity, a liquid metal has an electrical resistance that is larger
by a factor of (1 + M2/12) because its flow resists the current.
Transition to turbulence dramatically reduces the peak velocities for
4> 0 = -5.0because turbulent mixing reduces the velocity gradients
away from the boundaries. With a much smaller u, the induced
electric field opposing the electric current is proportionately reduced
and a much larger current can flow. The effective resistance of a
turbulent flow is much less than that of a laminar flow.
The velocities and velocity gradients are generally much larger
for MHD flow than for OHDflow. In the cases that we have presented
here, there is no direct effect of the magnetic field on the turbulence;
N is too small for Eq. (7) to be used and C\/N(l')+ 1 is very close
to 1. With the larger velocity gradients and with no magnetic field
suppression of turbulence, the turbulence is much stronger for the
MHD flows than for the OHD flows, and we can use Eq. (8) for vt

Fig. 11 Two-dimensional MHD case: current density


stream function contour for Re = 68,000, M = 5.0, and
= -5.0.
SLIDING ELECTRIC CONTACTS 187

with the integral length scale given in Eq. (5) at Re = 68, 000 with
confidence in the accuracy.
The effects of <J> 0 would be seen in the direction of the fluid flow.
For a <()0 > - 0.5, current flows from the rotor to the bottom stator,
which results in an electromagnetic body force that opposes the
Couette flow. As a result, the velocity is decelerated, In instances
where the electromagnetic body force is sufficiently strong, the flow
direction may be reversed from the direction of the rotor motion.
For 4) o < - 0.5, current flows from the bottom stator to the rotor. The
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

electromagnetic body force produced by a positive current and


magnetic field acts to accelerate the fluid. Regardless of the value
of <j)0, turbulence would redistribute momentum among the fluid
parcels, causing the velocity profile to flatten and the velocity gra-
dients along the solid surfaces to steepen.

Conclusions
The object of this work has been to resolve two questions
concerning turbulence in electrically conducting fluids with a mag-
netic field present. The more critical question is whether turbulence
is damped by the magnetic field to such an extent that the effect of
turbulence on the velocity and electric potential distributions is
insignificant. The second question concerns the turbulence model
itself. Here we ask whether the turbulence can be estimated by an
OHD model or whether the effects of the magnetic field must be
included.
To answer these questions, we have treated liquid-metal flows
in sliding electric contacts. Although the Hartmann number was
arbitrary, the magnetic field orientation was fixed as axial. Secondary
flows were neglected, and the free surface locations were taken as
y = 1, 6 < | z | < ( a + 6). A successive-over-relaxation finite-
difference code was developed to solve the nonlinear governing
equations which, due to the introduction of turbulence, included
relations for the turbulent viscosity and the integral length scale.
For the sake of comparison, OHD results were included along
with the MHD results. In the laminar OHD case, we observed Couette
flow in the core region with a relatively stagnant free surface/bottom
stator region. In the laminar MHD case, the velocity profile in the
core region was no longer typical of Couette flow, rather the flow
was dominated by the symmetric electromagnetic body force, which
gave it a parabolic profile. The magnitude of the peak velocity, which
occurs in the central region of the core, is five times that of the peak
velocity associated with the OHD flow for M = 5.0 and cj)0 = -5.0.
The free surface/bottom stator region was no longer stagnant: the
velocity magnitude was an order one quantity. When flow enters
188 G. TALMAGE ET AL

the turbulent regime, the velocity profile undergoes a dramatic


change: the gradients in the core region decrease, creating a flatter
velocity profile, and the velocity gradients near the rotor and bottom
stator increase. The change is more radical for the MHD case than
for the OHD case. In the MHD case, the velocity profile is no longer
parabolic and the peak velocity now occurs at or near the rotor
rather than in the central core region. The magnitude of the peak
velocity has decreased by a factor of five.
The answer to the first question is clear from this study. Under
the conditions that we have assumed, turbulence significantly
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

affects the velocity, electric potential, and current density stream


function distributions. Through turbulent mixing, the peak velocities
are reduced from their laminar values. The changes in the velocity
have a profound effect on the electric potential distribution and,
therefore, on the current density.
The question concerning the type of turbulence model to select
is not quite as clear. For our particular parameter range, where the
interaction parameter is an order one quantity, the OHD turbulence
model is sufficient. A comparison of the velocity, electric potential,
and current density stream function contours, using both the OHD
and MHD integral length scale definitions, proved indistinguishable.
For other parameter ranges, the issue is still open, although it
appears that the MHD turbulence model would be required for much
larger Hartmann numbers.

Acknowledg ments
This research was sponsored by Defense Advanced Research
Projects Agency (DOD), Naval Technology Office, Submarine
Technology Program.
References
1
Alty, C. J. N., "Magnetohydrodynamic Duct Flow in a Uniform Trans-
verse Magnetic Field of Arbitrary Orientation," Journal of Fluid Mechanics,
Vol. 48, Pt 3, 1979, pp. 429-461.
2
Hunt, J. C. R. and Williams, W. E., 'Some Electrically Driven Flows in
Magnetohydrodynamics. Part 1. Theory," Journal of Fluid Mechanics, Vol.
31, Pt. 4, 1968, pp. 705-722.
3
Hunt, J. C. R. and Malcolm, D. G., "Some Electrically Driven Flows in
Magnetohydrodynamics. Part 2. Theory and Experiment," Journal of Fluid
Mechanics, Vol. 33, Pt. 4, 1968, pp. 775-801.
4
Hunt, J. C. R. and Stewartson, K, "Some Electrically Driven Flows in
Magnetohydrodynamics. Part 3. The Asymptotic Theory for Flow Between
Circular Electrodes,1 Journal of Fluid Mechanics, Vol. 38, Pt. 2, 1969, pp.
225-242.
SLIDING ELECTRIC CONTACTS 1 89

5
Talmage, G.t Walker, J. S., Brown, S. H., and Sondergaard, N, A.,
"Liquid-Metal Flows in Current Collectors for Homopolar Machines: Fully
Developed Solutions for Primary Azimuthal Velocity," Physics of Fluids, Vol.
A1, No. 7, July 1989, pp. 1268-1278.
6 Sukoriansky, S., Zilberman, I., and Branover, H., "Experimental Studies
of Turbulence in Mercury Flows with Transverse Magnetic Fields,"
Experiments in Fluids, Vol. 4, 1986, pp. 11-16.
7
Sukoriansky, S., and Branover H., "Computational Model of Shear
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Flows with Body Force Induced Turbulence Anisotropy," Journal of Fluid


Mechanics (submitted for publication).
8
Yakhot, V., and Orszag, S., 'Renormalization Group Analysis of Tur-
bulence. I. Basic Theory," Journal of Scientific Computing, Vol. 1, No. 1,
1986, pp. 3-51.
9
Naot, D., and Rodi, W., "Interactions of Turbulent Eddies with a Free
Surface," Liquid-Metal Flows of Magnetohydrodynamics. Vol. 84, edited
by H. Branover, P. S. Lykoudis, and A. Yakhot, Progress in Astronautics
and Aeronautics, AIM, New York, 1983, pp. 98-112.
-t andMoreau, R., "Why, How, and When MHD Turbulence
Becomes Two-Dimensional," Journal of Fluid Mechanics, Vol. 118, 1982,
pp. 507-513.
11 Reichardt, H., "Gesetzmabigkeiten der Geradlinigen Turbulenten
Couette-Stromung," Max-Plank-lnstitut fur Stromungsforschung and
Aerodynamischen, Versuchsanstalt, Gottingen, Germany; Rept. No. 22,
1959.
Anisotropic Turbulence: Analogies Between
Geophysical and Hydromagnetic Flows
C. Henoch* and M. Hoffertt
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

New York University, New York, New York 10003


and
H. Branover$ and S. Sukoriansky§
Ben-Gurion University of the Negev, Beer-Sheva, Israel

Abstract
An analogy is proposed between two fluid dynamic systems which exhibit
similar properties. The analogy was developed, in part, to use experimentally
derived results from one system to help rationalize observations from the other.
The systems are: 1) turbulent liquid metal magnetohydrodynamic channel flow;
and 2) large-scale turbulent oceanic streams. The connection between them is
two-dimensional turbulence theory. This analogy consists of several parts: the
similarity of the forces affecting each flow and, hence, of the governing
equations: the resemblance of critical nondimensional parameters for each case,
and experimental and observational similarity between the two systems.

1. Introduction
The possibility of two-dimensional turbulence in natural phenomena was
postulated, by theoretical considerations, about 25 years ago by Batchelor1,
Kraichnan2 and others. Ordinary turbulence is generally defined by: a) both its
spatial and temporal chaotic three-dimensional motions, and b) its eventual
dissipation turbulent kinetic energy into heat at very small scales by the action
of molecular viscosity. Two-dimensional turbulence, on the other hand, exists
in a flow configuration that is confined, by either boundaries or body forces, or
both, to two dimensions, and is thus restricted from the above-stated
characteristics of three-dimensional turbulence.
Confining turbulent fluctuations to two dimensions results in the
anisotropy of both scales and of the spatial gradients of dynamic quantities such
as velocity. This leads to a profound effect upon the transfer of kinetic energy:
instead of the well-known energy flux from bigger to smaller vortices and the
ultimate dissipation of energy at molecular scales by molecular viscosity which
Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.
All rights reserved.
* Senior Researcher, Department of Applied Sciences.
fProfessor, Department of Applied Sciences.
tHead, Center for MHD Studies.
§Lecturer, Department of Mechanical Engineering.
190
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 191

occurs in homogeneous isotropic turbulence, in the two-dimensional case energy


is transferred in the opposite direction: from small vortices to larger ones. If
there is a continual injection of turbulent energy, there can be an inverse energy
cascade; kinetic energy is transferred toward the large-scale end of the kinetic
energy spectrum. In this way large-scale turbulent structures are enhanced and
they persist over long periods of time. This results in the growth and extended
lifetime of the two-dimensional eddies. The most important practical result of
this is the enhancement of transfer capability of passive scalars, such as heat,
across the two-dimensional plane.
Flows containing these large two-dimensional eddies have been observed in
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

large-scale oceanic and atmospheric streams. On the laboratory scale, they have
been observed in the flows of buoyancy stratified liquids, rotating liquids, and in
the flows of liquid metals under the influence of a uniform magnetic field. In the
latter case, the electromagnetic forces are acting in a manner analogous to the
action of the density gradient stratification in the ocean.

2. Theory
In addition to the forces acting on any fluid in motion (inertial, pressure, and
viscous), there is also a body force acting on a stably stratified oceanic stream
and another, analogous forces, acting on liquid metal magnetohydrodynamic
(MHD) flows. In the former case, the buoyancy force, due to gravity, and to the
stable density stratification of the ocean, acts in the vertical direction. It has the
effect of damping any vertical fluctuation (i.e., turbulence). In the MHD case an
electromagnetic force arises from the interaction between the currents generated
by the fluctuating motions and the magnetic field (Faraday's law). This force
dampens fluctuations in the magnetic fields direction, see Fig. 1.
These body forces appear in the governing equations for both systems. The
Navier-Stokes equation for the stratified ocean case is:

|j-+ (u-v)u = - - VP + vvn; + & |e.dz a>


where U is the velocity field, p density, P pressure, v the kinematic viscosity,
and g(3p/3z)dz is the buoyancy body force.
For the MHD case, the equation is

^JT+ (U-V)U = - - VP + vV 2 U + - jxB (2)

where jxB is the electromagnetic body force.


In Eq. (1) the last term is the buoyancy force term; it is what dynamically
distinguishes stratified from non-stratified flow.
To consider one critical dimensionless parameter for the existence of two-
dimensional turbulence in the stratified flow case, a scaling argument following
Gargett3 is used. The inertial terms can be assumed to scale as w2/h, where h
represents a characteristic vertical length scale, and w is the vertical velocity; h
« 1 (1 is the horizontal length scale) is one of the anisotropic conditions, the
other being that w « u~v, where u and v are the horizontal velocities. The
192 C. HENOCH ET AL.

Z U
particle of

dz
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

ELECTROCONDUCT1NG
PARTICLE

J=E+VXB

Fig. 1 Schematic of forces affecting buoyancy stratified flow and


liquid metal MHD flow.

buoyancy term can be assumed to scale as hN 2 , where


N = (^-^) . The ratio of the inertial to buoyant terms is then
p 3z >
The scale relation u/1 ~ w/h comes from the equation of mass conservation
or continuity
3ii 3v^ 3w
3x 3y ~ 3z
assuming isotropicity of the horizontal velocities (u ~ v) and scales (lx ~ ly ~ 1).
Then the above (w/hN)2 ~ (u/lN)2, which is equal to the Froude number squared.
This analysis allows the determination of a dimensionless parameter Fr2, the
magnitude of which can be seen to describe the relative importance of inertial or
nonlinear (i.e., turbulent) terms to buoyancy ones. The relevant idea of this is
to determine the range for which both of these terms can be considered
significant; this is the region of high inertial terms, i.e., high Reynolds number
— signifying a turbulent flow structure — and a high buoyant force, which
limits any turbulent fluctuations to the horizontal. This regime occurs when
Fr2- O(l) and thus Fr - O(l).
In Eq. (2) an analogous dimensionless parameter is determined by the ratio
of the inertial terms, which scale as u2/!, to the electromagnetic term, which
scales as auBo/p. Then the ratio of electromagnetic force to inertial (i.e.,
turbulent) forces is crB0l/pu = N, the interaction parameter. For both nonlinear
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 193

(ineitial) and magnetic forces to be of equal import, N should be of order 1. This


corresponds to a critical value of N, above which two-dimensional turbulence can
exist, if the Reynolds number, Re = (ul/v) (which is the ratio of inertial to
viscous forces), is high enough to constitute "turbulence." It is clearly seen that
these two dimensionless parameters, the Froude number Fr for the stratified flow
case and the interaction parameter N for the MHD case are derived identically
from the respective governing equations, and that both must be of order 1 to
initiate conditions suitable for two-dimensional turbulence to appear in their
respective cases.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

3 . Experiments and Observations


Recent experiments in liquid metal magnetohydrodynamic (LMMHD)
turbulent flows4 have provided several interesting results relevant to the analogy.
Figure 2 shows sketch of the overall facility and Fig. 3 shows a schematic of the
working channel in which the upper wall was heated isothermally and cross-
stream centerline temperature profiles were measured. Simultaneous wall
temperature measurements were made to provide heat transfer calculations.
These experiments were performed with high Reynolds number mercury flow,
with grid generated turbulence under no-field and high magnetic-field conditions.
If, indeed, two-dimensional turbulent eddies existed under the strong
magnetic-field conditions, we expected to see better "mixing" along the two-
dimensional plane — the plane perpendicular to the magnetic-field lines. Figure
4 shows graphs of the thermal eddy diffusivity, KT, as a function of cross-stream
distance for the no-field and the 0.6T case. This calculation was made using the
heat conduction equation,

3T

I .COOLING WATER

Fig. 2 Schematic diagram of experimental facility.


194 C. HENOCH ET AL

where q is the heat flux at the wall and 3T/3y was calculated from the
temperature profile using a finite difference method.
The plots show that near the top heated wall there is no enhanced diffusivity
in any case (here eddies are constrained by the proximity of the wall) but that
after about 5 mm from the wall the magnetic-field case always exhibits a greater
value of KT due to enhanced turbulence. This enhancement comes from the
presence of vortices whose axes are oriented parallel to the magnetic-field lines;
they are created by the magnetic field. In the channel center the diffusivity is
about 6-10'4 m2/s. For comparison the molecular thermal diffusivity is
4.64-10"6 m2/s, so the enhancement of heat transfer due to cross-stream
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

turbulence is about two orders of magnitude.


In the ocean, large "mesoscale" or synoptic eddies exists with axes parallel
to the gravitational vector. It has been estimated that these large horizontal
eddies in the ocean exhibit a thermal eddy diffusivity of up to 1010 m2/s, about
eight orders of magnitude greater than vertical diffusivities.5 This shows then,
that in both the MHD case and the ocean case, these large, highly anisotropic or
two-dimensional eddies are very important in the transport of heat and indeed of
any other passive scalar field such as dissolved chemical species.
The following results are from dynamic LMMHD experiments where a hot-
film anemometric probe was used in single bar grid generated turbulent channel
flow of mercury under no-field and strong magnetic-field conditions. Figure 5 is
a schematic of the working channel for these experiments.
When turbulence intensities were measured with strong magnetic field
(0.6T) they reached values of almost 50%, even far downstream of the
turbulizing grid, as compared to -10-15% without the field. The eddies
responsible for this turbulence were large-amplitude, low-frequency fluctuations
and could be visualized by an oscilloscope. Figure 6 shows the results from a
calculation of "dynamic mixing length" using a Prandtl type argument that
u
' ~ Imix 9U/3y where u1 are the turbulent fluctuations and dU/3y the cross-
stream gradient of the mean velocity. This graph is for a downstream distance of
15 mm from the grid.

Thermocouples Heating Element


70 cm

2a = 2.65 cm

2b = 5.6 cm
u
B

Fig. 3 Schematic diagram of heat transfer experiment.


GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 195

o B « 0.0 T
• B » 0.6 T

• o

3-
o o
o o
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1- o o

1 2
y [mm]

139
o B-O.OT
119
• B.0.6T
^ 99 •
<s
£ 79-
59-
2
W* 39-

19-

-1
3 4 5 6 7 8 9 1 0
y [mm]

• • I
500-

400-
0 <

300- 0

200-
I o B«O.OT
100- • B.0.6T
5
n
10 20 30 40

y [mm]

Fig. 4 Thermal eddy diffusivity K T vs. cross-stream distance y for


B = O.OT and B = 0.6T.
196 C. HENOCHETAL

= 2,65Cm

2b = 5,6CIH

\J
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Hot Film Probe Single Bar Grid


Fig. 5 Schematic diagram of dynamic experiment.

20

B = O.OT

B = 0.4 T

I 10 H

10 -, 20 30
y [mm]
r

Fig. 6 Dynamic mixing length L m i x vs. cross-stream distance y


for B = O.OT and B = 0.4T.

The calculation gives some indications of the eddy sizes, for field and no-
field cases, under the assumption that uf ~ v' with the field and u' ~ v' ~ wf
without it (it is in theory only u1 and vf which comprise the structure of the two-
dimensional eddies). The graph indicates that even relatively near the grid, at any
cross-stream distance the eddies under the field's influence have length scales up
to two times those without the field. At its maximum, a region of high d\J/dy
exists (i.e., high shear) and the length scales reach about 2/5 of the channel
width.
Again, large, long lived eddies on the scale of the mean flow have
counterparts in the ocean. A well-studied example are the cyclonic eddies which
form along the Gulf Stream. The turbulence associated with the Gulf Stream
probably occurs on many scales, but the large-scale eddies are the cyclonic, cold
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 197
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

e'o

Fig. 7 Satellite data of 15 degree isotherms in the Gulf Stream,


showing meanders and cold core rings (from Ref. 7).

core rings which appear to "pinch off1 from meanders along the main current (see
Fig. 7). These large vortices are referred to as frontal synoptic eddies or rings.
They probably result from baroclinic instabilities6 (of extremely low frequency)
of the strong mean flow, which is the Stream itself, and they reach diameters of
about 200 km (Ref. 7). The scale, then, of these eddies is on the order of the
scale of the mean flow. Surface speeds reach values of ~ 150 cm/s, and their
lifetimes can last 2-3 y (Ref. 8). Gulf Stream rings are a special type of eddy;
they are the most energetic eddies in the ocean.
According to Richardson et al.7 the rings interact with "other rings, with the
Stream, and they generate mean flow; they help drive the Gulf Stream and
enhance its transport."
Calculations made about 20 years ago9 showed that the fluctuating eddy field
could transfer momentum to the mean field; momentum transfer from turbulent
to mean flow is referred to as negative eddy viscosity.10 Figure 8 shows results
from these calculations based on measurements taken near the Gulf Stream in the
ocean. In the figure, A = -puV (3u/3y) is the rate at which the energy of the
large-scale turbulence is converted into kinetic energy of the mean flow. A is
seen to go to negative values between 35 and 70 km offshore; this indicates a
transfer of momentum from the eddies to the mean flow: a negative eddy
viscosity.
198 C. HENOCH ET AL.

In the MHD experiments mean velocity profiles were measured downstream


of the single bar grid, shown as a composite in Fig. 9. It is seen that the wake,
under strong magnetic field, was very long lived as compared to the no-field case.
In the no-field case, the wake behind the obstacle is seen to flatten out and
disperse classically. The turbulence intensity was also very strong and long
lived under the magnetic field; Fig. 10 shows a composite normalized
turbulence intensity profile for three cases: no field, 0.39 T and 0.6 T.
It was decided to look for the effect of negative viscosity as a factor in the
promotion and/or maintenance of the mean flow (the wake). The turbulent
Reynolds stresses, Ty = -pu'iu'j* are proportional to vx (3ui/3xj) (where vx is the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

eddy diffusivity of momentum, and p is the density), according to traditional


closures. Instead of using the Reynolds stresses which were not measured, to
find the diffusivity of eddy momentum (i.e., the "turbulent viscosity") the
following procedure was done. The velocity profiles were graphically integrated
to determine an average velocity value at a fixed x downstream. The two-
dimensional momentum equation was written in terms of the mean variables
du du 1 3P + d V du
u —+
ox
v —
dy
= —T~ T- T~
p ox dy oy
where v is the total (molecular plus turbulent) viscosity. The term involving v
was dropped, and the pressure gradient in x was assessed using Bernoulli's

2 40 r -I960

200 800

16 0 - 64 0

80

4 0 -

Shore X
4 0 x -%6 0 A 80
N *' ^ > \ S
-4 0

Fig. 8 Cross stream profiles of mean flow 7T, momentum flux


and A (dashed line) in the Gulf Stream (from Ref. 9).
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 199

a)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

15 30 120
b)

15 30 60 120

c)

x, downstream, [mm]

120

Fig. 9 Composite normalized velocity profiles downstream of bar


(in scale): a) B = O.OT; b) B = 0.39T; c) B = 0.6T.
200 C. HENOCHETAL

a)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

15 30 60 120

b)

B 0

15 30 60 120

c)

x, downstream, [mm]

15 30 60 120

Fig. 10 Composite normalized turbulence intensity (dashed line,


20%) profiles downstream of bar (in scale):
a) B = O.OT; b) B = 0.39T; c) B = 0.6T.
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 201

equation 0 0
P2 (U 2 x2-U 2 xl)
P
where U represents the average velocity which is a function of x alone. It was
found that if a sufficiently small Ay was used, the error in dv(~Av) was small,
so the term Av/Ay was dropped. The second derivative of u was expanded and
solved at each point by using a finite difference approximation. Figures 11 and
12 show, for different cross-stream distances, the downstream profile of turbulent
eddy viscosity. Value of y/L were chosen to include only regions "within" the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

wake of the grid bar to avoid wall effects.


At each y/L the eddy viscosity, vx, is seen 1) to always be lower for the
enhanced turbulent case than for the no-field case, 2) to be more negative for
small x, distances close to the grid. As a reference point, v molecular is
1.15-10-3cm2/s.
A negative eddy diffusivity of momentum is then evidence of the transfer of
momentum from the fluctuating field (turbulence) to the mean field. Thus, the
situation of the Gulf Stream — where a narrow, strong, long-lived jet-like
current is fed momentum from the large, two-dimensional eddies (created by
stable density stratification) with which it interacts — is very similar to the
MHD wake experiments wherein the long lived wake of the mean flow behind a
single bar grid is fed momentum by the two-dimensional eddies which are created
by the magnetic field.
Another demonstration of the agreement of data on two-dimensional
turbulence with theoretical prediction can be provided by the experimental spectra
of kinetic energy. Since an understanding of turbulence involves the
classification of eddies by their size, the Fourier transform of U(r,t) —> U(k,t)
from position to wave number space is used. Then the kinetic energy per unit
mass E becomes
E = - < U 2 (r, t) > = - J U 2 (k, t)dk = /E(k, t)dk

where brackets refer to mean square values and E(k,t) is the energy spectrum and
characterizes the energy distribution among the different scales of motion.
Energy and enstrophy are related the wave number spectrum by the relations

E = |E(k)dk

fi = Jk 2 E(k)dk

By scaling arguments which follow Kolmogorov11, Kraichnan2 showed that


there must exist, in the wave number or frequency spectrum of energy of two-
dimensional turbulence, two formal inertial ranges.
Kraichnan's hypothesis consists of the following prediction for the form of
the wave number energy spectrum:

E(k) = cte2/3k-5/3 k 0 < k < k inj (3a)


2/3 3
pii k- kinj<k<kd (3b)
202 C. HENOCHETAL.

• B = O.OT

• B = 0.39 T
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

"6 OH

10 20 30 40 50 60 70

x [mm]

0)
1.5

• B-O.OT
1.0-
• B-0.6T

0.5-

0.0

6 -0.5
o

-1.5
10 20 30 40 50 60 70

x[mm]

Fig. 11 Eddy diffusivity of momentum V T vs distance x downstream


at cross-stream values y/L: a) 0.268; b) 0.34.
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 203

B = O.OT

2- B = 0.39T

B*0.6T
1 -
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

0 -
e
o
-1 -

-2-

•3
10 20 30 40 50 60 70

x [mm]

b)
2.5 -


• B-O.OT /*
2.0-
• B-0.39T •*

1.5- • B-0.6T ..•*
..•-'
1.0-
....-•"
^
0.5- ..«••...-••
<N
.•••••
g ...
^
0.0- ..................... ... ...... •••••'
..•••

. . . . • • * • • • • • • • • • • • •

">
-0.5 -
.•-""

10 20 30 40 50 60 70

x[mm]

Fig. 12 Eddy diffusivity of momentum V T vs distance x downstream


at cross-stream values y/L: a) 0.446; b) 0.5.
204 C. HENOCH ET AL

where a and P are dimensionless constants, ko and kd are the low wave number
limit and the high wave number dissipative limit of the spectrum, respectively;
kinj is the wave number of the turbulent energy injection scale; E(k) is spectral
energy density; and k is wave number and is related to an eddy's size I, by
k - 2n/l.
The physical interpretation of these two inertial ranges is that with
continuous forcing or injecting of turbulent kinetic energy at some certain wave
number, kinj, in a two-dimensional turbulent flow, two transfers occur.
Enstrophy (mean square vorticity) is transferred to larger and larger wave number
or smaller and smaller scales. This is referred to as a direct cascade of enstrophy
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

and is manifest by a k'3 slope on the wave number spectrum of kinetic energy in
the inertial range Eq. (3b) above. Energy is transferred to smaller and smaller
wave numbers or larger and larger scales; this is referred to as an indirect or
inverse cascade of energy. Eddies of the scale of the forcing wave number can
coalesce and grow to whatever limiting size constrains them and this is manifest
by a k~5/3 slope on the wave number spectrum of kinetic energy in the inertial
range of Eq. (3a) above. On the wave number spectrum the range (3a) occurs to
the left of the forcing wave number kinj, and range (3b) to the right. These
features of energy transfer of the main hallmarks of two-dimensional turbulence.
Figure 13 shows a spectra12 taken from the ocean near the Gulf Stream.13
The large peaks occurring at the high-frequency end correspond to energy input
from the tides. This energy is dissipated by the normal three-dimensional
Kolmogorov viscous dissipation downscale at molecular levels, as evidenced by
a -5/3 slope. There is then an inertial range with a -3 slope over several decades

frequency, &> (cycles/h )

0.001 0.01
-i——i—rn——
10 - nominal depth-100m
10

10 10

10 id1
10' 10"r2

io"3 I i i 10•r3
1000 100 10
period, T(h)
Fig. 13 Kinetic energy spectrum from the western Atlantic Ocean
which exhibits a -3 range at lower frequencies, indicative of two-
dimensional turbulence, and a -5/3 range at higher than tidal
frequencies, indicative of homogeneous turbulence (from Ref. 13
adapted by Hoffert 12 ).
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 205

of wave number. This corresponds to the direct enstrophy cascade of two-


dimensional turbulence and occurs at very large scales. It may be that baroclinic
instability, that is, internal waves generated by the intersection of isobaric
surfaces with isothermal ones, act as a source of low wave number, for the two-
dimensional turbulence. It has further been suggested6 that the potential energy
associated with the slope of the isopycnal surfaces may serve as a source. In any
case, the -3 slope at low frequencies in Fig. 13 is indicative of an enstrophy
cascade of two-dimensional turbulence.
Figure 14 presents energy spectra taken during the MHD experiments. All
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

spectra were taken at the cross-stream centerline. The field is directed normal to
the plane of the paper. The figure shows four spectra at a 30mm distance
downstream of the grid, taken at four different magnetic field strengths. Several
points are clearly visible. 1) At all downstream distances energy is transferred to
lower frequencies (i.e., lower wave number, and larger scale) as the magnetic
field increases. The slope is near -5/3 in the spectra toward the low-frequency
end; the tendency is for the slope to approach -3 as the field is increased, and an
inverse energy transfer seems to clearly develop. 2) At any downstream distance,
energy injection peaks are visible at the "original" forcing wave number
(corresponding to the grid bar size) when a strong enough field is applied. This
means that eddies in this band of wave number, i.e., of the injection scale, are
unaffected by molecular viscosity for very long time scales. The longevity of
these eddies is further evidence of their two dimensionality.
Viewed in order of ascending magnetic field strength, the trend is for energy
to become more and more highly concentrated at the low-frequency end of the
spectra (frequency is proportional to wave number by the Taylor hypothesis,
k = 27i;f/u, and lower frequency corresponds to a greater scale). This is one
indication of the development of the inverse cascade of energy. Another such
indication is seen by comparing the lines representing of slope -5/3 and -3. In
Fig. 14a, the no-field case, there appears a region where a slope of -5/3 exists
over a wide range of frequencies. This is presumably a region of the classical
Kolmogorov/Richardson initial range for three-dimensional homogeneous
turbulence with a direct energy cascade. In Figs. 14b and 14c, with a magnetic
field strength of 0.22 and 0.4T, respectively, several peaks of energy injection
occur. The first of these peaks occurs at a frequency corresponding to the
injection length scale, the scale of the grid bar. The next few peaks, at higher
and higher frequency correspond to harmonics of this injection scale. These
"harmonic peaks" occur in an inertial range of slope near -3. This is a region of
direct enstrophy transfer for two-dimensional turbulence as discussed above.
This result is interpreted as meaning that a large amount of energy is
concentrated at the scale of the vortices initially created by the bar in a region of
direct energy transfer, these vortices then halve and then quarter themselves so
that peaks of high energy density also occur at frequencies corresponding to 1/2
and 1/4 of the injection scale.
In Figs. 14c and 14d, it is seen that both a -5/3 and a -3 slope appear on the
left and the right side of the injection frequency, respectively. This is in
agreement with the theoretical predictions in that for two-dimensional turbulence
with a steady injection of energy, energy will be transferred inversely (upscale)
on the low wave number side of the forcing k and enstrophy will be transferred
directly to the high wave number side of the forcing k. The wave number at
206 C. HENOCHETAL

1QO U

10-2 -
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

10-4 -

10'6

Frequency [Hz]

b)

C
ID

(ti
15

10-6

Frequency [Hz]

Fig. 14 Evolution of spectral energy with increasing field


strength in cross-stream centerline, downstream at x=30;
a) B = O.OT; b) B = 0.22T; c) B = 0.39T; d) B = 0.6T.
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 207

C)

10-2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

.0
<

10-4

10-6

d)

c
D

100
Frequency [Hz]

Fig. 14 (continued) Evolution of spectral energy with increasing


Held strenght in cross-stream centerline, downstream at x = 30;
a) B = O.OT; b) B = 0.22T; c) B - 0.39T; d) B = 0.6T.
208 C. HENOCH ET AL.

which the transition occurs is (supposed) to correspond to a critical value of the


interaction parameter, N of order 1. This would imply a regime where both
turbulent (nonlinear) and electromagnetic forces are of equivalent import as
discussed above. In both the ocean and the MHD experimental spectra, the
distribution of energy is indicative of two-dimensional turbulence at low
frequencies; these respective spectra provide evidence of theoretical predictions.

4. Discussion
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Results experimentally demonstrated for two-dimensional turbulence in


MHD flows could be extended to large-scale flows in the oceans, based on the
analogous underlying nature of these two systems.
Similarities between the two systems have been shown in the governing
equations and similarity was seen by comparing analogous dimensionless
parameters. Evidence backing the analogy was seen through experimental results
in MHD of enhanced heat transfer, enhanced turbulence, unusual momentum
transfer, and spectral energy distribution; all of which have counterparts in
geophysical regimes.
Further experimentation in MHD would be extremely interesting if all three
fluctuating components of velocity were measured: this could reveal the details
of energy transfer which are not clear at present; and it would have some
relevance to geophysical flows in that the energy injected into oceanic streams
by a variety of sources at a variety of scales is redistributed to turbulent energy,
potential energy, and heat energy.
The tunability of liquid metal MHD experiments with respect to the
strength of the magnetic force, the scale of the channel, and the scale of the
turbulent forcing makes it ideal for investigating specific phenomena in
geophysics. In this work the turbulent wake created behind a single obstacle
turned out to be especially applicable to narrow streams, e.g., the Gulf Stream,
in the ocean.
Geophysical flows exhibit such a vast range of phenomena that a
comprehensive theory is untenable, yet it may be that two-dimensional
turbulence, with its ready reproducibility in the MHD laboratory, can serve as a
model for not only narrow streams in a stratified ocean, but also for explaining
other natural phenomena such as planetary atmospheric flows, and perhaps even
the formation of galaxies and other cosmic events.
References
^atchelor, G.K., "Computation of the Energy Spectrum in Homogeneous Two
Dimensional Turbulence," Physics of Fluids, Supp. 2, Vol. 12, 1969, pp. 233-239.
2
Kraichnan, R.H., "Inertial Ranges in Two Dimensional Turbulence," Physics of
Fluids, Vol. 10, No. 7, 1967, pp. 417-423.
3
Gargett, A.E., "The Scaling of Turbulence in the Presence of Stable
Stratification," Journal of Geophysical Research, Vol. 93, No. C5, 1988,
pp. 5021-5036.
4
Branover, H., Henoch, C., Greenspan, E., Klaiman, D. and Sukoriansky, S.,
"MHD Enhancement of Heat Transfer," Proceedings from the 27th Symposium on
Engineering Aspects of MHD, Reno, Nevada, 1989, pp. 8.221-226.
5
Monin, A.S., and Ozmidov, R.V., Turbulence in the Ocean, D. Reidel,
Dordrecht, Holland, 1985.
GEOPHYSICAL AND HYDROMAGNETIC ANALOGIES 209

6
Gill, A.E., Green, J.S.A., and Simmons, A.J., "Energy Partition in the Large
Scale Ocean Circulation and the Production of Mid Ocean Eddies," Deep Sea Research,
Vol. 21, 1974, pp. 499-528.
7
Richardson, P.L., Cheney, R.E., and Worthington, L.V., "A Census of Gulf
Stream Rings," Journal of Geophysical Research, Vol. 83, No. C12, 1978, pp. 6136-
6144.
8
Wyrtki, K., Magaard, L., and Hager, J., "Eddy Energy in the Oceans," Journal of
Geophysical Research, Vol. 81, No. 15, 1976, pp. 2641-2646.
9
Webster, F., "Measurements of Eddy Fluxes of Momentum in the Surface Layer
of the Gulf Stream," Tellus, Vol. 17, 1965, pp. 239-245.
10
Starr, V., Physics of Negative Viscosity Phenomena, McGraw Hill, New York,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1968.
n
Kolmogorov, A. N., "Local Turbulent Structure in Uncompressible Liquid at
Very Great Re Numbers," Reports of the Academy of Sciences, USSR, Vol. 30, No. 4,
1941, pp. 299-303.
12
Hoffert, M., "The Ocean in 1 Dimension," 3rd International CO2 Cycle
Conference, Hinterzarten, Germany, Oct. 16-23, 1989.
13
Thompson, R., "Topographic Rossby Waves at a Site North of the Gulf
Stream," Deep Sea Research, Vol. 18, No. 1, 1971, pp. 1-19.
Turbulent Electrically-Induced Vortical Flows
V. N. Vlasyuk*
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Vinica Institute of Teaching, Vinica, Ukraine


and
E. V. Shcherbinint
Latvian Academy of Sciences, Riga-Salaspils, Latvia

Abstract
This paper discusses the numerical calculation of the axisymmetric vortical
flow in molten metal induced by an axial electrical current. While the whole
bottom plate of the vertical container serves as an electrode the second electrode
occupies but a small part of the top of the container. An azimuthal
(axisymmetric) magnetic field is self-induced and interacts with the turbulent
flow. A comparison is made between two flow regimes: laminar and turbulent,
and between two mathematical models of turbulence: the k-e model and the k-W
model. The possibility that the circulation level is augmented by the
magnetohydrodynamic turbulence is suggested.

Introduction
Electrically induced vortical flows (EVF) are related to an original class of
flows generated by electromagnetic forces. An electric current has its own
magnetic field. Under certain general conditions an electromagnetic force
emerging due to the interaction of the electric current with its self-magnetic field
becomes vortical, and if the electric current goes through an electrically
conducting fluid, the latter must come into motion.
There is a thorough survey of the work devoted to the theory of electrically
induced vortical flows in the book Electrically Induced Vortical Flows.1 In this
work we will concentrate on typical situations accompanied by the EVF arising
(Fig. la). This situation is also a model of a number of electrometallurgical
furnaces. In a cylindrical container filled with melted metal, the bottom is one
of the electrodes. The second electrode has a smaller cross-sectional area than

Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Docent
fProfessor, Head of Laboratory, Institute of Physics.
210
INDUCED VORTICAL FLOWS 211
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 1 The EVF in a cylindrical container: a) the lines of the


electric current (dashed line), the circulation shape and the velocity
z-component profile (solid line); b) the lines of equal values for the
velocity z-component according to the k-w model (solid line), to
the k-e model (chain-dotted lines), to the laminar computation
(dashed line), the numbers at the curves -10 V/V m .

does the container. Naturally, in the vicinity of the smaller electrode the electric
current will be nonuniform. This condition is sufficient for the EVF initiation
having the shape of a toroidal vortex with the motion directed towards the
decrease of the electric current density in the axial zone.
The electromagnetic force generating the motion of the melt is proportional
to the square of total current I. A non-dimensional parameter of the EVF
S = |o.0I2/47c2pv2 characterizes the ratio of this force and the inertial or viscous
one. Depending on the numerical value of this parameter, different flow regimes
occur. Computations in terms of the laminar modes2 show that, at a small value
of parameter S (S ^ 102 corresponds to the total current in a physical mercury
model I = 0.6A for H/R = 1, ro/R=0.2), the linear Stokes regime of flow occurs
(Fig. 5). In this regime the melt velocity is proportional to the first degree of S
(or to the square of the total current). In the range of 102 < S < 105 the inertial
term of the flow equation plays a more important role: It is a transitional
regime. At S > Kf-106 (I ~ 20-60A) the velocity and other flow characteristics
become proportional to V S or to the first degree of the total current; i.e., the
flow regime becomes nonlinear.
The most significant property of the nonlinear regime is self-similarity of
the velocity field by Vs. In other words, the velocity field normalized by v S
does not change in the core of the flow except for a thin region near the wall.
This property of the EVF is quite convenient for the purpose of modeling
electrometallurgical processes. It is sufficient to investigate the structure of a
non-linear flow and its characteristics on a physical model and then to extrapolate
the results to the industrial current in a geometrically similar real device.
212 V. N. VLASYUK AND E. V. SHCHERBININ

However, the data obtained by the laminar model cannot be extrapolated to


large values of the current. Experiments show3 that if the current I « 100A, the
velocity in the EVF reaches 3xlO-2 m/sec. Thence the Reynolds number
calculated by the radius of the experimental model (R = 3.4xlO'2m) is Re « 104.
At such Re numbers the flow will likely be turbulent.
According to this, there is a necessity to develop turbulent models of the
EVF. Two-parametric models are mostly widespread. The turbulent EVFs were
calculated using the k-e model in Refs. 4 and 5 or the k-W model in Ref. 6. A
comparison of the calculation results according to the k-e model with the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

experiment (see Ref. 7) shows good correspondence, in spite of a rough method


used for the velocity measurements. Recently very thorough measurements have
been made of the EVF velocity in a cylindrical container.3 This provides the
possibility of choosing the most suitable model as well as a numerical
calculation method to investigate the effect of both boundary conditions and the
sensitivity to the variation of constants k and 6, w for the EVF. This is one of
the aims of this work.
Apart from this, the EVF is quite a specific object from the point of view
of the MHD-turbulence theory. First, the electromagnetic force is a source of
motion and, consequently, of turbulence. On the other hand, the azimuthal,
closed at itself and not crossing the flow area boundaries magnetic field
responsible for the electromagnetic force, must interact with the turbulent
disturbance.
It is known that a quasi-azimuthal magnetic field substantially suppresses
the turbulence,8 and the minimum drag corresponds to Ha/Re = 0.8xlO'3. For
the EVF an analogue of the Hartman number is Ha = (|j,0I/27t) Va/pv but in
the nonlinear flow regime Re = (I/v) VpVp. Ratio Ha/Re = VJVJV = Vp
(P is the Batchelor number) is a constant that is independent of the total current.
In particular, for mercury Ha/Re = 4xlO'4; therefore the effect of the turbulence
suppression is not very important for the EVF.
Second, a specific feature of the averaged flow having a toroidal vortex
form in a closed volume is the flow convergence to the symmetry axis in the
part of the container near the smaller electrode, and the divergence in the
remaining part. This is related, for example, to the boundary layer thickness
change with an increase of the current (or of number S) in the mentioned part of
the container.9 Evidently these features of the EVF should exert an influence on
its turbulent characteristics.

Transport Equations
Numerical simulation of the turbulent flow is performed on the basis of a
two-parametric k-z model and the Navier-Stokes equations in terms of an average
velocity vorticity and average stream function (co-\|/) supplemented by two
equations for turbulent kinetic energy k and turbulence parameter Z = km^ where
m, n are constants and I is a scale of turbulence. The scalar field of the turbulent
INDUCED VORTICAL FLOWS 213

viscosity is expressed by k and Z. These differential equations make the average


flow rheology substantially nonlinear.
In the computational procedure for the EVF in a cylindrical container of
radius R, the variables are transformed in a nondimensional form:
... VR f w , coR2 t r
V'=——, V=J~> ®-——> r =—
v vR v R
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

v 2m R n-2m

Henceforth the prime of the nondimensional quantities is dropped.


A general representation of the governing equations in the axisymmetrical
case in the cylindrical coordinate system is given by the following differential
operator*:

Particular expressions for the transport equations are defined by the


corresponding coordinate dependences of parameters a, b, and c.
The nondimensional vorticity transport equation is expressed in the form

L(r 2 ,r 2 ,v e )~ = Qco (1)


r
where

is a source of electromagnetic force curl, Xj/j, is an electric current function that is


the solution of equation

and v e = 1 + v t is the effective intensity. The radial and axial velocity


components (u, v) are found by the use of stream function \jr.
1 dw 1 d\\f
u = — -I-, v = —-1-
r oz r or

which is the solution of


214 V. N. VLASYUK AND E. V. SHCHERBININ

Boundary conditions for Eqs. (1) and (2) are the usual no-slip ones: the radial
velocity is zero, and y and co are at the symmetry axis. Two variants of the
boundary condition for vorticity co and the rigid wall are used: Thorn's variant
and Spolding's variant. The results of Thorn's variant have been found to
correspond better to the actual flow.
In the k-w model instead of second parameter Z turbulence parameter
W = k//2 (m = 1, n = -2) is used, and the turbulent velocity is expressed as
v t = k/w172. The parameters k and w are found by solving the following
equations:
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

L(l,rk,l)k =
(3)
L(l,rw,l)w =iQw,

where Fk=vt/ak, Fw=vt/aw, ak and aw are turbulent Prandtl-Schmidt numbers.


The expressions for the source terms in Eqs. (3) and (4),

Qk = G-C D kw 1/2

contain the dissipative function G:

1 __ /* I v T
I . I ^** I I ** I . •*• I ** T an ^ 2 '
**"

,9z

For k and w the boundary conditions at step h from the wall are expressed
by the tangential stresses at the wall:

1 W ^W

~ C2CDh2

and the above stress is computed according to the formulas of a turbulent flow in
a tube: Tw = 0.0225 V7/4lr1/4. The tangential to the wall velocity v magnitude is
taken at distance h from the wall. On the symmetry axis the nonpermeability
condition is used.
In the k-£ model second parameter Z is replaced by the rate of the turbulent
energy dissipation e = k372// (m = 3/2, n = -1) which is generated by the equation

L(l,re,l)e = iQe (5)


INDUCED VORTICAL FLOWS 215

where Te = vt/cre, and the right side of Eqs. (3) and (5) are defined by expressions
Qk = G-e, Qe = e(C1G-C2e)/k. The turbulent viscosity is given by expression

v t = CDk2/e

The boundary conditions (for e the values at distance h from the wall are
used) are
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where A = 0.42 is the von Karman constant.


In addition to the previously mentioned boundary conditions for k and Z,
computations have been made also under zero conditions at the wall: kr = 0,
Zr = 0. It has been found that, in the case of the k-w model, these give a better
convergence, and vice versa; for the k-e model the convergence is significantly
worse.
The constants for the k-w model are Q = 1.04, C2 = 0.17, C3 = 3.5,
CD = 0.09, ak = aw = 0.9; for the k-e model they are Q = 1.44, C2 = 1.92,
CD = 0.09, a k = 1, ae = 1.3 (i.e., the same as for the plane turbulent flow
computations10). In order to reveal the sensitivity of the results to the choice of
constants, the value of CD is replaced according to

3v
dr
|Vm|

where Vm is the maximum velocity magnitude in the EVF and coefficient K<. is
chosen to be dependent on n

fKco for 0<r<0.25


K c = \ 0.5(1 + sin 2itr)Kco for 0.25 < r < 0.75 (6)
[0 for 0.75<r<0

where K^ is a variable parameter. This scheme will be referred to as a corrected


scheme.11
Equations (3-5) do not account for the Joule dissipation and other possible
effects related to the turbulent parameters.
Samarskij devised a scheme (SS; see Ref. 12) used in the computational
procedure of the laminar EVF; for the turbulent flow the scheme with upwind
differences (SUD) is used. However, it is well known that the SUD introduced a
numerical viscosity that could well exceed the turbulent viscosity and thus
distort the results. An estimate of its influence on the turbulent EVF has been
made by computing the flow according to the k-e model and using a modified
216 V. N. VLASYUK AND E. V. SHCHERBININ

Samarskij scheme (MSS) (see Ref. 11) in the vorticity transport equation. This
scheme is obtained by adding to the central difference scheme a viscous term that
coincides to the second-order accuracy with the difference between the Samarskij
scheme and the central-difference scheme.

Computational Results: Correspondence to the Experiment


Practically all the following results are related to the case of r0/R = 0.2,
S = 108, which permits us to concentrate on the most important aspects.
From Fig. Ib the main difference between the laminar and turbulent flows
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

is evident: the turbulent flow in the axial region is more fused, and the reverse
flow at the side wall of the container is closer to it, if compared to the laminar
flow. There is no significant difference between the results obtained by the k-e
and k-w models both for the field of average velocity and for the maximum
velocity magnitude. This is so despite the fact that the turbulent viscosity
magnitude level is three times higher for the k-w model than for the k-e one
(Figs. 2a and 2b). However, the numerical experiment has shown that, even
when the maximum turbulent viscosity magnitude increases 10 times (that could
be obtained by setting Kco = 0.1 in Ref. 6), the maximum axial velocity
magnitude decreases by only 15%.
From Figs. 2a and 2b, it also follows that there is practically no difference
in the distribution of relative turbulent viscosity vt/vtm. The maximum values
are generated in the vicinity of the maximum velocity gradient values in the
near-axis jet and in the reverse flow region at the side wall of the container (see
Fig. 3). A similar increase of the viscosity takes place at the bottom of the
container, where, as shown in Ref. 9, the thickness of the boundary layer in the
flow diverging from the axis is significantly lower; hence, the velocity gradient
is higher than that in the converging flow at the smaller electrode.

b)

Fig. 2 The isolines of v t : a) according to the k-w model, v t = 89;


b) according to the k-e model, v t m = 30, the numbers at the
curves -10 v t /v tm .
INDUCED VORTICAL FLOWS 217
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-2-

Fig. 3 The profiles of the velocity z-component in the section of


Z - 0.5: according to the laminar computation (chain-dotted lines),
V m = 4.56xl04; according to the k-e model (the SUD, dashed line),
V m = 5.25xl0 4 ; according to the k-e model (the MSS with a
correction, solid line), Vm = 5.46xl04.

Because of an insignificant influence of the model type on the average flow


structure, we will compare the experimental data with the numerical simulation
results using the k-e model. Figure 3 shows the axial velocity profile along the
radius in the plane of Z = 0.5. It is evident that the maximum experimental
velocity magnitude is 12% higher than that computed according to the laminar
theory. The results are closer to the maximum velocity computed with the use
of the SUD; however, a noticeable deviation from the experiment is observable
in the mid-radius region. The MSS scheme with the correction is found to be
the most suitable.
Thus, the computations demonstrate that the two-parametric turbulence
models can be controlled to make those agreeable with empirical results. The
second example of which an adjustment is shown in Fig. 4, where the
distribution of the axial velocity along the r=0 axis is represented. According to
the computations, the maximum velocity is positioned in Z = 0.59 for the
laminar flow; for the turbulent one by the SUS scheme Z = 0.565; by the MSS
scheme with the correction and also according to the experiment Z = 0.43.
218 V. N. VLASYUK AND E. V. SHCHERBININ

v-icf
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

O 0,5

Fig. 4 The distribution of the velocity z-component along the


symmetry axis: according to the k-e model (the SUD, solid line), to
the k-£ model (the MSS with a correction, dashed line).

Fig. 5 The dependence of the maximum velocity z-component at the


symmetry axis on parameter S by the laminar computation (solid
line) and by the k-£ model (dashed line).
INDUCED VORTICAL FLOWS 219

It is of interest to find out the moment when the laminar flow regime
changes into the turbulent. As follows from the data of Fig. 5, the moment
comes practically simultaneously with the beginning of the nonlinear flow
regime, which agrees with the experimental data. According to these data,
within the interval of S = 107-108 the measure of turbulence (Vv' 2 /v) is
approximately 0.5x10~3.

Conclusions
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

The most surprising result of the numerical simulation is a more intensive


circulation level for the turbulent flow (also in the experiment) than for the
laminar one and the displacement of the velocity maximum point within the
turbulent flow toward the small electrode. The latter can be related to the
convergence of the flow to the symmetry axis at the small electrode, yet the
circulation intensity increase needs additional information to explain this
phenomenon.
Of course, the two-parametric turbulence models do not pretend to describe
a fine structure of turbulence, although they permit some quantitative relations
to be revealed. Thus, a connection between the position of the turbulent
viscosity magnitudes and the velocity gradients is evident; the similarity theory
and the direct numerical simulation allows the dependencies of the turbulence
parameters to be established for the EVF in the magnitude of total electric
current (k « I2, E » I3, v t » I), typical velocity (v « I), etc. However, the
construction of the turbulent EVF theory needs both new theoretical approaches
and a thorough experimental investigation of the turbulent characteristics for that
special flow.

References
iBojarevich, V. V., Freiberg, Ja. Z., Shilova, E. E., Shcherbinin, E. V.,
Electrically Induced Vortical Flow, Kluwer, Dordrecht, The Netherlands, 1988.
2
Vlasjuk, V. H., "The Influence of Consumable Electrode Radius on Electrical
Vortex Flow in Cylindrical Container," Magnitnaya Gidrodinamika, Vol. 4, 1987,
pp. 101-103.
3
Chudnovskij, A. Ju., Shcherbinin, E. V., et al., "The Experimental
Investigation of Velocities Field in Axisymmetric Electrical Vortex Flow in
Cylindrical Container," Magnitnaya Gidrodinamika, Vol. 3, 1986, pp. 110-116.
4
Dilavari, A. H., Szekely, J., "A Mathematical Model of Slag and Metal Flow in
the ESD Process," Metallurgical Transactions, Vol. 8B, 1977, pp. 227-236.
5
Dilavari, A. H., and Szekely, J., "Electromagnetically and Thermally Driven
Flow Phenomena in Electroslag Welding," Metallurgical Transactions, Vol. 9, 1978,
pp. 77-87.
6
Kreyenberg, J. and Schwerdtfeger, K., "Stirring Velocities and Temperature
Field in the Slag During Electroslag Remelting," Archives Eigenhiittenwessen, Vol.
50, No. 1, 1979, pp. 1-6.
220 V. N. VLASYUK AND E. V. SHCHERBININ

7
Medovar, B. I., Szekeley, J., Shcherbinin, E. V. et al., "The Comparison of
Results for Physical and Mathematical Modeling of Velocities Field in a Slag Bath at
the ESD," Problemy Spetsial'noj Metallurgii, Vol. 17, 1982, pp. 9-15.
8
Branover, G. G., Vasil'ev, A. S. Gelfgat, Yu. M., Shcherbinin, E.V.,
"Turbulent flow in a plane perpendicular to the magnetic field direction," Magnitnaya
Gidrodinamika, Vol. 4, 1966, pp. 78-84.
9
Shcherbinin, E. V., Jakovleva, E. E. "Electrical Vortex Flow in a Spheroidal
Container," Magnitnaya gidrodinamika, No. 4, 1986, pp. 64-69.
10
Launder, B. E., Spalding, D. B., "The Numerical Computation of Turbulent
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Flows," Computer Methods, Applied Mechanics and Engineering, Vol. 3, 1974, pp.
269-289.
n
Vlasyuk, V. N. "Turbulent Electrical Vortex Flows in a Cylindrical
Container," Magnitnaya gidrodinamika, No. 3, 1988, pp. 76-82.
12
Samarskij, A. A., "The theory of finite difference schemes," Moscow, 1983.
Dissipation Length Scale Dynamics
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

D. Naot*
Center for Technological Education Holon, Holon, Israel
and
N. Yacoubt and D. Maron Moalem$
Tel Aviv University, Ramat Aviv, Israel

Abstract
A transport model equation for the high Reynolds
number dissipation length scale was derived from the
k-e model equations. It was shown that for
homogeneous turbulence energy in local equilibrium in
a fully developed flow, the dissipation length scale
becomes geometric and depends on the field geometry
and boundary conditions only. The choice of the k-e
model coefficients, Cel < 1.5, and CJE > ak, was
reasoned on the background of the dissipation length
dynamics. Some convective flows with different
relative importance of the convection were analyzed
showing that the dissipation length scale always has
an important geometric constituent with an effect of
secondary importance due to the dissipation length
dynamics.
Introduction

Experience in numerical simulation of channel


flows1"4 left the impression2 that, in order to reduce

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
* Associate Professor, Department of Mechanics.
fDoctor, Department of Mechanics, School of Engineering.
tFull Professor, Department of Mechanics, School of Engineering.
221
222 D. NAOT ET AL.

the cost of a parametric study for engineering


purposes, one may consider the replacement of the
dissipation transport model equation by a geometric
specification of a length scale. Here an assessment
of such a possibility is outlined.
To this end the two transport model equations for
the energy and dissipation were used to formulate a
single transport model equation for the dissipation
length scale. With the present formulation, it is
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

possible to identify the physical situation in which


the transport model equation for the length
degenerates to a partial differential equation that
predicts a length scale depending on the field
geometry and the boundary conditions and is
independent of the flow characteristics.
Referring to the solution of this degenerated state
as a universal geometric length scale, it is possible
to study the effects of the various physical
characteristics of the flowfield that deviate from
this ideal state, and base the assessment on such a
background.
A few examples are elaborated: 1) nonconvective
fully developed flow in one-dimensional open channel
and closed duct flows, 2) fully developed flow in two-
dimensional channel flow1 with weak convection due to
the secondary currents, "4 and 3) the convective
recirculation zone formed at the slug front in
horizontal slug flow5'6.
These examples encompass a variety of different
physical situations from the nonconvective fully
developed flow to the convective recirculation bubble,
and represent a wide range of applications.

Length Scale Transport Equation


The model transport equation, discussed here, for
the high Reynolds number dissipation length scale

(1)
DISSIPATION LENGTH DYNAMICS 223

is derived from the model transport equation for the


energy
D ( 3 v 3k ^
— k = —— -^ —— + n - e (2)
Dt Ux.I a,
K
3 x1. )

and the model transport equation for the dissipation


D (3 v 3 e ^ e e.2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

_____ p = I _______
, + Cel. - n - C
e2 , — (3)
Dt Ux.i ae 3xJ k k
In the above equations
D 3 - 3
— « — + U. —— (4)
Dt 3t * 3xi

v - C k2/e (5)

——— 3U.
n = - U.U.
1 3 —^ (6)
3x..
CU , aK , aC , CC ,± , CG «, are the model coefficients and x
is Von Karmann Constant. Ihese are given in Table 1.
A transport equation for the dissipation length is
obtained by multiplying Eqs.(2) and (3) by

C 1/2a 3 C 1/2 a k
~^-T—*- — and -^———s. —2 (7)
X2 2e x2 e

Table 1 Model Coefficients

C C C
, *k *e el s2 X

0.09 1.225 1.225 1.44 1.92 0.40


224 D. NAOT ET AL.

respectively, and substracting the results


D 31
1/4 1/2
C Xk Dt 3x
(8)
1 "*" 2 * 3 "*"

The five parameters of Eq.(8) are


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1/2
(9)
X2

1/2
V*
2 . (z
x
(3.-C
el)) I--
'
do)
f a. 1 f1 3 (X 3k }1
1 --*- -——- P- 7— (ID
X2 I aj Le 9xi Uk 3x.JJ
, i2 rsk i f3k
^ 4 =-^ (12)

and
2.1 (13)

These five parameters express different physical


aspects of the flowfield and illuminate some
characteristics of the k-e turbulence model that are
hidden in its original Eqs.(2) and (3), and are
discussed here.
Geometric Length Scale
The starting point of the present discussion is the
observation that with vanishing five parameters Fi,
i » 1-5; Eq.(8) degenerates for fully developed flow
to the form

+ 1= 0 (14)
3x. 3x.
DISSIPATION LENGTH DYNAMICS 225

expected to yield solutions that depend on the field


geometry and boundary conditions totally independent
of the turbulence energy structure. Some analytic
solutions, as well as numerical solutions for Eq.(14),
are shown here. The importance of these relies on the
observation extended later that Eq.(14) becomes exact
under the assumption of homogeneous turbulence energy
in local equilibrium. It is, therefore, suggested to
refer to the solutions of Eq.(14) as the universal
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

geometric length scale and study the deviations that


are due to the fact that real turbulence is not in
this state.
Eq.(14) was solved analytically for one-dimensional
fully developed channel flow for the following two
cases:
H (n }
1 = - sin -y (15)
n Ip J
for closed duct the width of which is H, and
H ( n }
1.07 - sin ——— yy (16)
n U.07H j
for open channel, the depth of which is H. Eqs.(15)
and (16) were subjected to the vanishing length,
1 = 0, condition at solid walls and to the condition
of 1 = 0 at y = 1.07 H above the open surface as
discussed in Refs. 5 and 7. Two features of the
analytic solutions should be noted: for vanishing
length Eq.(14) becomes dl/dy * ±1 and, hence, adjacent
to the solid walls both Eqs.(15) and (16) predict
1 = y. The maximum values of 1 are: 1 ~ 0.318 H at
the symmetry line of the closed duct, and 1 ~ 0.341 H
at the depth of 0.465 H below the free surface of the
open channel.
Wall Proximity Parameter - FX
Eq.(9) for the parameter F1 contains model
coefficients only and does not depend on the field
characteristics. With nonvanishing value for FI the
numerical value of the coefficients of Eq.(14) may
change, resulting in different but still geometric
226 D. NAOT ET AL.

solutions. The standard model coefficients choice,


however, sets F1 = 0 in order to maintain the 1 = y
behavior of the geometric solutions adjacent to solid
walls where logarithmic profile, together with local
equilibrium turbulence energy, is expected. The
Fl « 0 choice is done in context of the present choice
for the coefficients of Eq.(l) for the length, and the
use of k » v^/NC as boundary condition adjacent to
//
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

solid walls. With these, using Eq.(l) for s, and v*


dU/dy for n, the local equilibrium condition n = 8
turns to an equation for the velocity dU/dy = v*/Xl
and the demand for 1 = y becomes obvious in
logarithmic flow.
Observing Eqs.(2) and (3) it is evident that with
Ccl being different fromC e2 , the possibility was
excluded that the dissipation becomes homogeneous and
in local equilibrium simultaneous with the energy,
thus limiting the scope of the k-e turbulence model.
With the present formulation the equivalent
restriction is the statement that with FX * -1 a
uniform length is not a solution for the geometric
scale.
Spectral Dynamics Parameters F2 and F3
The two parameters F2 and F3 reflect the influence
of the absence of local equilibrium on the dissipation
length scale presumably due to the spectral dynamics
associated with the energy sources.
Eq.(lO) for the parameter F2 depends on the
inequality between the turbulence energy production n,
and its dissipation rate s. Expecting the energy
production to be associated with an increase in the
large eddies population, F2 is expected to become a
source in the dissipation length scale equation in
regions where an excess in energy production prevails.
Indeed, for n > e Eq.(l) predicts such a trend as far
as Cel < 1.5, and the choice of Cel < 1.5 bears
physical interpretation. Still, with the standard
choice of the model coeffcients |F | is limited and
n
is smaller than 0.13 for 0 < — < 2. The effect is not
8
pronounced when compared to the heterogeneity effects
DISSIPATION LENGTH DYNAMICS 227

shown later. Generally, in channel flow away from


walls n/E becomes small and F2 turns to a small sink
for the length in Eq.(8).
Eq.(ll) for the parameter F3 depends on the ratio
between the sources of the energy due to the turbulent
diffusion and the energy dissipation rate. Expecting
the large eddies to contribute to the turbulent energy
diffusion more than the smaller ones, F3 is expected
to become a source for the length scale equation in
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

regions where the shortage in turbulence energy


production and convection is compensated by an
intensive turbulent di ffusion. Indeed, Eq.(11)
predicts such a trend as far as ae > a k , and the
possibility that ae > crk bears physical interpreta-
tion.
Eq.(ll) also explains the sensitivity of the
numerical simulations to the difference ae - ak. Note
that for <re - ak greater than the small value

e - *k > 1 - f-C el - 0.04 (17)

with vanishing production compensated by turbulent


diffusion, |F3 | becomes larger than |F2|. For example,
the recommendation cre - ak = 0.3 may lead in such a
case to |F3 |/|F2 | * ?!5 turning F3 to a significantly
large source.

Energy Heterogeneity Parameters - F4 and F5


Mathematically the two parameters F4 and F5 stem
from the use of the high Reynolds number approximation
for the eddy viscosity for the modelling of the
turbulent diffusion of both the energy and the
dissipation. Physically, these parameters reflect the
influence of the turbulence heterogeneity. Note that
for homogeneous energy both Eqs.(12) and (13) for F4
and F5 vanish. The two parameters do not depend on
the model coefficients and, hence, cannot be tuned by
the choice of these. Estimations for F4 and F5 in
one-dimensional channel flow based on the geometric
scales gradients and an approximation for the energy
k = kB [1 + 2.3 (1-y/B)2] (18)
228 D. NAOT ET AL.

.4

-.4

-.8
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig.l Estimated
heterogeneity
-1.2 parameters for one-
dimensional flow:
-1.6 _____ closed duct,
0 .2 .4 .6 .8 y/B _ _ _ _ _ open channel.

are shown in Fig.l. In Eq.(18) B = H for the open


channel flow, and B = H/2 for the closed duct flow.
By definition in Eq.(12), F4 depends on the energy
gradient only and is always a sink in the length scale
equation. However, F5 also depends on the length
scale gradient and may change sign accordingly.
In both cases of channel flow adjacent to the wall,
these gradients are of different sign and F5 becomes a
sink in the length scale equation. However, away from
the wall in open channel flow, the length scale
gradient alters sign and F5 turns to a source. In
both cases F5, shown in Fig.l, seems to be the largest
among the Fi terms of Eq.(8).
One-Dimensional Channel Flow
Hie overall effect of the dissipation length
dynamics is demonstrated for fully developed channel
flow by comparing dissipation length scale calculated
with Eqs.(2) and (3) with the geometric scale
calculated with Eq.(14). In Fig.2a for closed duct
flow a reduction of about 40%, due to a large sink, is
evident, presumably due to the large and negative
values of the heterogeneity parameter F5 . In Fig.(2b)
for open channel flow a reduction of 20%, due to a
large sink, shows up close to the wall, and an
DISSIPATION LENGTH DYNAMICS 229

a)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

b)
Fig.2 Dissipation
length scale 1/H in
one-dimensional flows:
a) closed duct and
b) open channel.
____ geometric,
calculated
0 .2 .4 .8 y/B with k-s model.

increase of 30%, due to a large source, shows up close


to the open surface, presumably due to the alternating
sign nature of the heterogeneity parameter F5. Still,
the qualitative agreement between the scales is
remarkable.

Two-Dimensional Channel Flow


The effect of the dissipation length dynamics is
demonstrated for fully developed flow in two-
dimensional channel. The length calculated with
Eqs.(2) and (3) in the context of a calculation of
three-dimensional flow that includes secondary
currents using the algebraic stress model given in
Refs. 1 and 2 is compared with the geometric scale
calculated numerically with Eg.(14). In Fig.3a for
closed square duct flow a monotonic large reduction,
similar to that observed in the one-dimensional case,
is shown. The effect of convection of large scale by
the secondary currents from the channel center towards
the corner also shows up. In Fig.(3b) for open
230 D. NAOT ET AL.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig.3 Dissipation length scale 1/H in two-dimensional


flows: a) H x H closed square duct and b) H x 2H open
channel.
_____ geometric, _ _ _ _ _ _ calculated with k-s model.

channel flow, the reduction close to the wall


accompanied by an increase close to the open surface
typical of the one-dimensional case shows up together
with the effect of the convection of large scale by
the secondary currents towards the corner. Similar
effects are also shown in Fig.4 for wide open channel.
Now, however, the agreement between the geometric
scale and the scale calculated by a k-e model is
remarkable. Moreover, note that calculations reported
in Ref. 2 indicate that, although the flow is
dramatically affected by wall roughness heterogeneity,
the dissipation length is almost unaffected and the
scope of this remarkable agreement may be extended.
Horizontal Slug Front
The effects of the dissipation length dynamics is
demonstrated for the convective field formed at the
front of the horizontal slug flow6'7 shown in Fig.5,
by comparing the length calculated with Eqs.(2) and
(3) with the geometric calculated with Eq.(14) shown
DISSIPATION LENGTH DYNAMICS 231
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

IE-

Fig. 4 Dissipation length scale in H x 4H wide open


channel flow:
____ geometric, calculated with k-s model.

y =H

GAS Fig.5 Streamlines


pattern for the
horizontal slug flow,
VTH/v = 457000.
y - o;
x = 0 x = 10H

in Fig.6. In this case the convection is a dominant


factor and the state of fully developed flow is
reached only at the right side of the solution domain
x ~ 10 H. Indeed, at this side both lengths become
similar to the corresponding scales calculated for the
one-dimensional closed duct case. However, at the
recirculation zone an effect of a large source is
evident overruling the intensive convection of small
length scale from the flow origin at the left
232 D. NAOT ET AL.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

8 x/H

Fig.6 Dissipation length scale 1/H in the horizontal


slug flow:
a) geometric and b) calculated with a k-e model.

boundary, presumably due to a large heterogeneity


parameter F5.
Concluding Remarks
Within the variety of the flow configurations
analyzed here, ranging from nonconvective regions with
local equilibrium turbulence to convective regions
with intensive turbulence production, it is suggested
that the pure geometric behavior of the dissipation
length scale is the dominant factor in its evaluation
and that the dynamic processes are of secondary
importance. The general impression is that the
standard k-8 turbulence model, together with the high
Reynolds number approximations, contains a basic
geometric rule for the specification of the
dissipation length scale subjected to relatively small
corrections due to the length scale dynamics.
Simple phenomenologic explanations lead to the
identification of the terms of the length scale
equation that describe spectral processes associated
with deviations from the state of local equilibrium
turbulence, as well as the coefficients of the k-e
turbulence model that quantitatively control these
effects. Now, the choice of Cel somewhat smaller than
1.5, and the choice of a , being somewhat larger than
DISSIPATION LENGTH DYNAMICS 233

tfk, gets physical interpretations. With better


understanding of the spectral effects on the length
scale, one may better tune the k-e model coefficients.
The dominant role of the turbulence energy
heterogeneity effects in forming deviations from the
geometric behavior of the length scale showed up in
all the cases examined. As these effects do not
depend on the k-e model coefficients, it is impossible
to tune their contributions to the calculation of the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

length.
The expectations that the dissipation length scale
calculated for the fully developed channel flows will
show the smallest deviations from the geometric scale
did not come true for the closed duct cases, due to a
monotonous energy heterogeneity effect. Still, the
qualitative agreement is good. For the open channel,
however, the energy heterogeneity effect alters sign
and the deviations from the geometric scale indeed
become small.
On the other hand, the horizontal slug flow case
demonstrates a field with strong convection, strong
energy production, strong diffusion, and steep energy
gradients, each of which may cause a large deviation
between the dissipation length scale and the geometric
scale. Strange, but such large deviations did not
show up, due to the opposing effects neutralizing
each other.
References
1
Naot, D., and Rodi, W., "Calculation of Secondary
Currents in Open Channel Flow," Journal of the
Hydraulic Division, ASCE, Vol.108, HY8, 1982,
pp.948-968.
2
Naot, D., "Response of Turbulent Open Channel Flow
to Roughness Heterogeneity," Journal of Hydraulic
Engineering, ASCE, Vol.110, No.11, 1984, pp.1568-1587.
Naot, D., and Emrani, S., "Numerical Simulation of
the Hydrodynamic Behaviour of Fuel Rod with
Longitudinal Fins," Nuclear Engineering and Design,
Vol.73, 1982, pp.319-329.
4
Naot, D., "Sensitivity of Turbulent Channel Flow
to the Interactions at the Perimeter," Single and
Multi-Phase Flows in Electromagnetic Field, edited by
H. Branover, P. Lykoudis, and M. Mond, AIAA Progress
Series, Vol.100, 1985, pp.202-212.
234 D. NAOT ET AL.

5
Naot, D., Yacoub, N., and Maron Moalem, D., "Open
Surface Boundary Conditions for the k-e Turbulence
Model," Proceedings of the Twenty-Third Congress of
the International Association for Hydraulic Research,
Vol.
6
A, Ottawa, Canada, 1989, pp.293-299.
Yacoub, N., Naot, D., and Maron Moalem, D.,
"Towards the Numerical Simulation of Horizontal Slug
Front," Int. J. for Numerical Methods in Fluids,
Vol.13, 1991 (in print).
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

7
Maron Moalem, D., Yacoub, N., Brauner, N., and
Naot, D., "Hydrodynamic Mechanisms in Horizontal Slug
Pattern," Int. J. Multiphase Flow, Vol.17, 1991,
pp.227-245.
Towards Quasi-Isotropic Algebraic Stress Model
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

for Magnetohydrodynamic Channel Flow


D. Naot* and J. Tannyt
Center for Technological Education Holon, Holon, Israel

Abstract
The applicability of the stress transport models,
as well as the derived algebraic models, to the early
stage of the laminarization of high turbulent Reynolds
number turbulence in magnetohydrodynaadc flows is
further developed here. Numerical solutions for
unidirectional two-dimensional flow subjected to two
orientations of the magnetic field, longitudinal and
transverse, are shown; and the influence of the
diffusional magnetic field fluctuations on the
hydrodynamic turbulence anisotropy was studied. The
application of the results for the algebraic stress
modelling of channel flow is discussed and outlined.
Introduction
The present work is associated with the expansion
of the1 2 applicability of the stress transport
models, ' as well as the related methods such as the
algebraic stress models3 or the subgrid turbulence
models, to magnetohydrodynamic flows.
Experiments showed that magnetohydrodynamic
turbulent flow in channels is characterized by a
combined phenomena of the electromagnetic-induced

Copyright © 1991 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Associate Professor, Department of Mechanics.
fLecturer, Department of Mechanics.
235
236 D. NAOT AND J. TANNY

laminarization4'5 and the enhancement of anisotropy


even to the extent of turning turbulence two-
dimensional.6"8 For a comprehensive literature survey/
see Ref.9. A need for a stress transport model or a
related model is evident. An important step in the
development of these is the study of the set of
theoretical test cases of turbulent shear flows.
Here/ the response of the turbulence anisotropy
typical of unidirectional two-dimensional flow to
diffusional electromagnetic fluctuations is studied/
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

and the possibility to derive an algebraic stress


model for channel flow with weak lateral motion is
outlined.
In developing the present theory1°'11 use was made
of three concepts: local equilibrium/ local
homogeneity/ and quasi-isotropy. Consequently/ the
application of the results is restricted. Still, for
high level turbulence with microscales that are much
smaller than the channel dimensions, these concepts
become plausible; it is suggested11 to distinguish
between the early stage of laminarization of channel
flow high level turbulence/ and the final stage of
laminarization of grid generated low level turbulence/
and apply the present model to the first only.
Subjected to these restrictions/ numerical
solutions for longitudinal and transverse magnetic
fields are presented here and the possiblity of
turning these into an algebraic stress model is
discussed and outlined.
The Stress Transport Turbulence Model
The exact stress transport equations are derived
from the instantaneous linear moraentum equations
3 1 2 1
—at"y + (y-v)y
~ ~ = -p- VP + w ~v +p -~j x~§ (D
where V is the fluid velocity/ P is the pressure/ p is
the fluid density/ v is the kinematic viscosity/ J is
the electric current/ and B is the magnetic flux
density. Each instantaneous variable was replaced by
an averaged variable plus a fluctuation/ and the
stress transport equations were elaborated using an
MAGNETOHYDRODYNAMIC CHANNEL FLOW 237

ensemble averaging12 process

— v. v. + vu —— v.x v. = n.x . + T.13. + P.1] . + L.


1] . + M.
13 . (2)
a*. 1 3 * av 3 3

where
3V.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

r.. -- rI-3 —— ir a —
v vJ + -n I —— pv + —— pv J
a —}i
(2b)
P \oX.
Jiv liv dX. )\
I I

i rav. av."!
P.13. = - p —*- + -J- (2c)
P Ux,
-j axj
™i

a2
L. . - v ———— v. v. - 2 (2d)
1D L<Jx
L 3 k dx
^ 9 ^k X- D•> v,dxk; ^ ^ ^ j
and

V.b n J. * V.

Here ci .k is the permutation symbol13 1) sijk = 1 if


ijk form an even perautation of I f 2 r 3 ; 2) eijk » -1 if

ijk font an odd perautation of I r 2 ,3 ; **** 3^ eijk " °


if ijk do not form a perautation of I r 2 f 3 -
The terms listed represent the following physical
processes2 :
1) n,. . represents production by the mean velocity
gradients;
2) TV . represents diffusional processes due to
turbulent"3 fluctuations;
238 D. NAOTANDJ.TANNY

3) P t . expresses turbulent velocity redistribution


by pressure fluctuations;
4) L. . describes loss by molecular diffusion and
dissipation; and
5) M. . counts for the magnetohydrodynamic
interactions.
Diffusional Magnetic Fluctuations
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Systematic order of magnitude estimation10'11 for


diffusional magnetic fluctuations characterized by
small magnetic Reynolds number, suggest for the direct
electromagnetic interactions

M.i j. » - [v.
i J mj n
B }mn
e. + v.
3 J mj nB iron
e. v (3)
'
P \ /

Using the equation for the electric current


fluctuations

v x b - v x b (4)

further restricted by counting the largest constituent


only

j = a(v x B) or jm = 0vkBl 8 m k l (5)

Equation (3) becomes


2a
M. J. - —
P (yTMj + vTMi)
|i- |v,v, B,B, + v,v, 8,8, | -

Equation (6) can be implemented directly in a Reynolds


stress closure with no need for further modelling.
To find the indirect mode of electromagnetic
interactions1 ar11 Poisson equation for the pressure
fluctuations ° was derived from the momentum
equations

P
"
MAGNETOHYDRODYNAMIC CHANNEL FLOW 239

with

3V_ 3v, 32 f —
I v.k vn - v^v
k « II (8a)
vww/
3x
*

and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

1
S =-V-jxB + Jxb+jxb-jxb (8b)
P

Equation (7) was then split into


1 32
» PI = si
p axk3x and -7-^— =S
p 3X.9X, P 2 2
k

defining two constituents of the pressure fluctuations


P - Pi + P2 (10)
As a result, one may also split Equation (2c) for the
pressure redistribution

p.. =?.<«> + P.;M) (ID


counting for the hydrodynamic and the magnetic
contributions to the pressure strain correlations,
respectively. To obtain these, a turbulence model is
needed. Here, the outcome of the quasi-isotropic
model14 is used.
Quasi-isotropic Approximations
The quasi-isotropic model for the magnetic pressure
redistribution developed in Refs. 10 and 11 is

~ ~ r

-J-I ^ ^. (12a)
240 D. NAOT AND J. TANNY

+ b kfB.B. - J-B S. .1 (12b)

[f-k 4lj - V^ (120

Equation (12) depends on three coefficients a, b, and


c, which, in turn, depend on the macroscale <(>
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

27 - 36<|>
a = ————— - (13a)
105
- 3 - 24*
b - ————— - (13b)
105
and

It is the same <f> that is later used to determine the


three coefficients of Pi(.H) .
The quasi-isotropic model for hydrodynamic pressure
redistribution shown in Refs.l, 2 and 14 is

vdiere

3V. ——— 3V.


D. , = - v.
x lv. —— + 3v.v.
l ——i (14a)
3x. 3x.

and

3V. 3V.
(14b)
3x.. 3xt
MAGNETOHYDRODYNAMIC CHANNEL FLOW 241

The three model coefficients depend on the macroscale


*
76-8<|>
a = ——— (15a)
105
22+64$
e = ——
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

and
9+72 <f>
v - —— use)
The dissipation is modeled2 by

L
ij -- v A + -.kd-A)*^ (16)

where e is the energy dissipation rate. The pressure


redistribution term suggested by Rotta15 is
incorporated in Eq.(16) extending the interpretation
of Eq.(16) to the total return to i sot ropy model. The
model coefficients calculated for + - -0.5 are given
in Table 1.

Table 1 Ttie pressure redistribution model coefficients


for <*> - -0.5

Hydrodynamic Electromagnetic
redistribution redistribution
a - 0.762 a - 0.429
3 - -0.095 b - 0.086

Y - -0.257 c - 0.143
242 D. NAOT AND J. TANNY

Flow Field Specification


The present calculations are confined to two-
dimensional unidirectional mean flow in the x
direction U(y,z) shown in Fig.l, fully specified by
means of a single velocity gradient:

puy pir
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

* • • + • • (17,
UyJ Liz,

plus an inclination angle also shown in Fig.(l)

(3U / 31T>
* = arc tg — / — (18
l3z / 3yJ
Two orientations of the magnetic field were
considered: longitudinal shown in Fig. la , and
transverse shown in Fig. Ib . In the second case \|> = 0
describes a normal magnetic field applied to shear
flow, and i|/ =« 90° describes an azirauthal magnetic
field applied to shear flow.
The magnetic field intensity and direction are
specified by means of a local turbulent Hartman number
represented here by the parameter Ht
ak-2
a - 2— B (19)
pe
and a normalized magnetic tensor

B,, = B, BV(B U a ) (20)

The choice of Equation (19) for Ht to specify the


magnetic field intensity stems from the nondimensional
form of the stress transport equations. Still, it
contains some arbitrariness. For exact local
equilibrium this is not important, as the solutions
relate uniquely to all other possible suggestions for
such a specification.11 Using an algebraic stress
model, however, local equilibrium is not necessarily
MAGNETOHYDRODYNAMIC CHANNEL FLOW 243

a) LONGITUDINAL
b
> TRANSVERSE
MAGNETIC MAGNETIC
FIELD FIELD
AZIMUTHAL
MAGNET
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

NORMAL
MAGNET
z, w z,w

Fig. 1 Two-dimensional unidirectional flow with


a) longitudinal magnetic field b) transverse magnetic
field.

expected in most parts of the field and the uniqueness


of these relations is not maintained. It is believed
that H is the best choice for the presentation of the
magnetic field intensity in a calculation where e and
k are the basic variables.
Local Equilibrium

The analysis is also confined to the state of local


equilibrium where the production n balances the
dissipation E, plus the suppression by the magnetic
field M
e -i- M (21)
where
n = r- (22a)

e » - T (22b)

and

M = - T (22c)
244 D. NAOT AND J. TANNY

Using the abbreviation

N
iJ - 2-
V
jVk
(23)

and noting that

v
V
B = vvlvvk D B
kl = N.11 (24)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where VB is the velocity fluctuation in the magnetic


field direction, Eq.(6) contracts to

e Ht (1 - /2k) (25)
By substituting Equation (25) in Equation (21) a
relation between the dissipation and the production
needed to establish local equilibrium is obtained

e/n = - vB2/2k)] (26)


and shown in Figs. 2 and 3. The e/n ratio shown in
Fig.2 for a longitudinal magnetic field does not
depend on \|/, as the ratio u2A is theoretically11
independent of this angle. However, for transverse
magnetic field a weak dependency is shown in Fig.3.

.001
100 Hfc 1000

Fig. 2 Local equilibrium condition e/n for longitudinal


magnetic field
MAGNETOHYDRODYNAMIC CHANNEL FLOW 245

The values for vB2/2k used to yield Figs. 2 and 3


were taken from the numerical iterative solutions
later described.
In local equilibrium state the production should
balance both the dissipation and the magnetic
attenuation. It is apparent that with the increase of
the turbulent Hartman number, the relative part
allocated to balance the dissipation decreases until
finally the production mainly balances the magnetic
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

attenuation.
Mathematically local equilibrium state defines the
relative values of the turbulent stresses v"v"/k
only. Their absolute values are obtained Vrom
additional information. Using a full k-e transport
model, an excess in the production results in an
increase in the energy level, and an excess in the
dissipation results in laminarization. it is the
solution of the full transport equations that
determines the energy level. A need for further
adjustment of the e equation is emphasized in this
context. With 8 equation tuned for nonmagnetic flows
applied to magnetohydrodynamic channel flow,
laminarization will start for relatively low turbulent
Hartman number, just as soon as the excess in M+e
becomes effective. However, with e equation adjusted
to describe the "Reverse Cascade" the value of e will
become considerably smaller with an increase in H due
to energy transfer from small scale eddies to large
scale ones; and the excess in M+e will become

.1

.01

.001 .1 10 100 H 1000

Fig. 3 Local equilibrium condition e/n for transversal


magnetic field
246 D. NAOT AND J. TANNY

effective only for much larger values for Ht. The


quality of the energy prediction depends on our
ability to properly describe the Reverse Cascade.

Iterative Procedure
In order to obtain numerical solutions/ the stress
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

transport equations for the state of local equilibrium

R + P i < H » +Pi (M, + M I . +Li . =0 (27)

were reorganized accumulating all the terms that


include v. v. explicitly on one side of the equation,
thus yielding

a-

- C f-k &., + N.. ]}/[A + Ht(l-C)l (28)

With the observation that Eq.(27) may only result in


relative values for viv.A once a condition relating 8
to the flow parameters3 is met/ the following strategy
was practiced. At each iteration old values of v. v,
were used to calculate the right hand side of Equation
(28) and obtain new values for vtv. . These were
under relaxed by 60% and were renormallzed keeping a
fixed value for the turbulence energy. At the sane
step the dissipation was recalculated by Equation (26)
using new values for n and v B 2 and was also under-
relaxed by 60%. The results shown here were obtained
after 100 iterations. To verify convergency, sane
cases were run for 500 iterations underrelaxed to 80%.
MAGNETOHYDRODYNAMIC CHANNEL FLOW 247

Longitudinal Magnetic Field


The turbulent stresses calculated for uni-
directional flow with longitudinal magnetic field are
shown in Figs. 4 and 5 in terms of Ht and \j/. Indeed,
u2/k is independent of \j/ and the other stresses show
trigonometric dependency on \f/ as was shown in Ref.ll.
All the components, apart from u2A/_decrease with
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

the Ht increase. The large ratio u2/k typical of


nonmagnetic flow, now becomes larger and the flow
tends to become practically one-dimensional. Still,
even for infinitely large Ht the other components do
not vanish and show a limiting new structure. Another
important observation is the fact that the stresses
that_ control_ the secondary currents in channel flow
v^w2 and vw, decrease monotonically with the H
increase. Therefore, we should expect attenuation of
the secondary currents with the application of a
longitudinal magnetic field.

i 1 i

1.4 10 _
2
1
v2/k w /k
1.2

- vu
1
1 -

.8 _
G2/k
.6 H =0

"C^^ ^-^^^^
.4 - ^=s^^^\^^^
— - ^ 1 0 — -— zrr^io"^^—
———^m^-— oo
.2 ~
"
1 \ I 1——i——— i i i i i i i i i i
15° 45° 75° 15° 45° 75° 15° 45°

Fig. 4 The turbulent normal stresses for longitudinal


magnetic field
248 D. NAOTANDJ. TANNY
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

15° 45° 75° 15° 45° 75° 15° 45° 75°y

Fig. 5 The turbulent shear stresses for longitudinal


magnetic field

Transverse Magnetic Field


The turbulent stresses calculated for
unidirectional flow with transverse magnetic field are
shown in Figs. 6 and 7 as functions of Ht and \j/. Now,
with the increase of Ht the relative values of the
components u2/k, w*/k and -uw/k decrease; the compo-
nents v*/k and -uv/k increase; and -vw/k is practical-

15° 45° 75° 15° 45° 75° 15° 45

Fig. 6 The turbulent normal stresses for transverse


magnetic field
MAGNETOHYDRODYNAMIC CHANNEL FLOW 249

ly unaffected. In this case/ u2A remains large in


spite of the magnetic attenuation and the flow tends
to become practically two-dimensional. The phenomena
is more pronounced for azirautal magnetic field
(\|> « 90°) than for normal magnetic field (\|; « 0).
Still, a limiting structure for infinitely large Ht
exists with nonvanishing values for all the
components. It is also of interest to note that ^-w2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

which controls the secondary currents is substantially


affected by the magnetic field, while vw is not.
Therefore, unlike the case of longitudinal magnetic
field, we should expect an increase in the secondary
currents with the application of transverse magnetic
field.
rks Towards Algebraic Stress Modelling
An important objective in the development of an
algebraic stress model is to find ways to extend the
use of computation techniques developed for the k-e
turbulence model. An important example is the tunning
of the eddy viscosity concept. Using local
equilibrium solutions for the stress transport
equations, and the definitions

15° 45° 75° 15° 45° 75° 15° 45

Fig. 7 Hie turbulent shear stresses for transverse


magnetic field
250 D. NAOTANDJ.TANNY

.1
Fig. 8 The eddy
viscosity coefficients
.01
for longitudinal
magnetic field
.001 10
.1
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

100 Hfc 1000

— k2 3u (29a)
uv - - C — —
"7 e 3y

and
— k2 3U
uw = - C z — — [29b)
" 6 3z
it is is possible to estimate Cyt/y and C//* . Similar
practice16 for nonmagnetic flow showed that
C «• C zz - C (0) are not functions of V/. In Ref.ll it
i/y //
was shown for longitudinal magnetic field that C
equals Cv- both depend on Ht , and are totally
independent of x|/. The numerical evaluations confirm
these findings and are shown in Fig.8.
For transverse magnetic field, however, such a
theory does not exist and the two coefficients were
estimated and shown in Fig.9. These were found to be
practically independent of \J/ for the present choice of
model coefficients. Attempts to use the near wall set
of coefficients of Ref.1 showed weak dependency on y.
However, as solutions for the longitudinal magnetic
field failed to converge for this set, these results
are not discussed here.
The analytic description16 of the stresses that
control the secondary currents, v2, w2 and vw, in
in terms of k, e, 3U/3y, and 3U/3z, was extended in
Ref.ll for the longitudinal magnetic field. The
information needed for this description is restricted
to two values to be derived from the local equilibrium
MAGNETOHYDRODYNAMIC CHANNEL FLOW 251

Fig. 9 The eddy


viscosity coefficients
.01 r for transversal
magnetic field
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

.001
10 100H 1000

solutions, presumably in the form of two empirical


correlations of Ht. Unfortunately, similar analysis
for transverse magnetic field does not exist and the
elaboration of the description of these stresses, on
the base of local equilibrium, is expected to be more
complicated, as it may involve empirical correlations
of both * and Ht.
Finally, the wide range of values for Ht presented
here may not be practiced in real channel flow. In
fact, proper description of the Reverse Cascade,
expected to control the dissipation and the
laminarization process, is also expected to restrict
the range of Ht materialized. This, in turn, may also
affect the turbulence unisotropy shown here to be
strongly influenced by Ht.

References
1
Launder, B., Reece, G., and Rodi, W., "Progress in
the Development of a Reynolds Stress Turbulence
Closure,11 J. Fluid Mechanics, Vol.68, 1975,
pp.537-566.
2
Wolfshtein, M., Naot, D., and Lin, A., "Models of
Turbulence," Current Topics in Thermal Science, edited
by C. Gutfinger, Hemispher Co., Washington. 1975,
pp.3-45.
3
Naot, D., and Rodi, W., "Applicability of
Algebraic Models Based on Unidirectional Flow to Duct
Flow with Lateral Motion," Int. Jour, for Numerical
Methods in Fluids, Vol.1, 1981, pp.225-235.
252 D. NAOT AND J. TANNY

4
Narasimha, R., "Relaminarization - Magnetohydro-
dynamic and Otherwise," Liquid Metal Flow and MHD,
edited by H. Branover, P. Lykoudis, and A. Yakhot,
Progress in Astronautics and Aeronautics, Vol.84,
AIAA,5
New York, 1983, pp.30-52.
Sukoriansky, S., Zilberman, I., and Branover, H.,
"Experiments in Duct Flows with Reversed Turbulence,"
Single and Multi-Phase Flows in an Electromagnetic
Field, edited by H. Branover, P. Lykoudis, and M.
Mond, Progress in Astronautics and Aeronautics,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Vol.100,
6
AIAA, New York, 1985, p.111-124.
Moreau, R., "Why, How and When MHD Turbulence
Becomes Two-Dimensional." Liquid Metal Flow and MHD,
edited by H. Branover, P. Lykoudis and A. Yakhot,
Progress in Astronautics and Aeronautics, Vol.84,
AIAA, New York, 1983, pp.20-29.
7
Sommeria, J., "Two-Dimensional Behaviour of
Electrically Drived Flows at High Hartman Number,"
Single and Multi-Phase Flows in an Electromagnetic
Field, edited by H. Branover, P. Lykoudis and M. Mond,
Progress in Astronautics and Aeronautics, Vol.100,
AIAA,8
New York, 1985, pp.77-88.
Caperan, P., and Almany, A., "Transition from
Three-Dimensional to Quasi-Two-Dimensional MHD Grid
Turbulence," Single and Multi-Phase Flows in
Electromagnetic Field, edited by H. Branover, P.
Lykoudis a n d M ^ M o n d , Progress in Astronautics and
Aeronautics, Vol.100, AIAA, New York, 1985, pp.89-99.
*Tsinober, A., "MHD Flow Drag Reduction," Viscous
Drag Reduction, edited by D.M. Bushnell, and J.N.
Hefner, Progress in Astronautics and Aeronautics, to
be published.
*° Naot, D., and Peled, A., "Magnetohydrodynamic
Redistribution of Three-Dimensional Turbulence,"
Current Trends in Turbulence Research, Vol.112, edited
by H. Branover, M. Mond, and Y. Unger, AIAA, 1988,
pp.448-457.
xl
Naot, D., Peled, A., and Tanny, J., "Response of
Shear Flow Turbulence to Diffusional Electromagnetic
Fluctuations," Applied Mathematical Modelling, Vol.14,
April12
1990, pp.226-236.
Hinze, J.O., Turbulence, McGraw Hill, New York,
1959.13
Aris, R., Vectors, Tensors and the Basic
Equations of Fluid Mechanics, Prentice-Hall, Englewood
Cliffs, N.J., 1962.
MAGNETOHYDRODYNAMIC CHANNEL FLOW 253

14
Naot, D., Shavit, A., and Wolfshtein, M., "Two
Point Correlation 1Model and the Redistribution of
Reynolds Stresses/ The Physics of Fluids, Vol.16,
1973, pp.738-743.
15
Rotta, J., "Statistical Theory of Non Homogeneous
Turbulence,"
16
Physik, Vol. 129, 1951, p.547.
Naot, D., "Two-Dimensional Unidirectional
Turbulent Flow in a Local Equilibrium," The Physics of
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fluids, Vol.18, 1976, pp.1813-1814.


Two-Phase Grid Turbulence
Th. Panidis* and D. D. Papailiouj
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

University ofPatras, Patras-Rion, Greece

Abstract

The results of an experimental study on water-air bubble two-


phase turbulence are presented. The experiments were conducted in
a water channel with a vertical test section. The nearly isotropic
turbulence flowfield was generated by a grid located at its entrance
part. The water constant volume mean flow was directed upwards,
whereas the air bubbles, consisting in the dispersed phase, were
injected from the grid at a variable rate. Laser-Doppler velocimetry
has been applied to measure instant and mean velocities of the
liquid phase. Turbulent statistical characteristics such as turbulence
intensity, autocorrelation, skewness and flatness factors, as well as
turbulence macroscales, were measured for this phase. Gas phase
measurements, including void fraction, bubble velocities, and bubble
sizes, were also conducted by applying photographic techniques.
Certain differences between the presented experimental results and
those existing in the literature, concerning the dependence of
measured quantities on void fraction, have been identified and
discussed. The obtained measurements indicate that the second
phase introduces drastic changes in the turbulence structure, mainly
through buoyancy-originated mechanisms. These changes are evident
in both the mean velocity measurements and those of the turbulence
statistical characteristics.

Copyright © 1991 by Th. Panidis and D. D. Papailiou. Published by the American


Institute of Aeronautics and Astronautics, Inc. with permission.
*Research Assistant, Mechanical Engineering Department, Laboratory of Applied
Thermo dynamic s.
fProfessor, Mechanical Engineering Department, Laboratory of Applied Thermo-
dynamics.
254
TWO-PHASE TURBULENCE 255

Introduction

In the last two decades, several attempts have been made in the
direction of basic understanding of two-phase flows. Measurements
and computations of the local properties of simple two-phase flow
configurations have been conducted by several investigators.1"13 A
brief summary of the results reported in these investigations follows.
Serizawa et al., " in the first of a sequence of three papers,
presented a thorough assessment of techniques in use for two-phase
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

measurements. In the second paper, a study of the mean and


statistical characteristics of an upward water-air bubble flow in a
pipe was presented. No clear connection between void fraction and
turbulence intensity was found in these experiments. The third
paper3 addresses the problems of heat diffusion and bubble
dispersion in two-phase flows. It also offers a mixing length analysis
for momentum transport.
Theofanous and Sullivan investigated the turbulence structure of
the continuous phase in water-nitrogen bubble flow, using Laser-
Doppler velocimetry. Contrary to the findings reported in Ref. 2,
they reported a strong dependence of turbulence intensity on void
fraction. They also proposed a model for the prediction of
turbulence level in two-phase flow. Similar measurements were
conducted by Wang et al., in water-air bubble up and down flows
in circular conduits. A single and three sensor hot film anemometry
were applied to measure Reynolds stresses and to derive a
relationship between the void fraction distribution and the
turbulence structure in the continuous phase.
An approach similar to that described in the present work was
adopted in the experiments reported by Lance et al., Lance and
Bataille, and Marie. These investigators used the same test rig, a
grid turbulence water channel in which air bubbles were injected
from the grid. These experiments were conducted at constant
velocity at the center for varying void fraction. In the first two
papers, ' measurements of the longitudinal turbulence intensity,
skewness and flatness factors, autocorrelation, and one-dimensional
spectra were obtained by using primarily hot film anemometry.
Bubble diameter and velocities were also determined with the aid
of a two hot film arrangement and photography, whereas an optical
probe was used to measure void fraction. Their results indicated the
strong dependence of the turbulence intensity on the square of the
void fraction and little or no dependence of the autocorrelation and
macroscale on void fraction. In Ref. 8 similar measurements were
reported in which Laser-Doppler velocimetry was applied.
256 TH. PANIDIS AND D. D. PAPAILIOU

The approach adopted in the present experiment is that of


examining the influence of a second dispersed phase on a simple,
well-investigated turbulent flow. Following the historic development
of ordinary, single-phase turbulence, the nearly isotropic field
created behind a grid was chosen as the reference flow.
The selected isotropic turbulent field is expected to alter in the
presence of the second phase for the following reasons. Because of
buoyancy effects, the isotropy of the flow is altered since bubbles
behave as agitators transferring energy to the mean flow in a
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

preferred direction. Also, the presence of the bubbles is responsible


for changes occurring in the flow structure since the continuity of
the liquid phase is destroyed at length scales related to the size and
motion of the bubbles.
The aim of the present continuing effort has been to conduct
extensive and as complete as possible experimental measurements
on the mean and statistical characteristics of the two-phase flow
under consideration. It is expected that understanding the turbulence
structure and transport processes of the selected simple case would
provide valuable information to be used in more complex two-phase
flow configurations. Of particular interest would be the study of the
mechanism by which the dispersed phase influences the cascade
processes, resulting in changes in the distribution of turbulence
energy among spectral wave numbers. However, certain differences
were identified between the present measurements and
corresponding ones reported in the literature. ' These differences,
which emphasize some important aspects of the interaction
processes between the two phases not readily observable in previous
experiments, suggested a more complete, yet gradual, investigation
and presentation of the flow under consideration.

Experimental Arrangement and Procedure

The experiments were conducted in the water channel of the


Laboratory of Thermodynamics, having a vertical test section of
300 x 300 x 1200 mm . A nearly isotropic field was produced by a
grid located at the lower part of the test section. The grid was
biplane, made of copper tubes with an outside diameter of 5 mm,
crossed at a mesh spacing, M = 30 mm. It was also used for the
injection of the dispersed phase. Hypodermic needles with inside
diameters of 0.2 mm were attached to the tube intersections of the
grid. Air, fed through the tubes, was eventually injected in the flow
at a rate controlled by pressure regulation.
TWO-PHASE TURBULENCE 257

The mean flow was directed upwards so that buoyancy effects, due
to the presence of the bubbles, had no influence on the symmetry of
the flow.
The measurements were conducted at a constant water volume
flow rate. In this way, the reference field, i.e., the grid-produced
turbulent field, was kept unaltered.
The obtained measurements referred to the flow characteristics
of both the liquid and gas phases. All measurements were made at
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

a distance from the grid, x/M = 30.


The water velocity field was monitored with a dual beam forward
scatter LDV. A 15 mW Spectra Physics Laser and TSI optics were
used. The measuring volume had 57 fringes at 3.18 |im spacing. The
signal from the photomultiplier was transmitted to a TSI 1980-B
counter. The amplitude limit control was used to ensure that only
signals from the water phase would be processed. The analog signal
of the time for eight cycles of the Doppler-burst, produced by the
counter, was checked for proper data rate with an oscilloscope and
fed to the computer. For this purpose a 14-bit A/D converter was
used (the counter produces the analog signal from a 12-bit digital
one). The sampling rate was constant and could be set by software
up to 16 kHz. Sets of 20,000 samples were taken, at a rate of
100 Hz, for the estimation of mean values and statistical moments
and at 2 kHz for that of the autocorrelation. Data points more than
30% apart from the mean value were rejected as probable bubble
measurements.
Bubble measurements were conducted with double-exposure
photography.13 A double light source flashing at a controllable time
lag produced a vertical plane light beam twice. This way, two
successive images of the bubble field were monitored on the same
frame. Bubble diameters, bubble velocities, and void fraction values
were calculated from the enlarged prints. The void fraction, thus
obtained, is spatially averaged. Considering that the averaging takes
place within a finite area (50 x 50 x 10 mm3) located at the center
of the channel, and that the flow is nearly homogeneous there, the
estimated values of void fraction should be very close to the local
value at the center of the channel. It was found that bubble mean
diameter was 3 mm, whereas bubble mean slip velocity was
approximately 25 cm/s.
For the computation of statistical moments and autocorrelations,
homemade software in C language was used, according to standard
algorithms.14'15
258 TH. PANIDIS AND D. D. PAPAILIOU

Discussion of the Experimental Results

Single Phase Measurements

Measurements of the characteristics of the water phase grid


turbulence, with a zero void fraction, were conducted at a distance
from the grid, x/M = 30. These consist of mean local velocity and
turbulence intensity distributions (Figs. 1 and 2), skewness and
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

flatness factors (Fig. 3) and autocorrelation (Fig. 4). It can be seen


in these figures that the variation of the mean velocity and the
turbulence intensity far from the wall was negligible. Skewness and
flatness factors also exhibit flat distributions in this area. The
measured autocorrelation at the center was found in good agreement
with the reported measurements by Van Atta and Chen at
Re = 25,000 and by Stewart17 at Re = 5300 in air, whereas a
significantly different distribution was reported by Lance and
Bataille7 at Re = 10,000 in water (Fig. 4).
In conclusion, the obtained measurements indicate that the

o ReM = 8000
o
0)

-4-^
>> 300
*0
O^
<D

C
O
0)

200
0.00 0.10 0.20 0.30 0.40 0.50
Test Section Width (y/D)

Fig. 1 Single phase mean velocity distribution.


TWO-PHASE TURBULENCE 259

structure of turbulence in the channel, and especially at its central


part, was nearly isotropic and, therefore, suitable for serving as a
reference for subsequent two-phase measurements.

Two-Phase Flow Measurements

Measurements at the center of the channel indicate that the


introduction of the dispersed phase results in an increase in the
mean velocity of the continuous phase, depending on the void
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

fraction. As shown in Fig. 5, these measurements exhibit two distinct


regions. At very low void fractions the mean velocity increases
slowly. Beyond a certain value, which in the present experiments is
found to be approximately equal to 0.1, it changes linearly at
drastically higher rates. This observation, which is consistent with
the measured turbulence statistical characteristics, is discussed in the
following.
In the present experiments, conducted under conditions of
constant water volume rate, the observed increase in the mean
velocity is partly due to the additional bubble volume injected in the

0.10
o ReM = 8000

CO
c
CD

o
0.05 - o
c

0.00
0.00 0.10 0.20 0.30 0.40 0.50
Test Section Width (y/D)

Fig. 2 Single phase turbulence intensity distribution.


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

K)
CT>
O
Autocorrelation Skewness Flatness
O -* to o ro ^
b b o o
oo o o o
b ,————————————————————————————
i I i | i | i ——-
o
£
« o a

p H
9 o o a
HQ

I* 0)
8 0) D
0)
9* CO P
o a a o z
I D
s.8 *T"| (y) D
0*
D O ^
I; o*
1 £^" ^ D
$ O
f 0 D »» TJ
ft ^ 0)
i rr F
jj O
1 <?
§*
§ r-f °
g. S ^k o a II II
0* O
00 CX)
• 00
00
00
IlII p
In
r*» o a
TWO-PHASE TURBULENCE 261

flow. A second contributing factor is the dispersing action of the


bubbles and the eddy structure created by their motion, which is
responsible for alterations occurring in the mean velocity profile.
Existing experimental evidence2' ' indicate that, at low values of
the void fraction, its distribution is very flat. This affects the mean
velocity profile that exhibits a wider flat portion, in the case of
shear turbulence, than that corresponding to the single phase. ' It
appears, therefore, that the observed slight increase of the mean
velocity at low-void fractions is produced by the flattening effect of
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

the dispersed phase on the mean velocity profile. Subsequent


increase of the void fraction results in a decrease of the flat portion
of the profile, which changes to a dome-shaped one at high-void
fractions.2'5 It is, therefore, evident that the linear increase of the
mean velocity at high-void fractions is due to the rapid domination
of the dispersed phase on the turbulent flowfield of the continuous
phase.
As already mentioned, measurements of turbulence statistical
quantities, conducted at the center of the channel, appear
compatible to those of mean velocity. Turbulence intensity, skewness
and flatness factors, macroscale, and autocorrelation measurements

1.6
o Present Work U0 = 0.29 m/s
^ 1.5
* Serizawa et al. U0 = 0.74 m/s
o D Serizawa et aL U0 = 0.88 m/s
o
1.4
c
D

1.3

o
c
.2 1.2
CO
c
1 1.1
C
o
1.0

0.9
0.00 0.01 0.02 0.05
0.03
0.04
Void Fraction
Fig. 5 Mean velocity variation with void fraction.
262 TH. PANIDIS AND D. D. PAPAILIOU

o Present Work U0 = 0.29 m/s


* Serizawa et al. U0 = 0.74 m/s
° Lance & Bataille U0 = 0.30 m/s
0.15

CO
c
CD

0.10
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(D
O
C

"5
.£>

0.05

0.00
0.00 0.01 0.02 0.04
0.03
0.05
Void Fraction
Fig. 6 Turbulence intensity variation with void fraction.

(Figs. 6-10), indicate the existence of a two distinct region pattern,


similar to that identified in the mean velocity measurements. The
void fraction value, corresponding to the point separating the two
regions, is found to be approximately the same as in the velocity
measurements, except in the case of macroscales, where it appears
slightly increased. However, contrary to the velocity measurements,
drastic changes in the measured quantities, indicating substantial
changes in the turbulence structure due to the presence of the
dispersed phase, occur at low-void fraction values.
The influence of the dispersed phase on isotropy is evident in the
skewness and flatness measurements presented in Figs. 7 and 8. At
low-void fractions the isotropy of the turbulence field is severely
damaged. However, a trend of gradual return to isotropy with
increasing void fraction is observed in the high-void fraction region.
In general, high values of skewness are associated with convection
of turbulence energy from regions of high intensity to regions of low
intensity. Also, increased values of flatness indicate the existence of
inhomogeneity in the turbulence field. Therefore, it can be surmised
that the dispersed flow influences the structure of the turbulence by
introducing a flow inhomogeneity in the liquid phase, causing a
TWO-PHASE TURBULENCE 263

Void Fraction

o ReM = 8000

o 1.00
-M
O
D
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

CO
CO
0)
c
£
0)
£0.50

0.00
0.00 0.01 0.02 0.03 0.04 0.0
Void Fraction
Fig. 7 Skewness factor variation with void fraction.

simultaneous transport of turbulence energy. Further void fraction


increase results in an increased domination of the dispersed phase
on the turbulent structure of the liquid phase. The observed trend
in skewness and flatness factor measurements in the high-void
fraction region possibly indicates that a state of equilibrium might
be reached between the dispersed phase and the grid turbulence at
void fraction values not attained in the present experiments.
Further information on the nature of the energy transfer
mechanism between the dispersed phase and the grid-generated
turbulence can be inferred from the autocorrelation and macroscale
measurements (Figs. 9 and 10). Although further measurements
related to the statistical character of the turbulence field, such as
probability functions and turbulence spectra, are needed for a final
conclusion to be reached, it appears that energy transfer occurs as
a result of an interaction between the dispersed phase and the small
wave number (larger scale) energy containing eddies. At low-void
fractions the formation of two distinct size eddies exist as the
18
autocorrelations' double-structure shape and the reduced
macroscale values indicate. In this context, the reduction in the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Autocorrelation Flatness Factor


O Ol C
b b c
0 0 C

i
SI ^
00
0
a 8 0
g
ST O
g o < b O
0)' ^ O
ft. r+- o
Q
CL
0 p
D 1
O -i o p
CD 1Q b 0
O ND T)
|. r-H
3 |o' o °
D
I 3 0
s a 73 O
b O (D
OJ
pf
0 H
p
o CD
I b a
O

co
0
b ——————————————————————————————
en
TWO-PHASE TURBULENCE 265

o ReM = 8000
20

D
O
CO
2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

O
D

D
Q.
CO

0.00 0.01 0.02 0.03 0.04 0.05


Fig. 10 Spatial macroscale variation with void fraction.

turbulence macroscale to a size comparable to that corresponding to


the introduced bubbles should be noted. At higher-void fractions the
increasing presence of the bubble-generated eddy structure gradually
results in more uniformly sized eddies, as the autocorrelation
changes to a single-structure shape.
As presented in this work, autocorrelation and macroscale
measurements are in serious disagreement with those reported in
Refs. 6 and 7. According to these authors, measured autocorrelations
show little or no void fraction influence beyond an initial drastic
change occurring at very low-void fraction. Notwithstanding the
different experimental conditions under which the two experiments
were conducted, it is difficult to explain the reported results since
they should be interpreted as indicating no changes in the
turbulence structure at high-void fractions.
The measured turbulence intensities at the center of the channel
for increasing void fraction values, reflect the discussed changes in
the structure of the grid turbulence caused by the presence of the
dispersed phase. At low-void fractions corresponding to the first
region, the small increase of the turbulence intensity is due to the
energy dispersing action of the bubbles. At higher-void fractions the
increased presence of the bubbles, and that of the energy containing
266 TH. PANIDIS AND D. D. PAPAILIOU

eddies produced by the bubble motion, result in an increased energy


transfer to the liquid phase turbulence, manifesting itself by the
drastic linear increase in turbulence intensity.
A comparison of the intensity measurements obtained in the
present experiments and those reported in Refs. 2 and 7 is also
shown in Fig. 6. It is evident that the present measurements and
those obtained at the center of a two-phase pipe flow ' at constant
water volume flow rates exhibit similar behavior. The turbulence
intensity measurements presented in Ref. 7 correspond to a constant
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

velocity at the center and, therefore, are not directly comparable to


the presented experimental results. They are not affected by the
velocity profile changes and give higher values for higher-void
fractions because they correspond to lower mean flow rates.

Conclusions

The conducted experiments have shown that the introduction of


the dispersed phase results in drastic changes in the turbulence
structure of the liquid phase, mainly through two buoyancy-induced
mechanisms. They are, respectively, based on the dispersing action
of the bubbles, causing the alteration of the mean velocity profile,
and on the interaction between the eddies generated by the bubble
motion and grid eddies of comparable size.
The obtained measurements clearly show the existence of two
distinct regions corresponding to low- and high-void fractions. At
low values of void fraction the influence of the dispersed phase
through the first mechanism manifests itself by mainly affecting the
mean velocity and the turbulence intensity. The second mechanism
is responsible for the substantial changes in the grid turbulence
structure, as inferred from the measurements of the turbulence
statistical characteristics. At high-void fractions the eddies generated
by the bubble motion dominate the transport processes. This
introduces a drastic increase in the mean velocity and turbulence
intensity, as well as a trend of gradual return to isotropy, possibly
indicating that a state of equilibrium might be reached at higher
values of void fraction.
Finally, the present experiments reveal certain differences to exist
with previously reported measurements partly due to differences in
the adopted experimental conditions.
TWO-PHASE TURBULENCE 267

References
^Serizawa, A., Kataoka, I., and Michiyoshi, I., "Turbulence Structure of Air-Water
Bubbly Flow - I. Measuring Techniques," International Journal of Multiphase Flow,
Vol. 2, 1975, pp. 221-233.
2
Serizawa, A., Kataoka, I., and Michiyoshi, I., 'Turbulence Structure of Air-Water
Bubbly Flow - II. Local Properties," International Journal of Multiphase Flow, Vol. 2,
1975, pp. 235-246.
Serizawa, A., Kataoka, L, and Michiyoshi, I., Turbulence Structure of Air-Water
Bubbly Flow - III. Transport Properties," International Journal of Multiphase Flow,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Vol. 2, 1975, pp. 247-259.


^Theophanous, T. G., and Sullivan, J., Turbulence in Two-Phase Dispersed Flows,"
Journal of Fluid Mechanics, Vol. 116, 1982, pp. 343-362.
5
Wang, S. K, Lee, S. J., Jones, O. C, Jr., and Lahey, R. T., Jr., "3-D Turbulence
Structure and Phase Distribution Measurements in Bubbly Two-Phase Flows,"
International Journal of Multiphase Flow, Vol. 13, No. 3, 1987, pp. 327-343.
Lance, M., Marie, J. L., Charnay, G., and Bataille, J., "Interaction of a Grid-
Turbulence and a Statistically Uniform Swarm of Air Bubbles," International
Symposium on Applications of Fluid Mechanics and Heat Transfer to Energy and
Environmental Problems, University of Patras, Patras, Greece, Vol. Ill, 1981, pp. 52-65.
Lance, M., and Bataille, J., Turbulence in the Liquid Phase of a Bubbly Air-
Water Flow," NATO Workshop, Shliersee, Germany, 1982.
Marie, J. L., "Investigation of Two-Phase Bubbly Flows Using Laser Doppler
Anemometry," PhvsicoChemical Hydrodynamics, Vol. 4, No. 2, 1983, pp. 103-118.
^Marie, J. L., and Lance, M., Turbulence Measurements in Two-Phase Bubbly
Flows Using Laser Doppler Anemometry," Measuring Techniques in Gas-Liquid Two-
Phase Flows, Springer, Berlin, Germany, 1984, pp. 141-148.
10
Durst, F., Taylor, A. M. K. P., and Whitelaw, J. H., "Experimental and Numerical
Investigation of Bubble-Driven Laminar Flow in an Axisymmetric Vessel," International
Journal of Multiphase Flow, Vol. 10, No. 5, 1984, pp. 557-569.
11
Durst, F., Schoenung, B., Selanger, K., and Winter, M., "Bubble-Driven Liquid
Flows," Journal of Fluid Mechanics, Vol. 170, 1986, pp. 53-82.
12
Gherson, P., and Lykoudis, P. S., "Local Measurements in Two-Phase Liquid-
Metal Magneto-Fluid-Mechanic Flow," Journal of Fluid Mechanics, Vol. 147, 1984,
pp. 81-104.
Mahalingam, R., Limaye, R. S., and Brink, J. A., Jr., "Velocity Measurements in
Two-Phase Bubble-Flow Regime with Laser-Doppler Anemometry," AIChE Journal,
Vol. 22, No. 6, 1976, pp. 1152-1155.
Durst, F., Melling, A., and Whitelaw, J. H., Principles and Practice of Laser-
Doppler Anemometrv, Academic Press, New York, 1976.
15
Buchhave, P., George, W. K., Jr., and Lumley, J. L., "The Measurement of
Turbulence with the Laser-Doppler Anemometer," Annual Review of Fluid Mechanics,
Vol. 11, 1979, pp. 443-503.
16
Van Atta, C. W., and Chen, W. Y., "Correlation Measurements in Grid
Turbulence Using Digital Harmonic Analysis," Journal of Fluid Mechanics, Vol. 34,
Pt. 3, 1968, pp. 497-515.
Stewart, R. W., 'Triple Velocity Correlations in Isotropic Turbulence,"
Proceedings of the Cambridge Philosophical Society. Vol. 47, 1951, pp. 146.
^Townsend, A. A., The Structure of Turbulent Shear Flow, Cambridge University
Press, Cambridge, U.K., 1976.
Abridged Octave Wavenumber Ring Models
for Two-Dimensional Turbulence
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

J. Lee*
Flight Dynamics Laboratory, Wright-P alter son Air Force Base, Ohio 45433

Abstract
From the two-dimensional Navier-Stokes equations in
spectral form, we constructed the abridged models by
retaining only the long *-rings separated by octaves.
The original system is recovered when all of the octave
k-rings are included, and our interest is in modeling
the two-dimensional flow with the fewest possible k-
rings. The efficacy of abridged octave *-ring models
was tested by numerical simulations. In the inviscid
case, the abridged model with the three longest chains
of octave *~rings (extending up to the wave number of
64) observes the equilibrium energy-enstrophy equipar-
tition. Further, the five longest chains of octave st-
rings (extending up to the wave number of 128) can ex-
hibit simultaneous development of the upward energy
(-5/3) and downward enstrophy (-3) cascade spectra when
the forcing maintains a constant flux of enstrophy
throughout the inertial range. These observations per-
sist as more octave *-rings are included in the abridg-
ed truncation model; hence, they are believed to be
intrinsic to the two-dimensional flow dynamics.
I. Introduction
Since the first simulation work of Lilly,1 numerous
investigators have performed the numerical integration
of two-dimensional homogeneous turbulence to simulate
the inviscid statistical mechanics,2"4 free decay,5 12
and forced quasistationary dynamics.13"17 In spite of
This paper is a work of the U.S. Government and is not subject to copyright
protection in the United States.
*Research Scientist, WL/FIBG.
268
MODELS FOR TWO-DIMENSIONAL TURBULENCE 269

ever refined numerics over the past two decades, the


two-dimensional flow has refused to reveal a unique ex-
ponent ft for the k~P energy spectrum, in parallel to
the *~ 3 energy spectrum for Kolmogorov's inertial
range in three dimensions. In fact, the previous simu-
lations have produced energy spectra with the exponent
in the range of 3<0<6. By Kolmogorov's type of reason-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

ing, Kraichnan18 argued that 0=3 represents the so-


called enstrophy cascade in two dimensions, whereby en-
strophy that is fed into the large eddies propagates to
high wave number without loss, only to be dissipated by
the small eddies lying beyond the inertial range. The
wide spread of observed exponents indicates that the
enstrophy cascade is not a universal eddy transport
process of the two-dimensional flow. Although this re-
futes the universality of Kraichnan's argument, we show
that the enstrophy cascade manifested by 0=3 is observ-
ed when a quasiconstant flux of enstrophy is sustained
throughout the inertial range.
In Sec. II, we obtain the isotropically truncated
system by explicitly enumerating the spectral equations
of motion in a circular wave vector domain inscribed by
an upper wave number. The enumeration has been perform-
ed by symbolic manipulations on a computer, which saves
the coupling coefficients in a one-dimensional array.
Although it essentially entails tedious convolutions,
one needs to carry it out only once. Later, for time
integration the triad-interactions in the right-hand
side are formed by multiplying the coupling coeffici-
ents by the Fourier amplitudes, which are also serial-
ized in a one-dimensional array. The isotropically
truncated system represents a dynamical system that
obeys exactly the conservation properties for each
triad-interaction in the inviscid case. Furthermore,
under forcing and damping it permits us to prescribe a
forcing procedure that maintains quasisteady enstrophy
dissipation, as dictated by Kraichnan's enstrophy cas-
270 J. LEE

cade.18 In spite of this, the isotropically truncated


system has a practical difficulty when the upper wave
number is large, for the array of coupling coefficients
becomes too large for efficient trajectory computation.
Therefore, this has led us to abridge the isotropically
truncated system by retaining only several of the long-
est chains of octave *-rings for computational expedi-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

ency (Sec. Ill ) .


Since the abridged /--ring truncation is motivated
by intuition, the efficacy of such truncation must be
examined by numerical experiment. We have first shown
in Sec. IV that the abridged model with as few as three
of the longest octave k -rings can exhibit the energy-
enstrophy equipartition up to the upper wave number of
64. Although this is encouraging, the inviscid statis-
tical mechanics has no direct bearing on the spectral
behavior of a dissipative system. We have thus shown
in Sec. V that at least five of the longest octave k-
rings are needed to simulate simultaneous development
of the k~* and *~3 energy spectra when a stationary
forcing is applied to the 8-ring. The validity of such
energy spectral behavior was acertained by increasing
the number of octave *-rings included in the abridged
truncation model (Sec. VI).
II. Isotropic Truncation
We begin with the incompressible Navier-Stokes
equations

ti-vti = -(I/P)VP + i/v2^ + P (i)


where P is the static pressure, p the fluid density,
and v the kinematic viscosity. Like the incompressibi-
lity V'&=0, body forces acting on the fluid are also
assumed 7^=0 . For a square flow region of side L with
cyclic boundary conditions, one may expand the two-dim-
ensional flowfield and body forces by

-*'- "here *-<


MODELS FOR TWO-DIMENSIONAL TURBULENCE 271

0,±1,±2,...) is the wave vector. By spanning the in-


compressible u(^) and i^(^) by the unit polarization
vectored) normal tot, i.e. , d(t)4(1)u (1) and£(t) =
e(*)/(i), we obtain from Eq. (1) the triad-interaction
representation of two-dimensional flow:19

(2)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where *=|1| and jk,* -» =(t-e(p) ) (e(1) -e(g) )+(t-e(g) )x


-^ ^ -> -> * I /•* » "
(e(*)*e(p)) is the symmetrized coupling coefficient.
Let us now enumerate the spectral equations of Eq.
(2) on a wave vector lattice of, say (-16, 16)x(-16, 16) ,
as shown in Fig. 1. Because of the reality requirement
u(l)=t/*(-t) , only a half of the wave vector lattice
points needs to be considered, for the remaining half
is redundant. If we further impose the isotropic trun-
cation to retain only the wave vectors within the upper

Wave Vector Lattice ITS(16)


Grid points = 1089(33x33) Wave vectors = 430

16 16

-16- 16 16

-16 -16

Fig. 1 Wave vector lattice of (-16,16)x(-16,16) with


1089 grid points. Only a half is needed for the direct
spectral equations of motion, and the wave vectors are
further reduced to 430 for isotropic truncation.
272 J- LEE

wave number K=16, there are now 430 wave vectors (Fig.
1). The isotropically truncated wave vectors can be
rearranged into contiguous k-rings, each of which in-
cludes the wave vectors with wave number in [(f2-*)1 >
z
* " ]. That is, the 1-ring has four wave vectors
V(8 • V(J) • V(-J with wave number ln
(0,,/2); the 2-ring has six wave vectors
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

V® ' V(o) ' V[-l) ' *10= [-2) W U h


(1/2,/6); etc. Let us denote by ITS(K) the isotropic
truncation system of K. Note that ITS(2) is the lowest
order system for u(\ ), . . . ,L/(t ) that has been found
ergodic by Kells and Orszag20 and Glaz.21
The direct enumeration of Eq. (2) has been coded
into a symbolic manipulation program to generate a one
dimensional array of coupling coefficients, from which
the spectral equations of motion can be constructed
along the £-rings. As shown previously,19 ITS(16) has
106,244 triad-interactions and the number of triad-
interactions increases to 1,598,084 for ITS(32). The
large memory requirement for such a long array of
triad-interactions is the obvious difficulty of the
present direct spectral formulation. However, we be-
lieve that it is the exact spectral representation
which retains all of the triad-interaction terms;
hence, there is no uncertainty of aliasing errors in-
herent in the pseudospectral procedure.22
III. Abridged Truncation Systems
As pointed out in Sec. II, ITS(32) requires about
1.6 megaword memory for the coupling coefficients,
which is certainly not an excessive requirement. How-
ever, as K increases to, say, 128, the length of
coupling coefficients becomes too large for efficient
in-core memory use in most supercomputers. Therefore,
we shall propose a way of paring down the coupling
coefficients by way of retaining only the long chains
of octave *-rings.
MODELS FOR TWO-DIMENSIONAL TURBULENCE 273

16

16 0 16 0 16 0

-16 -16 -16 -16


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

l-2-4-8-16-ring 3-6-12-ring 5-10-ring 7-14-ring

16 I , . . , . , , , , , , , , , , . 16
| , , M I M M M , , , , , 16 I . . . . . . . . . . . . . . .. 16

16 0 16 0 16 0 16

-16 -16 -16 -16


y-rmg 11-ring 13-ring 15-ring

Fig. 2 Decomposition of 430 wave vectors of ITS(16)


into eight chains of octave *-rings.

To illustrate this graphically, we have shown in


Fig. 2 how 430 wave vectors can be broken into 8 chains
of octave *-rings. First, the longest chain is the
l-2-4-8-16-ring, and the second longest is the 3-6-12-
ring followed by the 5-10-ring and the 7-14-ring. The
remaining four chains have only a single *-ring since
the lead-in wave number is greater than 8. Now, for
K=128 the leading chains of octave £-rings are

1 - 2 - 4 - 8 - 1 6 - 3 2 - 6 4 - 128-ring
3 - 6 - 12 - 24- 48 - 96-ring
5 - 1 0 - 2 0 - 4 0 - 8 0 -ring
7 - 14 - 28 - 56 - 112-ring
9 - 1 8 - 3 6 - 7 2 -ring
11 - 22- 44 - 88 -ring

127 -ring (3)


274 J. LEE

Each chain of octave k-rings begins with an odd wave


number; its length becomes shorter; and the last chain
has only the 127-ring.
Let us denote by ATS(K,N) the abridged truncation
system of ITS(K), which includes the first N longest
chains of octave *-rings. Although it is tempting to
leave out all but the longest chain of octave *-rings a
la the so-called cascade models of the two-dimensional
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

flow,23"25 it has been found that ATS(K,1) does not


contain enough triad-interactions to emulate correctly
the dynamics of ITS(K). Note, however, that the simi-
larity of ATS(K,1) and cascade model is only superfici-
al. The ATS(K,1) retains all of the triad-interactions
over the wave vectors of the 1-2-4-8-16-...-ring. On
the other hand, the cascade models involve a further
consolidation of ATS(K,1) in that the myriad of triad-
interactions is replaced by a local interaction among
the closest neighboring modes. We have shown in Table
1 how rapidly the number of triad-interactions grows,
as N is raised up to 9. For the first five chains of
Eq. (3) the octave brings of ATS(128,5) is given by:
1-2-3-4-5-6-7-8-9-10-12-14-16-18-20-24-28-32-
36-40-48-56- 64-72-80-96-112-128-ring (4)

And for ATS(128,6) the inclusion of the sixth chain of


Eq. (3) inserts four additional *-rings into Eq. (4),
and so forth. Finally, the parent system ITS(128) is

Table 1 Abridged truncation systems for K»128

No. of No. of
wave vectors triad-interactions Notation
5 2998 992,860 ATS(128,5)
6 3504 1,581,372 ATS(128,6)
7 4112 2,470,492 ATS(128,7)
8 4832 3,689,896 ATS(128,8)
9 5208 4,827,696 ATS(128,9)
MODELS FOR TWO-DIMENSIONAL TURBULENCE 275

recovered when all 64 chains of octave *-rings are


introduced.
As seen from Table 1, ATS (128, 5) has about 6000
(real) equations of motion and the number of equations
increases roughly by 1000 as N is incremented by one.
One may therefore suspect that the large-scale systems
of ATS would share certain dynamical behavior with the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

ITS(128). However, this is an unproven suspicion.


Although by construction ATS(128,N) should closely
approximate the ITS(128) in the limit as NF»64, we are
in reality more interested in the opposite limit of a
small N for maximal computational savings. Therefore,
we shall consider N as a parameter of numerical experi-
ment and observe the spectral energy distribution that
is invariant to the variation of N. Perhaps a pragmat-
ic justification for ATS is provided by the development
of chaos in two-dimensional flows. In the presence of
chaos, one must give up the notion of accurately com-
puting the flow trajectory and hence attemps to compute
the time-average of flow variables from a pseudo-orbit,
which is all that trajectory computation yields.26'27
Since the triad-interaction structure is preserved,
ATS(K,N) conserves its energy and enstrophy in the
absence of damping and forcing:

E = (1/2) Zji/^l2 (5)

Q = (1/2) V*|un|a (6)

where u =U(K ) and the summation is carried over all


wave vectors included for a given N. It also obeys the
classical Liouville theorem

where u? and u* are the real and imaginary parts of u ,


i.e., u = u*+iu*. Eq. (7) states the invar iance of
measure during time evolution. Since Eqs. (5-7) are
obeyed by each triad-interaction, they continue to hold
276 J. LEE

when triad-interactions are deleted from ITS(K) in


triplets.
IV. Equilibrium Energy Distribution
Kraichnan18 has constructed the canonical ensemble
with E and ft as the quadratic constants of motion to
deduce the energy-enstrophy equipartition, which in the
energy spectral form becomes
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

a*
f (*) = ————r (8)

where a and b are assumed non-negative constants. The


existence of canonical distribution requires that there
are sufficiently many degrees of freedom to validate
the Central Limit theorem. Using the ratio *2/*2 , where
*2= ft/E and k =K, Fox and Orszag2 showed that Eq. (8)
reduces to k in a large k region for *2/*2<0.1, and
k* in the entire k range for *2/*2=0.5. We have cate-
gorized in Table 2 the inviscid two-dimensional flow
simulations of previous investigators in terms of
*#/*2. When possible attempts were made to express the
total evolution time in units of the characteristic
eddy-turnover time given by r^}"""1'2 (which differs from
McWilliams9 by a 2rr factor). Note that the previous
simulations have attained the enstrophy equipartition
for *2/'*2 = 0.078, 0.11, and 0.17 and the energy equi-
partition for *2/*2= 0.48 and 0.55.
As indicated in Table 2, we have restricted K to 64
for the present inviscid flow simulation. The differen-
tial-equation solver ODE of Shampine and Gordon28(which
is also core module of the solver DEABM of the SLATEC
library) was used for time integration with the error
tolerance of 10~7. We have evolved ATS(64,3) from the
localized initial conditions <5(5) and <5 (40) , defined by:

( u =Cexp(i7r/4) for the r-ring, where C is


normalized so that E=l,
u =0 for all other k-rings
MODELS FOR TWO-DIMENSIONAL TURBULENCE 277

Table 2 Inviscid two-dimensional flow simulations

Source E(k] Comments


i
Fox and Orszag 0.11 Fig. 3; ki =64; evolution
(1973)2 time=2.5.
0.48 Fig. 2; k1 =64; evolution
time=2.5.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Seylor, Salu, 0.17 Fig. 3; run C; *|= 37.5;


Montgomery and k12=220; evolution time=34
Knorr (1975)3 or 59r.

Basdevant and 5 0.078 Fig. 5a; ^=2.43; k1 =8.71;


Sadourny (1975) evolution of 98,OOO steps.
0.55 Case B of Fig. 6; **=5.95;
k =8; evolution of 40,000
steps.

Present 0.0065 Fig. 3; *!=26.43; k =64;


simulation evolution time=1000r.
0.39 Fig. 4; ^ = 1603.3; k. =64;
evolution time=1000r.

Since 6(5) and 6(40) have the initial Q of 26.43 and


1603.3, they give rise to **/**=0.0065 and 0.39, re-
spectively (Table 2). By averaging over each *-ring in
Fig. 3, the equilibrium energy spectra sampled at 250r,
500r, and lOOOr are shown in Figs. 4 and 5. Clearly,
the evolution time of lOOOr is long enough for
ATS(64,3) to equilibrate toward the energy-enstrophy
equipartition, although fluctuations persist in the
small k region. To further suppress fluctuations, the
evolution of Fig. 4 was continued up to SOOOr, thereby
engendering a smoother energy distribution, as shown in
Fig. 2 of Lee.19 To see the effect of truncation, we
have further evolved ATS(64,N) for different values of
N. A general observation is that equipartition is at-
tained more quickly with smaller fluctuations when N>3,
and vice versa when N<3.
278 J. LEE

log(energy)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 3 Modal energies of ATS(64,3) over the 1-2-3-4-5-


6-8-10-12-16-20-24-32-40-48-64-rings. Considerable
fluctuations exist within each *-ring. A representative
energy distribution is obtained by averaging over each

In closing, we wish to comment on the relevance and


accuracy of the inviscid flow simulation. First, al-
though a successful verification of equipartition is
encouraging, it has no direct bearing on the actual
spectral energy distribution when forcing and damping
are present. What we have shown so far is that our
numerical integration does not violate the basic con-
straints of Eqs. (5-7) and that ATS(64,3) contains
enough degrees of freedom to satisfy the canonical dis-
tribution. Second, under the error tolerance of 10~7,
total E and ft have remained constant within 0.0025% and
0.25% of their respective initial values over the evo-
lution of lOOOr. However, this does not mean that we
can accurately compute the true trajectory over a very
long time. Since E and Q are obtained by averaging the
modal energies and enstrophies, the observed constancy
MODELS FOR TWO-DIMENSIONAL TURBULENCE 279

-1
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-4 J______l_l_ I I I I______
10 100

Fig. 4 Energy spectra of inviscid ATS(64.3) evolved


from the initial condition 6(5) with *|//r=0.0065;
time-averaged over 10r. A dashed line represents k
0=250r (evolution time=49); O=500r (evolution time=
98); *=1000r (evolution time=195).

-4

Fig. 5 Energy spectra of inviscid ATS(64,3) evolved


from the initial condition <5(40) with *|/*2=0.39; 1time-
averaged over 20r. A dashed line represent^ the i .
0=250r (evolution time=6.5); O=500r (evolution time=
13); *=1000r (evolution time=25).
280 J. LEE

should be attributed to the shadowing property29 of the


two-dimensional Navier-Stokes equations. It should be
noted that monitoring E and ft cannot be a diagnostic
means for controlling the numerical errors in the two-
dimensional21 and three-dimensional30 inviscid flows.

V. Quasi-Stationary Dynamics
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Since the first numerical simulation of Lilly1 on a


64x64 mesh grid, the resolution of spectral computa-
tions has been refined with the availability of improv-
ed computing power. In Table 3 are summarized the pre-
vious two-dimensional flow simulations in free decay in
order of increasing resolution. The energy spectra have
been represented by k~^ with the exponent in 3<0<6.
Free-decay simulations do not in general evolve the
two-dimensional flow into a quasisteady state that is
universal in the same sense as Kolmogorov's inertial
cascade flow in three dimensions. Note that 0=3 repre-
sents the enstrophy cascade18'31 and 0=4 the steepening
of vorticity gradients in the two-dimensional flow.32
More recently, McWilliams9 discovered the emergence of
coherent and isolated vortices that can give rise to
exponents in 4</?<6, and Benzi et al.33 attempted to re-
late such ft with the vorticity distribution function,
which, in turn, induces the velocity field. However,
one must point out that McWilliams' observation9 of
coherent vortices, though confirmed by Benzi et al.,33
has not been shared by Brachet et al.10
Therefore, we conjecture that the persistence of
coherent vortices is possible when the initial energy
input with large Q/E (%300) quickly transports enstro-
phy toward the high k, thereby engendering accelerated
vorticity decay to yield a "quiet" two-dimensional flow
field (ft/E«20), which is sparsely populated by isolated
vortices (Table 3). Our conjecture is based on the fact
that a dynamical system can have several basins of
MODELS FOR TWO-DIMENSIONAL TURBULENCE 281

Table 3 Free decay two-dimensional flow simulations

Spectral
Resolu- exponent
tion3 Source p Q/E Final time

(64)2 Lilly (1971)5 3 37 1200 steps


Fornberg (1977)6 3->4b 4->l 6840 steps
(series 3)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(128)2 Deem and Zabusky 4 390+136 1.7


(1971)7
Orszag (1976)8 4 13.5 2

(256)2 McWilliams (1984)9 ^5,6 300+20 40

(512)2 Orszag (1976)8 3 13.5 2


Brachet , Meneguzz i , 4*3 5.7 35
and Sulem (1986)10

The resolution can further be refined to (1024)2 if


certain fJpw symmetries are imposed (Kida ; Brachet
and Sulem11; Brachet et al.10).
b
"->" denotes transition from the initial to final
stages.

attraction. Whether or not Q/E is a right parameter by


which to categorize the basins of attractors is yet to
be seen. However, what is more unsettling is the possi-
bility that the observed spectra of Table 3 are still
transitory. In view of this, it is necessary to exam-
ine a dissipative system with the external forcing suf-
ficient to counteract viscous damping, thereby equili-
brating the eddy energy transport towards a quasisteady
state in some wave number range.
Forced two-dimensional turbulent flow was first
simulated by Lilly1 to test Kraichnan's hypothesis18 of
the simultaneous backward energy and forward enstrophy
cascades about a forcing wave number k^. Since then
high-resolution simulations have been performed, as
282 J. LEE

Table 4 Forced two-dimensional flow simulations

Spectral
exponent,
Forcing E cas- Q cas-
K Source wave number cade cade
15 Fyfe, Montgomery. 7.4<*rf?<8.36 5/3 3
and Joyce (1977)13
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

32 Lilly (1969)1 7.5<*f<8.5 5/3 3

64 Basdevant, Legras, 6.5<**<9.5 —— 4 & 6a


and Sadourny (1981)14 r

128 Frisch and Sulem 13<*rf<17 5/3 6


(1984)1*
Herring and 18.5<*rf<21.5 5/3 4b
McWilliams (1985)16

256 Legras, Santangelo, 7<*rf<10 —— 3.5 & 4.2C


and Benzi (1988)17

a
/?=4 for POW, PRW, PRI, and SRI, and ft=6 for POI and
SOI.
b
Case 2.
C
0=3.5 and 4.2 are Case II and III, respectively. Note
that Gilbert34 has attributed 0=3.5 to the development
of spiral vortices.

summarized in Table 4. Although Lilly1 and Fyfe et


al.13 have reported the simultaneous development of
*~5/3 and *~3, their results are inconclusive due to
low resolution. On the other hand, the high-resolution
simulations do not generally support Kraichnan's pre-
diction of simultaneous cascades. We first notice that
the inverse energy cascade in k<*f has been observed
only by Frisch and Sulem15 and Herring and McWilliams16
Furthermore, ft is typically greater than 3 for the
quasistationary energy spectra evolving in k>k*> just
as in the free-decaying flow (Table 3). Hence, enstro-
phy cascade is evidently not realized.
MODELS FOR TWO-DIMENSIONAL TURBULENCE 283

According to Kraichnan,18 the enstrophy cascade


requires a constant flux of enstrophy through the iner-
tial range, in parallel to constant energy flux in
Kolmogorov's inertial range in three dimensions. It is
therefore necessary to sustain constant upward energy
and downward enstrophy fluxes along the wave number
range. To specify /(t) to maintain stationary energy
and enstrophy dissipation rates, we single out the two
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

distinct Fourier amplitudes of wave vectors "$e and t^.


Using the notations i/e=u(te), /e=/(1e), *e=|2el and
similarly for aCJ . /,CJ . and kCJ . a general form of forcing
has been derived19

where E and Q are injection rates of energy and enstro-


phy, respectively. Here, e=2i/Q and 7j=2i/lS are the
respective dissipation rates of energy and enstrophy,
where $=( 1/2)S **|L/ |* . For a steady-state simulation
c/E/cft =cK)/c(t =0 , no energy and enstrophy need to be
introduced; hence, Eq. (9) reduces to

* (10)

This is the stationary forcing for maintaining constant


E and Q , for it replenishes exactly the energy and
enstrophy lost by viscous dissipation.
A. Ens trophy Cascade
To implement the stationary forcing of Eq. (10), we
pick out a pair of wave vectors t^
e and 3w in the k*-i
ring, such that the absolute difference of |i/e|-|i/0| is
the largest. Clearly, the designation of te and "1^ will
not remain fixed for all times and shift from one pair
284 J. LEE

_g -2

gg
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Jr3

-4

-5
10 100
k
Fig. 6 Energy spectra of ATS(128,5) evolved from the
initial condition <5(8) under stationary forcing Eq.
(10) applied to the 8-ring; time-averaged over ST. The
dashed lines represent *~5'3 and *~3. OlOOr (evolu-
tion time=12.5); O=168r (evolution time=21).

of wave vectors to another as time evolution proceeds.


With the /_C and fOJf continually updated using the cur-
rent u^, we integrate the spectral equations of motion
by the solver ODE, as in Sec. IV, but with a relaxed
error tolerance of 0.0001. We always start the time
integration with i/=0.0001 and then adjust it periodi-
cally to ensure that the enstrophy dissipation wave
number5 k_ = (^/i/3 )i/6 does not exceed K=128.
ATS(128,5) was evolved from the initial condition
6(8) by applying Eq. (10) to the 8-ring. We have shown
in Fig. 6 the evolved energy spectra at lOOr and 168r,
which clearly demonstrate development of the *~3 energy
spectrum in k>k*. Although the enstrophy cascade range
is kf<k<k , the range of k"3 observed in Fig. 6 is 14<*
<40. Beyond £=40 the energy spectrum falls off more
steeply than -3, which is perhaps due to ATS(128,5)
MODELS FOR TWO-DIMENSIONAL TURBULENCE 285

having only five octave *-rings (Sec. VI). Because of


the use of stationary forcing (Eq. (10)), E and ft have
remained within 0.0025% and 0.0046% of their respective
initial values for the entire evolution time. We now
present in Fig. 7 the evolution of e and ??, sampled at
every time interval of 0.5. Since Q is invariant, the
energy dissipation rate E should also be constant, but
it is not in the figure because v has been adjusted
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

during the evolution to ensure * <128. On the other


hand, one finds in Fig. 7 that the enstrophy dissipa-
tion rate 77 builds up rapidly to a steady level and
then oscillates with small amplitudes. Just as in E,
the oscillation of T? is in part due to the adjustment
of i/, necessitated by ensuring * <128. After a rapid
equilibration, the enstrophy dissipation attains a con-
stant level, reflecting a stationary flux of enstrophy
through the inertial range.

0,02

10 15 20
evolution time
Fig. 7 Dissipation rates of ATS(128,5) evolved from
the initial condition <5(8) under stationary forcing Eq.
(10) applied to the 8-ring. They are sampled at every
time interval of 0.5; hence, fluctuations are not shown.
O=energy dissipation rate e ; 0=enstrophy dissipation
rate 77.
286 J. LEE

B. Simultaneous Energy and Enstrophy Cascades


In contrast to rapid attainment of k~3 , the back-
ward energy cascade requires a very (infinitely) long
time to develop the *~5'3 energy spectrum in k<k^ . This
is because the lower wave number limit *0(<*f) °f ener-
gy cascade propagates to zero wave number bjr8
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Observe that the range of *~3/3 is 4<*<*f in Fig. 6.


Moreover, fluctuations are significant because there
are fewer degrees of freedom in k<k^ than in k>k^.
Although one can estimate *Q=0.09 (with £=0.014) at t=
21, the energy spectrum of Fig. 6 lacks sufficient
buildup in *<4 to show complete development of *~5 .
To accelerate the backward energy cascade, we repeat
the simulation of ATS(128,5) but by using the forcing
function of Eq. (9), which permits an arbitrary energy
injection by ctE/dt*Q. As shown in Fig. 8, the energy
spectra in *<*f exhibit continual buildup toward *~5//3
due to the energy injection, whereas the energy spec-
tra in k>k* have remained more or less stationary.
This clearly indicates backward cascade of the inject-
ed energy, independent of the forward enstrophy cascade.
Because of finite energy injection, Fig. 9a shows a
steady increase in E. Although the enstrophy Q initial-
ly remains constant (»64.17) up to the evolution time
of 40 (since c(ti/dt=Q) , it too increases to 68.44 toward
the end of evolution in the figure. We suspect that
this increase (%6.7%) is due to the excessive energy
pumping rate. The accompanying evolution of e and T? is
shown in Fig. 9b. Note that the constancy of energy
dissipation E has been punctured by the adjustment of i/ ,
as required by * <128. Although the enstrophy dissi-
pation 77 falls off gradually after having attained a
steady level, the overall falloff is apparently not
significant enough to upset the quasiconstancy of ^ ne-
cessary for the enstrophy cascade.
MODELS FOR TWO-DIMENSIONAL TURBULENCE 287
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 8 Energy spectra of ATS(128,5) evolved from the


initial condition <5(8) under forcing Eq. (9) applied to
the 8~ring; time-averaged over 6r. We started the time
integration with c/E/c/t=0, but then increased it gradu-
ally to rfE/rft=0.14 at the end of evolution. The dashed
lines represent k~* and *~3. 0=100r (evolution time=
12.6); <>=430r (evolution time=52.8); *=820r (evolution
time=92.8).

VI. Concluding Remarks


By bootstraping with only five chains of *-rings,
we have shown that ATS(128,5) can simulate the simul-
taneous backward energy cascade in k<k^ and forward en-
strophy cascade in *>**. The immediate question incum-
bent upon us is, "How well does ATS(128,5) approximate
the parent system ITS(128)?" Unfortunately, we cannot
answer this directly because the memory requirement for
ITS(128) is too large to carry out an efficient in-core
trajectory computation at the present time. Although
there is no a priori estimate on the minimum number of
octave *-rings, it appears that the first half of 64
octave *-rings represented by Eq. (3) would be suffici-
ent to simulate correctly the dynamics of ITS(128), for
288 J. LEE
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

50 Q

0
0 20 40 60 80
evolution time

b) 0.02

evolution time

Fig. 9 Evolution of ATS(128,5) under the conditions of


Fig. 8. Data plotted are sampled at every time step of
0.2: a) O=energy E; 0=enstrophy Q and b) O=energy
dissipation rate e; 0=enstrophy dissipation rate TJ .
MODELS FOR TWO-DIMENSIONAL TURBULENCE 289
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 10 Energy spectra of ATS(128,6) evolved from the


initial condition <5(8) under forcing Eq. (9) applied to
the 8-ring; time-averaged over 6r. The dashed lines re-
present *~5/3 and *~3. 0=113r (evolution time=14.2);
O=213r (evolution time=26.2); *=436r (evolution time=
48.2).

the octave *-rings of the second half all have only one
*-ring. Therefore, this lends some support to including
only the long chains of octave *~rings in ATS(128,N).
However, to provide a partial answer, we have re-
peated the simulation of ATS(128,N) for N>5. Typical
energy spectra of ATS(128,6) and ATS(128,8) are pre-
sented in Figs. 10 and 11. From Figs. 8, 10, and 11,
we find that the k~* energy spectrum is attained in
k<k* toward the end of evolutions. On the other hand,
the energy spectra evolving in k>k^ should be examined
more closely. To this end, we have presented in Fig. 12
only the final-time energy spectra of ATS(128,N) for N=
5-9. In spite of fluctuations, one observes development
of the k~3 spectral range in k>k*. For N=8 and 9, how-
ever, the spectral range of k~3 stops at around *=40
290 J. LEE
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-5

Fig. 11 Energy spectra of ATS(128,8) evolved from the


initial condition 6(8) under forcing Eq. (9) applied to
the 8-ring; time-averaged over ST. The dashed lines re-
present *~5/3 and *~3. 0=41r (evolution time=5.1);
O=132r (evolution time=16.5); *=228r (evolution time=
27.8).

and thereafter the energy spectrum falls off more


steeply with the -4 exponent. We believe that the com-
bination of k and k energy spectra has not been ob-
served in a two-dimensional flow simulation (Table 4).
One may ask if the *~* spectral range encroaches into
the *~3 range, so that eventually the entire inertial
range would have the k~4 energy spectrum as N increases.
This, however, does not appear to be the case because
the braking point of the k~3 and k"4 spectral ranges
moves away from k* as N is increased. Further work is
needed to ascertain the fate of simultaneous existence
of the k^3 and k~* spectral energy ranges.
Before closing, we wish to settle an issue raised
in our preliminary report in which it was observed that
the k 3 energy spectrum extends far beyond k*, hence
MODELS FOR TWO-DIMENSIONAL TURBULENCE 291

-2

-2

-2

-2
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

-2

-3

-4
10 100
k
Fig. 12 The final-time energy spectra of ATS(128,N)
evolved under the conditions of Figs. 8, 10, and 11.
The solid and dashed lines represent *~3 and £~4, res-
pectively. The ordinate axis has been shifted up to
separate the energy spectra of ATS(128,N) for N=5-9.
•=ATS(128,5) at evolution time of 820r; Q=ATS(128,6) at
evolution time of 436r; *=ATS(128,7) at evolution time
of 285r; O=ATS(128,8) at evolution time of 228r;
0=ATS(128,9) at evolution time of 62r.

the k~*^3 and *~3 energy spectra do not abut at k^


(Figs. 4 and 5 of Lee19). Although not explained expli-
citly in that report, we were then forced to further
simplify the spectral equations of motion by thinning
out wave vectors in the *-rings that lie in the high k
range. This was necessitated by the small internal
memory of Cray XM-P/12. Since then, the code has been
run on a Cray-2 with sufficient internal memory to in-
clude all of the wave vectors in the *-rings. The nu-
merical experiment has shown that the expedient thin-
ning out is responsible for the anomaly previously ob-
served.
292 J. LEE

Acknowledgment
This work was supported by AFOSR Task 2304N1.

References
1
Lilly, D.K., "Numerical Simulation of Two-Dimensional
Turbulence," Physics of Fluids, Supp.II, 1969, pp. 240-
249.
2
Fox, D.G., and Orszag, S.A., "Inyiscid Dynamics of
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Two-Dimensional Turbulence," Physics of Fluids, Vol.


16, Feb., 1973, pp. 169-171.
3
Seyler, C.E., Jr., Salu, Y., Montgomery, D., and
Knorr, G., "Two-Dimensional Turbulence in Inviscid
Fluids or Guiding Center Plasmas," Physics of Fluids,
Vol. 18, July, 1975, pp. 803-813.
4
Basdevant, C., and Sadourny, R., "Ergodic Properties
of Inviscid Truncated Models of Two-Dimensional Incom-
pressible Fluid," Journal of Fluid Mechanics, Vol. 69,
June, 1975, pp. 673-688.
5
Lilly, O.K., "Numerical Simulation of Developing and
Decaying Two-Dimensional Turbulence," Journal of Fluid
Mechanics, Vol. 45, Jan., 1971, pp. 395-415.
8
Fornberg, B., "A Numerical Study of 2-D Turbulence,"
Journal of Computational Physics, Vol. 25, Sept., 1977,
pp. 1-31.
7
Deem, G.S., and Zabusky, N.J., "Ergodic Boundary in
Numerical Simulations of Two-Dimensional Turbulence,"
Physical Review Letters, Vol. 27, Aug, 1971, pp. 396--
399.
S
0rszag, S.A., "Turbulece and Transition: A Progress
Report," Proceedings of the 5th International Confer-
ence on Mumerical Methods in Fluid Dynamics, edited by
A.I. Van de Vooran and P.J. Zandbergen, Vol. 59, Lec-
ture Notes in Physics, Springer-Verlag, New York, 1976,
pp. 32-51.
9
McWilliams, J.C., "The Emergence of Isolated Coherent
Vortices in Turbulent Flow," Journal of Fluid Mechanics,
Vol. 146, Sept., 1984, pp. 21-43.
10
Brachet, M.E., Meneguzzi, M., and Sulem, P.L., "Small-
Scale Dynamics of High-Reynolds-Number Two-Dimensional
Turbulence," Physical Review Letters, Vol. 57, Aug.,
1986, pp. 683-686.
11
Brachetf M.E., and Sulem, P.L., "Direct Numerical
Simulation of Two- Dimensional Turbulence," Single- and
Multi-Phase Flows in an Electromagnetic Field, Progress
in Astromautics and Aeronautics, Vol.100,e3ited by H.
Branover, P.S. Lykoudis, and M. Mond, American Insti-
tute of Aeronautics and Astronautics, New York, 1985,
pp. 100-110.
MODELS FOR TWO-DIMENSIONAL TURBULENCE 293

12
Kida, S., "Numerical Simulation of Two-Dimensional
Turbulence with High-Symmetry," Journal of Physical
Society of Japan, Vol. 54, Aug., 1985, pp. 2840-2854.
13
Fyfe, D., Montgomery, D., and Joyce, G.,"Dissipative,
Forced Turbulence in Two-Dimensional Magnetohydrodynam-
ics," Journal of Plasma Physics, Vol. 17, June, 1977,
pp. 369-398.
14
Basdevant, C., Legras, B., and Sadourny, R., "A Study
of Barotropic Model Flows: Intermittency, Waves and
Predictability," Journal of Atmospheric Sciences, Vol.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

38, Nov., 1981, pp. 2305-2326.


15
Frisch, U., and Sulem, P.L., "Numerical Simulation of
the Inverse Cascade in Two-Dimensional Turbulence,"
Physics of Fluids, Vol. 27, Aug., 1984, pp. 1921-1923.
16
Herring, J.R., and McWilliams, J.C., "Comparison of
Direct Numerical Simulation of Two-Dimensional Turbu-
lence with Two-Point Closure: the Effects of Intermit-
tency," Journal of Fluid Mechanics, Vol. 153, Apr.,
1985, pp. 229-242.
17
Legras, B., Santangelo, P., and Benzi, R., "High-
Resolution Numerical Experiments for Forced Two-Dimen-
sional Turbulence," Europhysics Letters, Vol. 5, Jan.,
1988, pp. 37-42.
18
Kraichnan, R.H. , "Inertial Ranges in Two-Dimensional
Turbulence," Physics of Fluids, Vol. 7, July, 1967, pp.
1417-1423.
18
Lee, J., "Steady-State Simulation of 2D Homogeneous
Turbulence," Physica D: Advance s in F1 u id Turbu1ence,
Vol. 37, edited by G. Doolen, R. Ecke, D. Holm, and V.
Steinberg, 1989, pp. 417-422.
a
°Kells, L.C., and Orszag, S.A., "Randomness of Low-
Order Models of Two-Dimensional Inviscid Dynamics,"
Physics of Fluids, Vol. 21, Feb., 1978, pp. 162-168.
21
Glaz, H.M., "Statistical Behavior and Coherent
Structures in Two-Dimensional Inviscid Turbulence,"
SIAM Journal of Applied Mathamatics, Vol. 41, Dec.,
1981, pp. 459-479.
22
Fox, D.G., and Orszag, S.A., "Pseudospectral Approxi-
mation to Two-Dimensional Turbulence," Journal of Com-
putational Physics, Vol. 11, Apr., 1973, pp. 612-619.
23
Gledzer, E.B., "System of Hydrodynamic Type Admitting
Two Quadratic Integrals of Motion," Soviet Physics -
Doklady, Vol. 18, Oct., 1973, pp. 216-217.
24
Lorenz, E.N., "Low Order Models Representing Realiza-
tions of Turbulence," Journal of Fluid Mechanics, Vol.
55, Oct., 1972, pp. 545-563.
25
Yamada, M., and Ohkitani, K., "Lyapunov Spectrum of a
Model of Two-Dimensional Turbulence," Physical Review
Letters, Vol. 60, Mar., 1988, pp. 983-985.
294 J. LEE

26
Benettin, G., Casartelli, M., Galgani, L., Giorgilli,
A., and Strelcyn, J.M., "On the Reliability of Numeri-
cal Studies of Stochasticity I: Existence of Time Aver-
ages," Nuovo Cimento B, Vol. 44, Mar., 1978, pp. 183-
195.
27
Benettin, G., Casartelli, M., Galgani, L., Giorgilli,
A., and Strelcyn, J.M., "On the Reliability of Numeri-
cal Studies of Stochasticity II: Identification of Time
Averages," Nuovo Cimento B, Vol. 50, Apr., 1979, pp.
211-232.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

28
Shampine, L.F., and Gordon, M.K., "Computer Solution
of Ordinary Differential Equations," Freeman, San
Francisco, 1975.
29
Hammel, S.M., Yorke, J.A., and Grebogi, C., "Do Nu-
merical Orbits of Chaotic Dynamical Processes Represent
True Orbits?," Journal of Complexity, Vol. 3, 1987, pp.
136-145.
30
Lee, J., "Development of Mixing and Isotropy in
Inviscid Homogeneous Turbulence," Journal of Fluid
Mechanics, Vol. 120, July, 1982, pp. 155-183.
31
Batchelor, G.K., "Computation of the Energy Spectrum
in Homogeneous Two-dimensional Turbulence," physics of
Fluids, Supp. II, 1969, pp. 233-239.
32
Saffman, P.G., "On the Spectrum and Decay of Random
Two-Dimensional Vorticity Distributions at Large
Reynolds Number," Studies in Applied Mathematics, Vol.
50, Dec., 1971, pp. 377-383.
33
Benzi, R., Patarnello, S., and Santangelo, P., "On
the Statistical Properties of Two-Dimensional Decaying
Turbulence," Europhysics Letters, Vol. 3, Apr., 1987,
pp. 811-818.
34
Gilbert, A., "Spiral Structures and Spectra in Two-
Dimensional Turbulence," Journal of Fluid Mechanics,
Vol. 193, Aug., 1988, pp. 475-497.
Solving Partial Differential Equations via Boolean
Automata: Statistical and Deterministic Approaches
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

K. Dang Tran,* A. Cosnuau,t J. Ryan,$ and Y. Morchoisne §


Office National d'Etudes et de RecherchesAerospatiales, Chatillon, France

Abstract
Cellular automata (locally interconnecting cells evolving in time by
mutual interactions) are used to solve partial differential equations
discretized by a finite-difference scheme on a nonuniform rectangular
grid. Boolean representations of real variables are introduced, and
logical intrinsic computer functions are used to achieve algebraic
operations of the finite difference algorithm. Convection and diffusion
problems have been solved by this method on a CRAY XMP and on an
INMOS massively parallel computer. Statistical errors and computing
costs are discussed.

Introduction
To solve complex partial differential equations such as those used in
fluid mechanics, even with modern high-performing computers
(reaching 200 Mflops), it is unlikely that finite difference or finite
element codes will take less than 10-6 second per point and per time
step. To give an accurate description of a three-dimensional complex
turbulent flow with no a priori modelization, several million points
would be necessary, which could require a thousand hours. With the
advent of massively parallel computers (Connexion Machine, optical
computers), such costs could be diminished, especially if new concepts
and algorithms are developed such as cellular automata, which are
closer to the basic binary (Boolean) structure of the computer.
The cellular automata concept is based on the idea that very simple
local laws can be used to determine a complex global behavior. Until

Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Head Theoretical Aerodynamics Branch 2, Aerodynamics Department, B.P. 72.
t Engineer, Computer Sciences Department, B.P. 72.
^Engineer, Aerodynamics Departmen, B.P. 72»
§Numerical Methods Coordiantor, B.P. 72; also at LAN, Paris VI University, 75230
Paris Cedex, France.
295
296 K. DANG IRAN ET AL.

now, use of cellular automata for solving partial differential equations


(PDEs) rested mainly on physical models, such as lattice gas models
developed by Frisch, Hasslacher, and Pomeau 1 and by D'Humieres
and Lallemand2, the accuracy of which is checked a posteriori.
Applying this concept to binary representations of real numbers, we
develop here a numerical approach using Booleans and probabilistic
properties. PDEs are first locally discretized (by finite volume, finite
element, or finite difference methods) on a nonuniform rectangular
grid. The discrete equations are then considered as a statistical
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

average of binary relations: on a conventional computer using floating


point arithmetics, real variables are replaced by their Boolean
representations, and real operations are replaced by logical ones.
Resulting algorithms constitute what we call Boolean automata.
The principle of the method is given below. Two different algorithms
are applied to the convection problem, and results are given for
convection and diffusion problems. Statistical errors and computing
costs are discussed. Finally, implementation on an INMOS transputer
system is described.

Principle of the Method

Boolean Representation of a Real Number


All real numbers are supposed to be within the interval [-1, 1] by an
appropriate rescaling: \U\ £ [0, 1]. Let [U] be the Boolean
representation of the positive real number U € [0,1]. If E (x) is the
integer part of x, U is defined as a random sequence ([U]j)j=i, N of N
digits 0 and 1 where the number N^u of digits equal to 1 is determined
by the value U:Nlu= E (U*N + 0.5).
The real value associated with a Boolean representation is obtained
by taking the average of the bit values. For example, the real number
U = 0.23 in a 10-bit word can be represented by [U\] = 0000100001 or
[f/2l = 0100100000. Both representations are equivalent. There are
10!/8!2! = 45 equivalent representations. Take a set containing 2 ones
and 8 zeros. Each equivalent Boolean representation can be seen as an
arrangement of these digits along a line, but the probability of picking
out a 1 out of this set in any of its arrangements will always be 0.2.
Thus, [(/] can be seen either as a sequence with probability U or an
algebraic binary writing of U where the position of the digits has no
meaning. These two interpretations will lead to two types of
algorithms, statistical or deterministic. The accuracy of such Boolean
representations is l/N where JV is the number of bits in the word.
A real number U € [-1,1] will be split into a positive part C7+ € [0,11
and a negative part U~ € [0,11: U — U+ - U~ , and represented by the
Boolean doublet ([t/+],[tT]).
BOOLEAN AUTOMATA 297

Boolean Operations
Standard logical functions are used to operate on Boolean
representations [[/]: AND, OR, NOT. SHIFTC (circular shift)
displaces each digit to the left, the leftmost becoming the rightmost;
this operator will be used for shuffling Boolean sequences. CSMG
(Cray Scalar Merge) completes a bit-by-bit selected merge between
two Booleans [Al and [B] controlled by a third one [C], POPCNT
counts the number of 1 in [[/]. It will be used to obtain the real value
attached to the Boolean sequence.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

To translate these Boolean operations into real operations, a small,


simple dictionary can be established if and only if the Boolean
sequences are independent in terms of statistics (see, for example,
Ref. 3).
Iff/, V, W are three real numbers with Boolean representation [if],
[V],[W],then:
[IH OR [V] is equivalent ioU+V-UXV
[U] AND [V] is equivalent to U X V
NOT [U] is equivalent tol-U
CSMG ([U], [V], [W]) = ([W\ .AND. [U]) .OR. (NOT .[W] .AND. | V])
is equivalent toW X U + (1-W) X V.
All these operations will appear in the schemes we use as we shall see
below.

Statistical Independence
In the case of nonlinear convection problems involving computation
of products U.U (see "Application") following a statistical viewpoint,
we need two statistically independent representations ([U] and [U]*)
of U to achieve a correct probability of the square. To meet this
requirement, a second representation of [IH is obtained by a random
mixing of the Boolean representation. A deterministic algorithm
relieves this necessity by dissociating advecting and advected
velocities.

Application to a One-Dimensional Convection Problem

Continuous Problem
Let us consider the one-dimensional periodic nonlinear convection
problem on [-1,11 X [0,T]
dU 3U
— + U— - 0
dt dx
(1)
U(x,t=0) =UQ
[/(-!, t) =U(l,t)
298 K. DANG IRAN ET AL.

Finite Difference Scheme


Let U" = U (xi, n&i) represent the velocity at grid point Xf and time
tn = nkt. An upwind finite difference approximation of the space
derivative and a Euler time scheme will give:
rrfl"l"l Jjn TT^ TT^
n
r rn ^^ f\
i A*
+u 1
X. - X .
' ,
-- un
i i-i (2)
n n
U n +l
- Lj 1 n
u - Un
+u
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Un < 0 ————— : 0
' Af *i+l-*i
Setting
A * A ^
A
nn un < (3)
A 1Al. — U .
Jjfl
l
,
x i -x i-l/ 1 l
' **-jl u-i
+J .iil -*•
-*•iI
The following equations are obtained

t/
'+A'f/'-1 (4)

As can be seen, in this particular problem, the sign of the convection


velocity U n must be known and its value must be split into a positive
part U+ and a negative part U~ as mentioned earlier. Equation (1) is
then replaced by:

———-^ + (U+ -U-)——————-=0 (5)


dt ax
which splits into
3U+ . dU +
—— + (U+ - U') —— = 0
dt dx (6)
dU~ ^ dU~
= 0
dt dx
and becomes
+
17?i >0 i/" +1 +
= (l-A")Un +
+AnU"
(7)
Uin ~ = 0 L (7i
n+ 1
~ = (1 - A ni) Uin ~-AniUni-i~

= 0
I ,-M (8)

1 1 n n
0 I i/"^ - = (1_B")[7' ~+B U ~
L z i i i f+1
These equations are more amenable to a Boolean treatment.
BOOLEAN AUTOMATA 299

Statistical Boolean Treatment of the Nonlinear Terms


When computing nonlinear terms such as A. n U", A*1 f / ^ j
in Eqs. (7) and (8), products Uf U^v Uf U^ and I//1 t//1 have to be
performed. The Boolean representation of U" , t/^j, and t/." 1, are
statistically independent. The first two products can therefore be
performed straightforwardly. But to achieve a correct probability of
the square U n U n we need two statistically independent represen-
tations of U? : ([I// 1 + UI//1 - D a n d t f t / / 1 + ]*, [I/.11 ']*). In the
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

statistical treatment proposed here, a random mixing of the previous


Boolean representation is used to obtain the following statistical
Boolean formulation:

[Un +
= CSMG ([(/" ,+ J , [Un +
J . , [An I )
1 fJ > 0 1-1 j ' j 17
(9)
[Un -]* = _ .

[U" +1* = 0 "+1+] [t/!1 +


] [Bp*)
(10)
[Un. ~ I* > 0 [U" +l +
] = CSMG ([[/^ ~ J , L t/ J1 "" ] [Bn. J*)

with

[A
" ]j =
X. — X .
.AND. [Un +
]* (ID
t Ji
z i—I

and
(12)

Deterministic Boolean Treatment of the Nonlinear Terms


The main inconvenience of the statistical treatment is the need for
random mixing of the Boolean representations at each time step, a
time-consuming operation that precludes any possible parallelization
of computations at the binary digit level.
A deterministic solver under study has been successfully applied to
the convection case. In this method, the independence of advected and
advection velocity is obtained by decoupling the convection Eq. (1) into
300 K. DANG IRAN ET AL.

two advection equations


dU_ dU
v— =
dt dx
dV dV
— + £7— = 0
dt dx (13)
U(x,t= 0 ) = V(x, * = 0 ) =
U(- l,t) = 17(1,0
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

which give the Boolean equations:

(14)
= CSMG _j

Random mixing of the third argument of the CSMG function if no


longer necessary to ensure statistical independence of advected and
advection velocities: numerical experiments show that a simple
circular shift to the right for [V*] and the same to the left for [LT n | is
sufficient.

10-1 -

10-2 -

10-3 -

10-4-

—1—
64 640 6400

Fig. 1 Statistical error versus the number of digits


of the Boolean representation.
convection case, statistical method
- convection case, deterministic method
diffusion case, statistical method
BOOLEAN AUTOMATA 301

Statistical Errors and Computing Costs


Comparisons between the finite difference solution and the two
Boolean solutions of the one-dimensional convection problem have
been performed: the initial condition is U (jc, 0) = sin (n x), x € [0, 1],
and a uniform spatial grid of 200 points is used with a time step At =
0.005 determined by the stability condition. Numerical experiments
show that the average quadratic errors (between finite-difference and
Boolean solution) are bounded in time. The evolution of these errors
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

versus the number N of digits of the Boolean representation is shown


in Fig. 1 after 100 time-steps for each Boolean solution. As can be
seen, the deterministic method is much more accurate than the
statistical one. The error is 0 (l/N) for the former, whereas it is
O(1A/N) for the latter.
Computing costs are given in Table 1 and Fig. 2. It can be seen that
the computing time per point and per time-step is proportional to N;
the random mixing used in the statistical method is responsible for the
cost of the method. The cost of the associated first order finite
difference scheme is of 10 7 s/point/time-step on a CRAY XMP, which
is similar to the cost of the Boolean determinist algorithm using
128 bits.

Extension to Higher Dimensions


To deal with a higher dimension convection equation, the
continuous equation is split along each space direction using a
fractionnary step method (Ref. 4).
For example in the two-dimensional case, if (U\, U%) are the two
velocity components in the (x, y) space, the split convection equation

Table 1 Computing costs for the convection case


(seconds/point/time-step versus number
of digits of the Boolean representation).______

Statistical method
——————————————————————— Deterministic
N Without With method
Random Mixing Random Mixing

128 1.4 X 10-7 2.6 X 10 ? 1 X 10-?


256 2.56 X 10-7 5.52 X 10-7 2.1 X 10 7
512 5.2 X 10-7 11.2 X 10-7 4.1 X 10 7
302 K. DANG TRAN ET AL.

10-6-
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

5.10-7-

1.10-7 -

I
128 256 512 N

Fig. 2 Computing costs (seconds per point and per time-step).


of the Boolean representation.
——•—— convection case, statistical method
- convection case, deterministic method
-\—— diffusion case, statistical method

will be the following:


dU dU
+ U. —— = 0
dt a* (15)
dU dU

dt dy

=0
dt doc (16)

—- + U^2 —— - 0
dt dy

Then each equation is translated into Boolean formulation in the


same manner as described above.
BOOLEAN AUTOMATA 303

Application to a One-Dimensional Diffusion Problem

Continuous Problem
Let us consider the one-dimensional periodic diffusion problem on

dU
— _v——
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

=0
dt dx2
U(x,t = 0) = U(0 (17)

U(-l,t) =

Finite Difference Scheme


Using a centered finite difference approximation of the space
derivative and a Euler time scheme, we obtain:
n n
i 17"1 + 1 -217"+
i Ut- _
(18)

Setting
2A<
(19)
- x)

Eq. (18) can be rewritten:


(Un , - 2 Vn + U" )
1 (20)
- = U" + A" -^————-—————

This equation can be rewritten in the following way, showing up two


CSMG.

Un +l = (A -An)U
I
n
I
+ AnI ((l - 1/2) t/" , + 1/2 Un (21)

which leads to the Boolean algorithm :

[V.\ = CSMG {[Ul_ J. , [Uni+


(22)
. ., ,
304 K. DANG IRAN ET AL.

Table 2 Computing costs for diffusion case

Statistical method

N Without With
Random Mixing Random Mixing

128 1 X 10 ? 2 X 10-7
256 2 X 10-7 4 X 10-7
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

512 4 X 10-7 8 X 10-7

Statistical Errors and Computing Costs


Comparisons between finite difference and Boolean solutions of the
previous one-dimensional diffusion problem have been performed: the
initial condition is U (x, 0) = sin2 (n x), x € [0, 1], v = 0.003; the time-
step is A£ — 0.005; 2 X 128 bits are allocated to each Boolean
representation; and 32 points in space are used. As for the convection
case, the average quadratic errors are bounded in time. Evolution of
this error versus N is shown in Fig. 1 after 100 time-steps.
Computing costs are given in Table 2 and Fig. 2. It can be seen that
the diffusion scheme for the statistical method is more efficient (Fig. 2)
though of similar accuracy as the convection scheme (Fig. 1).
The cost of the associated first-order finite difference scheme is 10 7
s/point/time-step on a CRAY XMP, which is similar to the cost of the
Boolean algorithm using 128 bits with no shuffling.

Extension to Higher Dimensions


To deal with a higher-dimension diffusion equation, the continuous
equation is split along each space direction, as in the convection
problem. In the two-dimensional case, if (U, V) are the two velocity
components in the (x, y) space, the split diffusion equation will be the
following:

BU
— ~ v —— = °
dt fa
2 (23)
dU au
— - v —— =0
dt ay2
BOOLEAN AUTOMATA 305

\ /•

)——————— <
J-1
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 3 Parallelism at the digit level.

av _ _
dt to2
dV (24)
-v -0

Then each equation is translated into Boolean formulation as


described above.

Implementation on a Massively Parallel Machine

Generalities
Three levels of parallelism for the present method may be exploited:
1) If no complex m a n i p u l a t i o n of the digits of Boolean
representations is needed (no random mixing in particular), each digit
[U"] or each spatial ensemble of jth digits [U "\ , / = i, i + L may be
allocated to a processor and computed in parallel (see Fig. 3).
2) Postive [U "*"] and negative elements [t/yi of the Boolean doublets
([U- + ], [U']) may be computed in parallel. If several 64-bits words are
used for the Boolean representations, these words may also be
computed in parallel.
3) Finally, a more classical geometrical parallelism (not particular
to this method) can be exploited by splitting the computational
domain into subdomains (slices, pencils, or cubes), the interiors of
which are handled in parallel with the others with information being
transferred through interfaces between subdomains.

Implementation on a Network of Transputers


The statistical approach for the diffusion problem above has been
extended to the three-dimensional diffusion case and coded in
306 K. DANG IRAN ET AL.

OCCAM2 on an INMOS ITEM 400 system of 40 transputers T800-20


distributed on 10 boards, using 64-bits words for Boolean
representations. Each transputer has 4 Kbytes on chip internal
memory and has access to its own 256 Kbytes external memory. On
each board, eight links may be manually connected to other
transputers to build the required network. For this particular parallel
machine, due to granularity considerations (amount of time spent on
communications versus computing in a parallel program), only the
third level of parrallelism has been exploited.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Three-Dimensional Distributed on a Ring of Processors


The physical three-dimensional domain, with n\ — 200, n% =10,
713 = 10 discretization points in jc, y, and z directions, is split in x
direction into 40 slices, so that an equal number (ni/40 — 5) of (y, z}
planes is allocated to each of the 40 transputers distributed on a ring
(Fig. 4). If n2 is the number of points of each subdomain, O (ft 3 )
computations and O (n2) communications are required.
In the case where five (yy z) planes are allocated to each transputer
(n\ = 5 X 40 = 200, n% — 10, 713 = 10), communications are
completely overlapped by computations and maximum speed-up is
reached, leading to a computing cost of 7.1 X 10-5 seconds per active
point and per time step. (An active point is a grid point with attached
unknown values of the variables.)

Three-Dimensional Domain Distributed on a Two-Dimensional


Array of Processors
The physical three-dimensional domain with n\ X n% X 713 = 54 X
40 X 28 discretization points is now distributed onto a two-
dimensional array of 9 X 4 — 36 transputers (Fig. 5). In the x
direction, the domain is split into 9 slices containing ni/9 — 6 (y, z)

Fig. 4 Domain distributed on a ring of transputers.


BOOLEAN AUTOMATA 307

Fig. 5 Domain distributed on a two-dimensional array of transputers.


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

planes. Each slice in the y direction is divided into 4 subdomains


containing n\/9 X fi2/4 X t t 3 = 6 X 1 0 X 2 8 active points. In fact, for
communication purposes, (n\/9 + 2) (n%/4 + 2)(n^ + 2) points are
allocated to each transputer.
More convenient for three-dimensional calculations, the use of a
two-dimensional array of processors leads also to a complete
overlapping of communications by computations and to a computing
cost of 1.71 X 10 5 seconds per active point and per time-step.
Compared to the cost obtained by simulating the same problem on the
Cray XMP (10-7 seconds per point and per time-step, using n — 64 bits
for the Boolean representation), the transputer system is 170 times
slower than the Cray.
Nevertheless, due to the complete overlapping of communications
by computations, the cost per time-step on this system is independent
of the total number of discretization points. For example, 1.037
seconds per time-step are needed if 33,600 points are computed,
allocated to 20 processors. Exactly the same time is needed if 60,480
points are computed, allocated to 36 processors. Thus, optimum speed
up of parallelism is reach* \ with this multidomain splitting applied of
the Boolean automata me od.

Conclusions and Perspectives

The proposed Boolean automata method can be implemented on


massively parrallel systems of elementary interger or Boolean
processors. In the case of the statistical approach, random mixing
within a word and between several words introduces a waste of time
and communication difficulties if Boolean processors are used.
Moreover, the accuracy in this case is only O (l/\/5v) if N is the total
number of bits used for the Boolean representation. Sensible
improvements have been obtained with the deterministic approach in
the convection case: accuracy is now O (l/N) and random mixing is no
more necessary. An equivalent approach is being studied in the
diffusion case.
308 K. DANG TRAN ET AL.

Test cases achieved on a transputer system, where only one


possibility of parallelism of the method is exploited, show that
complete overlapping of communications by computations can be
obtained leading to a decreasing computing cost per point and per
time-step versus the total number of discretization points.
For the moment and on computers not adapted to Boolean
operations (i.e., a CSMG is more expensive than the associated
floating point operation), costs of the Boolean automata technique are
similar to those of a first-order (in time and space) finite difference
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

scheme. On future massively parallel machines with low memory


processors optimized for simple binary operations, one can hope for a
great improvement of the ratio of Boolean costs over finite difference
costs.
Numerical schemes for the Navier Stokes equations adapted to the
Boolean automata formalism remain to be found and are the object of
current research. Feasibility tests on the "Connection Machine" are
also planned in the near future to exploit the other original
possibilities of parallelism of the method.

Acknowledgments
Financial support by DRET (Direction des Recherches Etudes
Techniques) for this work under contract is acknowledged.

References
iFrisch, U., Hasslacher, B., and Pomeau, Y., "Lattice Gas Automata
for Navier-Stokes Equation", Physical Review Letters, Vol. 56, 1986,
p. 1505.
2
D'Humieres, D., and Lallemand, P., "Lattice Gas Automata for
Fluid Mechanics", Physica 140A (Netherlands), Proceedings of the
16th International Conference on Thermodynamics and Statistical
Mechanics, Boston, 1986, pp. 326-335.
SPapoulis, A., Probability, Random Variables, and Stochastics
Processes, McGraw-Hill Series in Systems Science, McGraw-Hill,
New-York.
4
Yanenko, N. N., The Method of Fractional Steps, Springer-Verlag,
New York, 1971.
Rag Theory of Magnetic Fluctuations
in Turbulent Flow
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

A. A. Ruzmaikin*
Institute of Terrestrial Magnetism, Ionosphere, and Radio Wave Propagation,
Troitsk, Russia

Abstract

A framework for describing fluctuating magnetic fields in a


turbulent flow is developed. Issues as the temporal growth threshold in
the subcritical region where the Reynolds number is less than some
critical value, the role of boundaries, and the creation of intermittency by
a fast dynamo effect are addressed. Since the magnetic field structure in
a random flow is intermittent in the large Reynolds number limit, the
behavior of the magnetic field cannot be characterized solely by the
mean field and the second moment. A ratio of the moments is required
as is the distribution function of the random field, where possible.

Introduction

The dynamo problem is historically and mainly aimed at the origin


of large-scale cosmic magnetic fields. Presently, the study of small-scale
fluctuating magnetic fields is of great interest. The small-scale fields are
generated simultaneously with the large-scale fields in planetary cores,
stars, and galaxies, and they can be excited by turbulent motions even
without large-scale magnetic fields. The second possibility is of interest
in application to the clusters of galaxies1 and, in particular, to the
problem of magnetic field generation in large technical volumes of liquid
metal, for example, in the breeder reactors/
Fluctuation of magnetic fields in a turbulent flow of weakly
conducting fluid was first considered by Golitsyn.3
For dynamo excitation of the magnetic fluctuations in infinite fluid,
only a sufficiently large magnetic Reynolds number is needed. In a finite

Copyright <£> 1992 by the American Institute of Aeronautics and Astronautics, Inc.
All rights reserved.
*Professor, Chief of Laboratory of Magnetic Cosmical Research.
309
310 A. A. RUZMAIKIN

region, an additional condition appears. As a result of the diffusion of


the magnetic field through the boundaries a portion of the region must
be large enough to compare with a correlation length of the turbulence.
The magnetic Reynolds number is small in the laboratory or
magnetohydrodynamic (MHD) generators and very large in
astrophysical conditions, for instance in the solar case (108) (see Fig. 1). It
is interesting, however, that the Earth core and the modern breeder
reactors have comparable magnetic Reynolds numbers. Whereas the size
of the Earth's core is 105 times larger as compared with the size of a
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

breeder reactor, the turbulent flow in the breeder is moving faster to the
same factor. The planetary core is also similar to the metallurgy. The
main processes that drive the dynamo occur in a melted iron-sulphur
alloy freeing in the central part of the core.

Temporal Growth Under Threshold

To make the dynamo, i.e., the self-excitation of the magnetic field,


operative, it must pass a threshold, the so-called critical magnetic
Reynolds number. This number is determined theoretically by a
structure of flow and geometry of magnetic problems under
consideration. Typically, it is not less than several tens. Of course, there
is no problem in astrophysics and geophysics. However, the magnetic
Reynolds numbers that can be reached in a laboratory or technical plant
hardly exceed the critical one. What about the subcritical regime?
It is well known that a scalarlike temperature in the turbulent flow
without a source (for the scalar) will monotonously decrease with time.
Zeldovich4 and Pouquet5 noted an initial growth of the magnetic field in
a two-dimensional turbulence. The temporal growth of magnetic energy
in the three dimensional short-correlated turbulent flow was also
demonstrated (see Fig. 2).6
It may seem that a second critical magnetic Reynolds number exists
dividing the monotonous decrease and the temporal growth. However,
this number is dramatically dependent on a scale of the initial field
distribution. In fact, imagine a magnetic field of very large scale. Initially,

Lab Breeder Earth core Astrophysics


1———————————————^2 ^ Rfl

Fig. 1 Typical values of magnetic Reynolds number for different


conducting fluids.
RAG THEORY OF MAGNETIC FLUCTUATIONS 311

a) W (t)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

W max
b)
TV

4 --

2 ••

20 40

Fig. 2 a)The time dependence of the magnetic energy density for


increasing magnetic Reynolds numbers (from curve 1 to 3, R m
increases from 15 to 55); b), c) The maximums of W and a time of its
achievement vs R m .
312 A. A. RUZMAIKIN

this field evolves only under the action of turbulent velocity. The Ohmic
diffusion can be neglected until the scale decreases to a proper size,
depending on the magnetic Reynolds number. Namely, in this period of
evolution, the magnetic energy can rise, however, not exponentially, but
in, a power law (linearly). Mathematically, the field evolves as the scalar
plus a source determined by velocity deformations, it gives (at some
initial stage) a solution of the form t exp(-ydt) where Yd is the first
eigenvalue of the pure diffusion problem.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Diffusion of Magnetic Spots

Diffusion and dissipation of a large-scale (mean) magnetic field in


turbulent flow are determined by a turbulent diffusivity Vt (see, for
instance, Ref. 7). A characteristic time for the diffusion of the field of
scale A is A 2 /H. By considering the magnetic field at any scale, this
conception cannot be used directly. Instead, it is reasonable to treat
magnetic diffusivity as a function of the scale A,8 say,
if

and
D = vm + v t if X>£

i.e., it is the molecular diffusivity at small scales and the turbulent


diffusivity at large scales. Typically, vt / v m = Rm » 1.
This can dramatically change the characteristic time of the magnetic
field evolution in the turbulent flow. Imagine an initial field
concentration of a small scale A compared with a character scale of the
turbulent flow /. Then first the field diffuses under the action of the
molecular diffusivity, i.e., very slowly and only after the scale of the field
becomes comparable or larger than / does the field diffuse under the
action of the turbulent diffusivity, i.e., very fast.
To estimate a characteristic diffusion time, consider two close fluid
elements transporting a magnetic line. A mean-squared distance
between them evolves as d<r 2 > = 2Ddt. The integration from A to /
gives

Role of Finite Space

In a simple approximation, the finite turbulent fluid can be


described by an isotropic mirror symmetric short-correlated
nonhomogeneous velocity field. This velocity field is considered as given
RAG THEORY OF MAGNETIC FLUCTUATIONS 313

in the form of its correlation function, including a function describing the


nonhomogeneity (i.e., finiteness of the region).

(v t (x, 1,^(^12)) = qK^VyWSdj-ta)

where R = (x + y)/2, r = x-y and Vjj is the isotropic part of the tensor .
The solution of the kinematic eigenvalue problem for a correlation
tensor <HjHj> with <Hp> = 0 in this velocity field gives a critical
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

magnetic Reynolds number (Rm)cr « 50 and a critical ratio Acr = (l/L)cr «


(6Rm -i- 2)~1/2, where L is the size of the region under consideration and 1
is a correlation length of the turbulence.^When the magnetic Reynolds
number is large but less than the critical one, a temporal growth of the
magnetic fluctuations is possible. This point may be of interest in the
technical applications.
In the more complex case, the magnetic fluctuations are generated
also by deformations of the large-scale magnetic field. The evolution of
the magnetic fluctuations in the short-correlated isotropic flow having
the mean helicity in the presence of a large-scale magnetic field can be
described by a system of minimum two coupled differential equations
for the magnetic correlation function and the mean helicity. When the
conditions for magnetic field excitation are satisfied an influence of the
large-scale magnetic field becomes negligible in the limit t -> <*.
However the influence of the large-scale magnetic field on the magnetic
fluctuations is sensitive in the subcritical regime.

Speed of Magnetic Energy Propagation

In the supercritical regime, the magnetic field undergoes two main


actions: Instability with a growth rate y and diffusion D. It gives a
specific speed v = 2(2yD)~1/2 for the propagation of a boundary of mean
magnetic energy distribution.8 Earlier this effect was pointed out in the
nonlinear problem of propagation of bacteria and flames.10
Crudely, in a one-dimensional approximation the magnetic energy
for a distribution of scale exceeding / is

w(t) = w(0) (4rcDt)-1/2 exp(2yt-x 2 /4Dt)

where D =vm + vt .
Let us start with an initial distribution around x = 0. At a sufficiently
large time, the behavior of the surfaces of constant magnetic energy is
determined mainly by the lines of constant phase

yt - x2 / 4Dt = const
314 A. A. RUZMAIKIN
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 3 Asymptotic behavior of a line of constant phase determining


the speed of magnetic energy density propagation.

One of these lines is depicted in Fig. 3. Asymptotically, as t —> °c, it goes


to x = ± Vt, where

V = 2(2-yD>1/2

Thus, the magnetic energy distribution in the turbulent flow neither


diffuses nor can be transported by flow velocity. It propagates with the
specific speed determined by the diffusivity and an a priori unknown
rate of the magnetic field growth.

Creation of Intermittency by the Fast Dynamo

It may seem naive to believe that the increase of the magnetic


Reynolds number over the critical value will automatically increase an
efficiency of dynamo action. However, this type of behavior is an
inherent part of the magnetic field growth rate only in a finite interval of
the magnetic Reynolds numbers. Then two principal possibilities for the
dependence y(Rm) are known: the slow dynamo, after passing a
maximum, the growth rate tends to zero as Rm goes to infinity, and the
RAG THEORY OF MAGNETIC FLUCTUATIONS 315

fast dynamo, when the growth rate approaches some finite limit.11 This
is a conception. What is the reality?
Laminar flows of conducting fluid can act typically as the slow
dynamos. An example is the Ponomarenko screw dynamo and its
extensions developed, in particular at Riga Institute of Physics.
Turbulent flows instead are able to act as the fast dynamos. Initially
it was demonstrated by many authors that the mean field dynamo in a
helical turbulent flow has a finite growth rate in the limit Rm -> <*. From
a mathematical point of view, this problem is relatively simple because it
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

can be solved by standard methods of mathematical physics.


It is shown by attracting numerical methods that even in the absence
of helicity or the mean magnetic field, a growth rate for a correlation
function of magnetic fluctuations tends asymptotically to some constant
value.2'12
Now we are confident that not only the mean or mean-squared
magnetic field, but almost each realization of random magnetic fields in
turbulent well-conducting flows is growing in the fast dynamo style.
Formally, it is shown that, for a well-mixed random flow with a finite
correlation time by means of extension of the central limit theorem to a
product of random matrices, each matrix is determined by the velocity
field in a given correlation interval.8'13
The magnetic field amplification by random velocity flow is well-
illustrated with the help of topological transformations of the type first
presented in the Zeldovich eight dynamo and then developed by Finn
and Ott,14 who took into account that the magnetic loop can be
transformed nonhomogeneously, for instance, only the a («1) part of
the loop is extended.
The mean field < HI > and the second statistical moment < HjHj >
are not sufficient to characterize the behavior of the magnetic field in the
turbulent flow. According to numerical experiments15 and the dynamo
theorem, the magnetic field structure in the random flow is intermittent
in the limit of large magnetic Reynolds numbers. This intermittency
means a progressive growth of the magnetic statistical moments with an
increass in the number of moments. It is important to find a ratio of the
moments (for instance, the flatness) and, if possible, the distribution
function of the random magnetic field. A simple model of the distri-
bution, taking into account the back action of the magnetic field on the
motions, is constructed.16
The intermittent structure apparently occupies space in some fractal
set. The dimension of a set created by one of the magnetic loops that
undergoes to eight-type transformations in the model by Finn and Ott14
is

= l + ln2/ln[a(l-a)]1-1/2
316 A. A. RUZMAIKIN

It is interesting to find a dimension of the set in the more realistic


statistical case.
Some applications of these theoretical results to geophysics,
astrophysics, and plants are known yet, the others it is hoped will appear
soon.

References
^Ruzmaikin, A. A., Shukurov, A. M., and Sokoloff, D. D., "The Origin of
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Magnetic Field in the Clusters of Galaxies/' Monthly Notices of the Royal


Astronomical Society, Vol. 241,1989, pp. 1-41.
2
Leorat, J., Pouquet, A., and Frisch, U., "MHD Turbulence Near Critical
Magnetic Reynolds Numbers/' Journal of Fluid Mechanics, Vol. 104,1981, pp. 419-
445.
3
Golitsyn, G. S. "Fluctuations of the Magnetic Field and Current Density in a
Turbulent Flow of a Weakly Conducting Fluid," Soviet Physics Doklady, Vol.5,
1960 pp. 536-539.
4
Zeldovich, Ya. B., "The Magnetic Field in the Two-Dimensional Motion of a
Conducting Turbulent Fluid," Soviet Physics JETP, Vol.4,1957, pp. 460-462.
5
Pouquet, A., Journal of Fluid Mechanics, Vol. 88,1978, p. 1.
^Maslova, T.B. and Ruzmaikin, A. A., "Growth of Magnetic Fluctuations in
Turbulent Flow," Magnitnaya Gidrodinamika, No. 3,1987, pp. 3-7.
^Moffatt, H.K., Magnetic Field Generation in Electrically Conducting Fluid, Cam-
bridge University Press, Cambridge, UK, 1978.
8
Molchanov, S.A., Ruzmaikin, A .A. and Sokoloff, D.D., "Kinetic Dynamo in
Random Flow," Soviet Physics Uspekhi, Vol. 28,1985, pp. 307-327.
9
Maslova, T. B., Shumkina, T.S., Ruzmaikin, A. A. and Sokoloff, D. D., "Self-
Excitation of the Fluctuating Magnetic Fields in a Finite Random Flow," Doklady
ofAcadamy of Science USSR, Vol. 294,1987, pp. 1371-1376.
10
Zeldovich, Ya.B., Barenblatt, G. I., Librovich, V. B., and Makhviladze, G. M.,
Mathematical Theory of Combustion and Explosion, Plenum, 1985.
^Zeldovich, Ya.B. and Ruzmaikin, A. A., "The Magnetic Field in Conducting
Fluid in Two-Dimensional Motion," Soviet Physics JETP, Vol. 51,1980, pp. 493-
497.
l^Novikov, V.G., Ruzmaikin, A. A., and Sokoloff, D. D., "Kinematic Dynamo in
Reflection-Invariant Random Field," Soviet Physics JETP, Vol. 58,1983, pp. 527-
532
^Molchanov, S.A., Ruzmaikin, A. A. and Sokoloff, D. D., "Dynamo Theorem,"
Geophysics & Astrophysics Fluid Dynamics, Vol. 30,1983, pp. 242-259.
*4Finn, J. and Ott, E., "Geophysics and Astrophysics FluidDynamics," Physical
Review Letters, Vol. 60,1988, p. 760.
15
Meneguzzi, M., Fricsh, U.,and Pouquet, A., "Helical and Non-Helical
Turbulent Dynamo," Physical Review Letters, Vol. 47,1981, pp. 1060-1064.
16
Dittrich, P., Molchanov, S. A., Ruzmaikin, A. A., and Sokoloff, D. D., "The
Limiting Distribution of the Magnetic Field in Random Flow," Magnitnaya
Gidrodinamica, No. 3,1988, pp. 9-12.
Ballooning Instability in Fluid Dynamics
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

M. Mond*
Ben-Gurion University of the Negev, Beer-Sheva, Israel
and
E. Hameirit
New York University, New York, New York 10012

Abstract
Ballooning modes are combinations of waves that propagate along their rays
which are the streamlines of the flow. A stream function representation is used
in order to derive a set of ordinary differential equations, which describe the
variation of the amplitude of the ballooning modes along the streamlines.
Because of such a representation the resulting equations are simple, and the role
of various physical effects is clearly displayed. As an example, the ballooning
equations are derived and solved for the case of cylindrical swirling flows, and a
necessary condition for the stability of such flows is obtained.
I. Introduction
The stability of columnar vortices has been under extensive investigation
over the past 20 years.1"6 The study of such flows is of interest when
investigating such topics as transition to turbulence, the effect of trailing
vortices on a flying aircraft, and1 fluid flow in various medical devices.
Eckhoff and Storesletten, investigating the linearized Euler's equations,
obtained a necessary condition for the stability of a cylindrical swirling flow by
considering waves propagating one dimensionally along their rays. They
showed that the streamlines form a family of rays along which instability may
occur. To show that, they used a generalized Wentzel, Kramers, Brillouin
(WKB) method and obtained a set of ordinary differential equations along the
streamlines for the amplitudes of the perturbations. However, the various
physical effects such as the pressure and the streamline curvature are obscured
in the resulting equations, and their method cannot be easily extended to more
general flows.
Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.
All rights reserved.
* Professor, Department of Mechanical Engineering,
t Professor, Department of Mathematics.
317
318 M. MONO AND E. HAMEIRI

Investigating the same problem, Bayly2 employed the stream function


coordinates in order to construct unstable modes that are localized about the
streamlines of the basic flow. In his work, however, the ray equations were not
used, and only periodic basic flows were investigated using the Floquet theory.
In this work we combine the geometrical optics technics with the flux
(streamfunction) coordinate representation in order to obtain a set of ordinary
differential equations that describe the amplitude of the modes that are localized
about the streamlines of a basic flow. This combination generalizes the works
of Eckhoff and Storesletten1 and Bayly2 and makes the role of such physical
quantities as the pressure, vorticity, and streamline curvature more transparent
in determining the stability of the basic flow.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Examples of modes that are localized about rays can be found in the
magnetohydrodynamics (MHD) literature (in both magnetic fusion as well as
in space physics applications).7*12 In systems described by the ideal MHD
model, the magnetic field lines form a family of rays for the Alfven as well as
for the slow magnetosonic waves. The coupling of these waves along the
magnetic field lines gives rise to modes that are called ballooning modes and to
possible instabilities. The ballooning modes are considered to be most
detrimental to magnetic confinement, and their properties are usually taken into
account when calculating optimal parameters for fusion devices design.
In this study we use the technics and physical insight gained by
investigating the MHD ballooning modes in order to study their
hydrodynamics (HD) counterparts. The analogy between ballooning modes in
plasmas and in classical fluids was recognized in Ref. 13. Thus, since the
streamlines of the basic flow form a family of rays for the entropy as well as
for two vortex waves, it is expected that the coupling of all of those waves
gives rise to the ballooning modes and possible instabilities. The sound waves,
corresponding to the fast 6magnetosonic waves in MHD, propagate along
separate rays and are stable.
The properties of the basic flow and the linearized equations are discussed
in Sec, II. In Sec. Ill the ballooning equations are derived, and in Sec. IV they
are solved for the particular case of swirling flow in order to obtain a necessary
condition for its stability.

II. The Basic Flow and the Linearized Equations


The fluid is assumed to be inviscid and compressible and to obey the
Euler's equations:

=0 (2)

+u-VS = 0 (3)

where u, P, p, and S are the velocity, pressure, mass density, and entropy of
the fluid, respectively, The last three quantities are related via the following
BALLOONING IN FLUID DYNAMICS 319

equation of state:

P=A(S)pY (4)

where y is the specific heat ratio and A a function of S. It is noted that the
continuity equation can be reproduced from Eqs. (2-4),
The basic flow is determined from Eqs. (2-4) by setting all partial
derivatives with respect to time to zero. A convenient form is:

pu-Vu = - VP (5)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

V-(pu) = 0 (6)

u-VS = 0 (7)

As a result of Eq. (6), pu can be represented in terms of two Clebsch


potentials:
pu = V\|/xVcc (8)
where \(/ labels cylindrical surfaces of streamlines whereas a may not be single
valued. (ItwillbefurtherspecifiedinSec.lv.) Both y and a are constant on
each streamline.
Before turning to the linearized equations we derive two useful results. We
first notice that as a result of eq. (7) S may be chosen to be a function of \|/
alone, i.e., S = S(\|/). The second relation is a Bernoulli law. In order to derive
it we first rewrite eq. (5) in the following way;

u x (V x u) = - VP + V Q- u 2 ) (9)

where u is the magnitude of the velocity. Equation (8) is now multiplied by u,


Using the equation of state (4) we obtain:

which yields
V P 1
(11)

where B is an arbitrary function of y.


We now linearize Eqs. (1-4) around the basic flow described by Eqs. (5-
11). The perturbed quantities are denoted by a subscript 1, whereas the
equilibrium properties remain without a subscript. We obtain:

J u-Vii = - VPi (12)


320 M. MONO AND E. HAMEIRI

3P
+ yPl V-u + yPV-ii! = 0 (13)

r)S
^L + u 1 -VS + u - V S 1 = 0 (14)

where a linear relation between p l f P lf and Sj can be obtained from the


equation of state Eq. (4).
At this point it is convenient to express all of the perturbed quantities in
terms of the lagrangian displacement £(x,t). The latter expresses the shift at
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

time t, due to the perturbation, of a fluid element from its position at the same
time under the basic flow. We thus have:

PI = - S'VP - yPV-5 (15)

Inserting now Eq. (15) into momentum Eq. (12), the following equation is
obtained for £:

p + 2pu-V + F(§) + H(P,) = 0 (16)

where

(17)

(18)
and
a(5) = u-V^ + ^-Vu (19)
Equation (16) is written in such a way that the terms involving derivatives
along die streamlines (i.e., terms that contain the operator u'V) are separated
from those with derivatives in the other directions. The latter appear only in
the operator H through PI as expressed in Eq. (15). The benefit of this
separation will be made clear in the discussion of the ballooning modes in the
next section.
III. Ballooning Modes
As was previously discussed, ballooning modes represent waves that1
propagate one dimensionally along their rays. Eckhoff and Storesletten
showed that the streamlines of the basic flow form a family of rays for three
waves: an entropy wave and two vortex waves. The coupling of these waves
along the streamlines gives rise to the ballooning instabilities. It is, therefore,
BALLOONING IN FLUID DYNAMICS 321

the aim of this section to obtain a set of ordinary differential equations, called
the ballooning equations, that describe the variations of the amplitude of the
waves along the rays.
The ballooning equations are obtained by employing the geometrical
optics approximation:

x(*,t)] ft° + e*l + -) (20)

Pl=exp£%(x,t)] (pJ + ePj + -..) (21)


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

where %(x,t) is the phase function of the wave, £°(x,t) its amplitude, and e«l
the measure of the fast variations of the phase. For the waves that propagate
along the streamlines, the phase function satisfies the following equation:

f* + u-Vx = 0 (22)

According to the ray equations,14 3%/3t is constant along the rays; hence,
% satisfies the following equation:

u-V X = C (23)
where C is an arbitrary constant.
At this point we investigate the case C=0. Even though this choice may
appear restrictive, it will be demonstrated in the next section that in the case of
swirling flow it is possible to extend the resulting stability criteria to the case
of general C by a proper Galilean transformation.
Having set C to zero, it is easy to solve Eq. (23) for % and to obtain with
the aid of Eq. (8)

X = X(W») (24)
Inserting Eqs. (20) and (21) into Eq. (16) and using Eqs. (22) and (24), it
is noticed that the leading order terms in e come from H(Pi). This implies that
P° = 0. In addition, Eq. (15) results in

£°-Vx = 0 (25)
Thus, to O(l) we have:

§ + 2 p u - V I + F ) P£ = PJ V X (26)

where the superscript was dropped from £0 and P is a projection operator that
annihilates the components along Vx:
322 M. MONO AND E. HAMEIRI

(27)

such that P J; satisfied Eq, (25),


Operating with p on both sides of Eq. (26) eliminates the unknown P ,
and the result is the following ballooning equations for the two components of
J; normal to V%:
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

£=O (28)

It is first noticed that the ballooning Eq. (28) contains spatial derivatives
only along the streamlines, i.e., of the form u-V£. It is, therefore, an ordinary
differential equation that should be solved separately along each streamline.
We turn now to simplifying further the various terms in Eq. (28). For that
purpose, it is noticed that the value of V% on any particular streamline can be
written as
V% = XV\|/ + Vet (29)

where - °° < X < °° is a parameter that determines the spatial orientation of the
perturbation. Equation (28) is, then, a one-parameter family of equations and,
hence, for each streamline, stability for all X's has to be checked.
As in the case of MHD ballooning modes,10 the following representation
for £ is used as a result of Eq. (25):
£ = XN + Yu (30)
where

In a similar fashion as in Hameiri10 it can be shown that

N-V\)/ = 1
and
a(§) - (aX+Y')u + X'N
where
a = 2K-N - ~2N - u x £
u
K is the curvature vector of the streamlines, and £ = Vxu is the vorticity.
Now, since Eq. (28) has only components parallel to N and u, it is
multiplied by the two latter vectors separately. The two following equations
are obtained:
BALLOONING IN FLUID DYNAMICS 323

P T- f INI2 T^ + 2pN-KG(\|/)X + 2pN-Ku 2 (~- - 2N-KX ^ = 0 (32)


dt\ di / \dT /
where

A-B-B (33)
the derivatives along the rays are given by

£•!*••* <34>
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

B is defined by Eq. (11), and the dot represents the derivative with respect to \|/.
Equation (31) can be integrated exactly. We choose initial conditions for Y
such that the last term in Eq. (32) vanishes. It then reduces to a single second-
order ordinary differential equation for X:

0 (35)

Equation (35) is the desired ballooning equation that describes the variation
of the amplitude X of the perturbations along the streamlines. It is a second-
order equation and describes the propagation of vortex waves along the rays.
The third, the entropy wave, was decoupled.
To summarize, Eq. (35) has to be solved separately for each streamline and
for all values of X (the K dependence appears through N). The stability of
these solutions is necessary for the stability of the flow.
IV. Swirling Flows
As an example, the ballooning Eq. (35) is investigated for obtaining a
necessary condition for basic flows that are given by:

(36)
where r, 9, and z are the cylindrical coordinates; £1, W, and p arbitrary functions
of r; and 9 and z the unit vectors in the azimuthal and axial directions,
respectively. The pressure is determined through the steady-state momentum
equation:
P = pr Q2 (37)

For flows described by Eq. (36), the streamlines are helixes that form a set
of nested cylindrical surfaces. Hence, \|/ is a function of r alone. The winding of
the streamlines around the cylindrical surfaces is described by a, which is given
by:
a = z - q(r)9 (38)
324 M. MONO AND E. HAMEIRI

where according to Eq. (8)


XI//V\

(39)

and q is proportional to the pitch of the helical line. If q varies from one
cylindrical surface to the other the flow is said to be sheared, whereas if q'(r) »
0 the flow is shearless. It is in the latter case that the flow can be considered to
be periodic and the ballooning equation has periodic coefficients [constants for
the flow described by Eq. (36)].
Before turning to the investigation of the stability properties of swirling
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

flows, the implications of the particular choice C=0 made in Eq. (23) are re-
examined. Setting C to zero, in a frame of reference that is denoted by MI,
results in the following expression for V% (after redefining K):

(40)

In a frame of reference M2 that moves with a constant velocity - C z with


respect to M lt the velocity is given by:
u' = rd) + (W + C)z (41)
As a result

u'-Vjt = C
This means that the stability properties of a flow under u'*Vx = C can be
inferred from calculations conducted in a different frame of reference, obtained
by a proper Galilean transformation, in which irV% = 0. Since the stability
properties, as observed in different inertial frames of reference, should be the
same, examining the stability conditions obtained by setting C to zero and as
seen from all inertial frames is equivalent to considering all possible C's.
We start now the stability analysis by calculating N. According to its
definition after Eq. (30), we have:

(42)

It is obvious from Eq. (42) that for shearless systems, i.e., q1 = 0, N is


constant along the streamlines, whereas it varies in a nonperiodic way along
them otherwise. It is also noticed, since K is in the r direction, that the term
K-N in Eq. (35) is constant along the streamlines for both sheared as well as
shearless flows.
a. Shearless flows 1
For such flows q = 0 and N (and hence INI 2 ) are constant along the
streamlines. As a result, eq. (35) is a second-order ordinary differential equation
BALLOONING IN FLUID DYNAMICS 325

with constant coefficients and is given by:


H2Y
plNI 2 ^f+2N-KGX = 0 (43)

The solution of eq. (43) is of the form

X~e^ T

where }i is the solution of


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

plNI2|i2 + 2N-KG = 0 (44)

It is obvious that the perturbations decay exponentially and, hence, the


flow is stable for ballooning modes, if and only if

2N-KG > 0 (45)


is satisfied on each streamline, i.e., for each r.
The criterion of Eq. (45) is a necessary condition for the stability of
shearless flows. Using Eq. (37), it is given by:

^ (rQv +WW) + ^ - ~ r£l2 > 0 (46)

where c2 = 3P/3p is the sound speed and v = (r2 Q)'/r the z component of the
vorticity.

b. Sheared flows2
In this case INI is given by:
INI 2 = a + bG + c92 (47)

where a, b, and c are constants that can be obtained from Eq. (42).
The stability of the flow is determined by the behavior at T -» °°. In this
limit the ballooning equation is given by:

(48)
o
where P is a positive constant and use was made of the ray equations

f=rft(r) (49)

which implies that 6 is proportional to i.


326 M. MONO AND E. HAMEIRI

The solution of Eq. (48) is of the form

where the two possible |i are given by:

M- =-—=——^——~— (50)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

The stability condition is Re|a < 0. This implies

N-KG > 0
which is the same as the condition of Eq. (46) for shearless flows.
The stability conditions for sheared as well as for shearless flows, which
turned out to be the same (the shape of the modes, however, differs in both
cases), were obtained for the particular ballooning modes for which irV^ = 0.
Hence, according to the discussion earlier in this section, the expression on the
left-hand side of the condition of Eq. (46) should be minimized with respect to
W (for a shifted axial velocity with a given W1) in order to obtain a stability
condition for all of the ballooning modes. The result is:

+ rft2 - >0 (51)

where £ is the magnitude of the vorticity and v is its z component. The


criterion of Eq. (51) was obtained and discussed by Eckhoff and Storesletten1.

V. Conclusion
The general equations for ballooning modes were derived. These equations
describe the propagation of two waves along their rays, which are the
streamlines of the basic flow. A stream function representation was used in
order to obtain a simple ordinary differential equation for the variation of the
amplitude of the ballooning mode along the streamlines.
As a result of that representation the role of the various physical
characteristics of the basic flow, such as streamline curvature, vorticity, etc.,
are clearly displayed in the resulting ballooning equation.
As an example, a particular fluid flow, in which the streamlines form a set
of nested cylindrical surfaces, was investigated. An explicit stability condition
was obtained, which is a necessary condition for the stability of the flow. The
latter agrees with the necessary condition obtained by Eckhoff and
Storesletten1.
Acknowledgment
This research was supported by the U.S. Department of Energy under grant
no. DE-FGO2-86ER53223.
BALLOONING IN FLUID DYNAMICS 327

References
ifickhoff, K.S. and Storesletten, L., "A Note on the Stability of Steady
Inviscid Helical Gas Flows," Journal of Fluid Mechanics , Vol. 88, 1978, pp.
401 -411.
2
Bayly, B.J., "Three-Dimensional Centrifugal-Type Instabilities in
Inviscid Two Dimensional Flows," Physics of Fluids , Vol. 31, 1988,
pp. 56 - 64.
3
Leibovich, S. and Stewartson, K., "A Sufficient Condition for the
Instability of Columnar Vortices," Journal of Fluid Mechanics , Vol. 126,
1983, pp. 335 - 356.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

4
Singh, P.I. and Uberoi, M.S., "Experiments on Vortex Stability,"
Physics
5
of Fluids , Vol. 19,1976, pp. 1858 - 1863.
Howard, L.N., "On the Stability of Compressible Swirling Flow,"
Studies of Applied Mathematics , Vol. 52, 1973, pp. 39 - 43.
6
Eckhoff, K.S., "On the Stability of Symmetric Hyperbolic Systems: II,"
Journal of Differential Equations , Vol. 43,1982, pp. 281 - 304.
7
Coppi, B., "Topology of Ballooning Modes," Physical Review Letters ,
Vol. 39, 1977, pp. 939 - 941.
8
Connor, J.W., Hastie, R.J. and Taylor, J.B., "High Mode Number
Stability of an Axisymmetric Toroidal Plasma," Proceedings of the Royal
Society of London , Series A: Mathematical and Physical Sciences, Vol. 365,
1979, pp. 1 - 17.
9
Dewar, R.L. and Glasser, A.H., "Ballooning Mode Spectrum in General
Toroidal
10
Systems," Physics of Fluids , Vol. 26, 1983, pp. 3038 - 3061.
Hameiri, E., "On the Essential Spectrum of Ideal
Magnetohydrodynamics," Communications on Pure and Applied Mathematics ,
Vol. 38, 1985, pp. 43 - 66.
H
Hameiri, E. and Chun, S-T., "Stability of Ballooning Modes in a
Rotating Plasma". Physical Review A: General Physics, Vol. 41, 1990, pp.
1186-1189.
12
Lakhina, G.S., Hameiri, E., and Mond, M., "Ballooning Instability of
the Earth's Plasma Sheet Region in the Presence of Parallel Flow," Journal of
Geophysical
13
Research , Vol. 95, No. A7,1990, pp. 10,4441 -10,448.
Hameiri, E., "Ballooning Modes in Fluid Dynamics," Proceedings of the
1985 Sherwood theory conference, Paper 2Q - 17, Madison, WI, April 15-17,
1995.
14
Jeffrey, A. and Taniuti, T., Non-Linear Wave Propagation , Academic
Press, New York, 1964.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Instabilities of the Nonuniform Flows of a


Low-Temperature Plasma in MHD Channels
I. M. Rutkevich*
Russian Academy of Sciences, Moscow, Russia

Abstract

Instabilities of stationary nonuniform magnetohydrodynamic


(MHD) flows of slightly ionized plasma with equilibrium ionization are
considered. The main attention is directed to the influence of
longitudinal nonuniformities of MHD flow on the eigenfunctions,
eigenfrequencies, and increments of acoustic oscillations. A modified
method of geometrical acoustics is used for solving the eigenvalue prob-
lem. The important special type of acoustic instability of nonuniform
subsonic flows with respect to two-dimensional perturbations is
considered. This instability arises in the inner acoustic resonator between
two critical (cutoff) cross sections of a channel. A spectral instability of
supersonic flows in MHD channels is discussed briefly.

I. Introduction

The earliest theoretical studies of instabilities in low-temperature


magnetohydrodynamic (MHD) plasma were made in the 1960s
(acoustic1'2 thermal, 3 and ionization, 4'5 instabilities). In the open-cycle
equilibrium magnetohydrodynamic MHD generators, an ionization
instability is absent but other types of instabilities can arise for some
MHD flows «

Copyright © 1992 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
*Principal Scientist, Institute of High Temperature.
328
INSTABILITIES OF NONUNIFORM FLOWS 329

Analysis of stability of stationary flows of plasma with equilibrium


ionization is usually based on the single-fluid MHD approximation

3p / 3t + divpv = 0

p[dv / 3t + (vV)v] = - Vp +1 x B + divi


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

pT(3S / 3t -f (vV)S) = j2 / a - divq - Qr + Qd (1)

Here p is the mass density v the flow velocity; p, T, and S the


pressure, temperature, and unit-mass entropy; j , B, and a the electric
current density, magnetic induction, and conductivity; T, q, Qr and Qd
the tensor of viscous stresses, heat flux vector, unit-volume radiation
losses, and viscous dissipation, respectively. The equations of
electrodynamics will be used in low-frequency approximation

rotE = 0, divj=0, j + j x p = o(E + VxB) (2)

is fulfilled. Here t* is the characteristic time, L the characteristic length


of plasma nonuniformities, and c = 1 / ^/£0(I0 the speed of light.
Investigation of stability is based on the linearized equations (1) and (2).
The small perturbations having the form

f' = F(r;co)exp(-icot) ^

are considered. Here F is the vector eigenfunction, and co = cor + icDj is


the complex eigenfrequency. A flow is unstable if a certain eigenvalue
has a positive imaginary part (coj>0).
There is no universal method for calculation of spectrum for
arbitrary quasi-one-dimensional flow. However, for high-frequency
short waves, the asymptotic method of the WKB type can be used. In the
early papers, !~3' 9> 10 a local dispersion equation (LDE) was applied to the
description of short waves. LDE is the relation of the following type

A(co,k) = 0 (4)

Equation (4) can be obtained for the approximate local solutions


~expi(kr-G)t). For some nonuniform flows, the local increment of
traveling wave co(r) determined by the solution of LDE can change its
sign along the channel. For the analysis of stability of such flows the LDE
330 I. M. RUTKEVICH

approximation is invalid and the more rigorous methods should be used.


In the present paper, the asymptotic method of the WKB type is applied
to analytical investigation of instabilities of the quasi-one-dimensional
flows in MHD channels.
The organization of this paper is as follows. In sec. II, the modified
method of geometrical acoustics is developed for nonuniform plasma
flows. In sec. Ill the solution for two- and three-dimensional waves
traveling in MHD channels are constructed. The cutoff condition for
traveling acoustic waves in channels are discussed. In sec. IV the
analytical expressions for acoustic spectrum of subsonic flows
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

(eigenfrequencies and increments) are obtained for multidimensional


perturbations in the absence of critical cross sections. The spectrum of
two-dimensional oscillations localized between two cutoff cross sections
and the sufficient conditions for cutoff acoustic instability are considered
in sec. V. In sec. VII a spectrum of perturbations and instabilities of
supersonic MHD flows are discussed briefly.

II. Method of Geometrical Acoustics for Nonuniform MHD Flows

Linearizing Eq. (1) and (2) close to the stationary solution, one can
obtain a system of equations in the tensor form:

+ A iJj k (r) + BJi j ( r J) f : + - c i j k < + Dik = 0 (5)


dt dxk dx k I dx^ I dx k

Here f ' ^ V ' x / V ' y / V ' ^ p ' / S ' ) is the column vector of gasdynamic
perturbations, and (p' is the perturbation of electric potential (E = -Vcp').
The gasdynamic subsystem [Eq. (5)] is of parabolic type, while Eq. (6) is
of elliptic type. Parabolicity of Eq. (5) is caused by the effects of viscosity
and heat conductivity. We will neglect small dissipative terms in the
flow core and assume Ci-k^=0. This assumption is justified if the
Reynolds number and the Peclet number (by wavelength are much
larger than unity. In the case CijW=0, the gasdynamic subsystem (5)
becomes of hyperbolic type. The elements Aijfc determine the velocities
of propagation for different types of disturbances along the
characteristics of simplified equations. The terms B|j describe all the
effects related to the influence of gradients of^the background state as
well as the electromagnetic force j ' x B and Joule heating
(j/cr)'perturbations without taking into account the electric field
perturbation E f . The terms Dik3cpf /3xk describe the contribution of the
INSTABILITIES OF NONUNIFORM FLOWS 331

perturbation F into the perturbations j ' x B and (j^/a)'. In Eq. (6) the
coefficients ajj are the components of the electrical conductivity tensor.
The high-frequency gasdynamic perturbations satisfying the
condition

1 [aB2 (Y-l)j 2 V + al , ,«
= — max^
max ——
——, ^—— ^-,—
- —— 4«1 (7)
Icol [ p op L J

are considered. Here y is the adiabatic index, V the velocity of main flow,
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

a the sound velocity, and L the characteristic length of nonuniformity of


the main flow. Propagation of high-frequency perturbations can be
calculated by the method of geometrical acoustics approximation.
Solutions of Eqs. (5) and (6) may be in the form:

[f',<?'} = {F(r),4)(r)}exp[ico{e(f)-t}] (8)

Here co9 is the eikonal and F(r),<t>(?) are the slowly varying
amplitudes which may be sought in the form of the asymptotic series

= F0(r)(-ico)-s (9a)

T)(-m)-s~1 (9b)
s=l

The quantity 0(r) satisfies the eikonal equation (11)

a 2 (r)(V9) 2 = [l-V(f)Ve] 2 (10)

for acoustic waves and

V(r)V9 = l

for entropy and vortex waves. Realization of the WKB method in the case
of a current-carrying plasma is more complex than in classical
geometrical acoustics due to the presence of electric perturbations. The
substitution of Eq. (9) in Eq. (6) gives the expression for the first term E0
in the WKB expansion of the amplitude of electric field perturbation

_ - 0
= «... n, n= / (11)
(an-n) —-
332 I. M. RUTKEVICH

The substitution of Eq. (9) into Eq. (5) with allowance for Eqs. (10)
and (11) leads to the transport equation for the principal part, PQ, of a
slowly varying amplitude of pressure perturbation in an acoustic wave

W(r;n)VP 0 - X(f ;n)P0 = 0 (12a)

W = V(r) + a(r)n(T) (12b)

Here W is the group velocity and n is the unit vector determined in


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Eq. (11). The quantity X is the local increment of traveling wave in the
nonuniform MHD plasma. The expressions for K contain the following
three part:

where

-{[l + (HP)2 ]op / a - (HP)2 pp p } -f (nj )B{pp - p[l - (HP)2 ]op / a}

-p p (jB)(np)J}

+2(PP / P - op / o)[(nj )(n( j x p» -f- (5p)(nj)(j P)] - 2PPP j2 (nb)2 -

-2p-'a-' o[n( j x p> + (5B)((5j )(5P) - j P)]}

Kv = - - [divW - WV In{pa2(l - VV9)} + p^a'1 {p( VV)V + Vp}n

+c;1vvs]

Here a p = (3a/3p)s and P p =(3p/3p) s are the isoentropic


derivatives of conductivity and Hall parameter with respect to the
pressure, b = B / B is the unit vector directed along the applied magnetic
field. The quantities Xp and XQ describe the influence of electromagnetic
force and Joule heating perturbations on the amplification and
attenuation of the sound wave. The term Xv describes the contributions
INSTABILITIES OF NONUNIFORM FLOWS 333

of stationary gradients into the acoustic increment. Expression (13) takes


into account all the known mechanisms of acoustic instability of current-
carrying plasma with equilibrium ionization. !'2'6. In the case
j = 0, B = 0 , the relations

= 0, X v =--{divW-WVln[pa 2 (l-VV9)]}

are fulfilled and Eq. (12) turns into the classical transfer equation (11).
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

III. Propagation of Traveling Acoustic Waves in MHD Channels

For stationary quasi-one-dimensional flows depending on the x


coordinate only the eikonal equation (10) admits the three-dimensional
solutions of the following type:

c o G = k * ( x ; k i / ( 0 2 ) d x + k y y + kzz, kx* = 0 ) ^ (14)

*„,
M(x)-> = -, r, o l-(a 2 -u 2 )k 2 /co 2

Here u(x) is the velocity of stationary flows, and M(x) the Mach
number. Expressions (14) correspond to the simple traveling waves of
the type of Eq. (8) in an unbounded medium. Using the combination of
these waves with different values of ky and k z/ one can obtain the
waveguide solutions describing the propagation of traveling waves in a
channel lykh y (x), lzkh z (x):
X

Pi =C nK Y m (y / x)Z s (z / x)exp(-icot+ j[ik*(x)Tr(x)]dx), CTO = const


(15)
Here Y m = cos(7cmy /2h y ) for even values of m,
Y m = sin(7ony /2h y ) for odd m, Zs = cos(;isz/2hz) for even s, and
Zs = sin(rcsz / 2h z ) for odd s. The functions Ym and Zs provide for the
absence of normal component of gas velocity of the walls of a channel.
The functions -T+(x) and -F_(x) are the coefficients of spatial
amplification of the waveguide modes traveling downstream and
upstream, respectively. The expressions for F± have the form

u + an* ' k* (16)


334 I. M. RUTKEVICH

where X(x;n) is the local increment determined in Eq. (13), kj and k^


are the following wave vectors:

k* = k* + k J, k,1 = (-I)'"1 k y e y + kzez (t = 1,2)


l
^=(-iy~ kyQy-kzez (* = 3,4), ky=7cs/2hy, k z =7cs/2h z

Solutions (15) are valid for the conditions

dh dh, ljJBh yy ^ Icolmmh h z )


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

maxl —'- —- , ——— ^ y .


dx dx pa2 lu±al
(18)

The first inequality in Eq. (18) is necessary for using the WKB
approximation. The second inequality in Eq. (18) enables us to construct
the waveguide solutions by the combination of solutions (8) describing
the waves in unbounded plasma with kj =-k2y and ksy=-k4y12 If the
third conditions is fulfilled, relation (11) can be used for the elementary
waves [Eq. (8)]. In the present case yk*|h y )»l the electrodynamic
boundary layers of the small thickness ~ I/ k~| are being formed at the
electrode walls of an MHD channel y = ±h .13. These boundary layers
make a small contribution to quantities F± *3'14 (see also Ref. 15).
In the case coi=0, solutions (15) describe the propagation of acoustic
waves if R(x;k 2 /co 2 )>0, where R is the function determined in Eq.
(14). For the supersonic flows, the function R is always positive and the
cutoff is impossible. For the subsonic flows, the domain with R<0 can
exist. The local cutoff frequency is determined by the formula

coc (x) = k ± (x)Va 2 (x)-u 2 (x)

The equality ooc(x)=oo can be fulfilled at the point x=x*. Then the critical
cross section is created at x=x*. The wave numbers of the upstream and
downstream waves should be coincided at the critical cross section.
Moreover, the longitudinal group velocities of both waves become zero:
W£(x,) = W~(x,) = 0. Therefore, in the region R>0, the sound rays
incident to the critical cross section will be turned backward. In the
region R<0, the amplitude of the wave will be attenuated exponentially
in space.
In the case cor »lco i l^0, a physical picture of the reflection of the
wave from the critical cross section is analogous to the case .coi=0 The
position of critical cross section can be found from

r = R(X;ki/o> r 2 ) = 0 (20)
INSTABILITIES OF NONUNIFORM FLOWS 335

Reference 16 shows that for acoustic waves in the region Rr>0, the
WKB solutions (15) are true. In the narrow neighborhood of the critical
cross section determined by the inequality

( v/3
t-ffVMx
(21)
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

the WKB formulas (15) are invalid and the acoustic field can be
expressed in terms of the Airy function Ai[(-ico)2/3f;(x)] and its
derivative. The influence of the region [Eq. (21)] on the acoustic waves in
the region R r >0 can be taken into account by the use of an effective
reflection coefficient K. If the region Rr>0 is situated at x>x*, then the
effective boundary condition is (P'+ + KP'_)X=X = 0 with K = ein/2. 16.

IV. Spectrum of Natural Acoustic Oscillations for a Subsonic Flow

For the construction of eigenf unctions in the region of flow xi<x<X2,


the following boundary conditions at the ends of the gasdynamic line are
used:

(P'.+K.P1. ),=,,=<), (P 1 .+K 2 P 1 + ) XXXi =0 (22)

In the general case, the reflection coefficients KI and K2 are complex


and depend on CD. It is assumed below, that for the open-cycle MHD
generator, the cross section x=xi corresponds to the nozzle entry, and
x=x2 to the diffuser exit. In the absence of critical cross sections, the
eigenfunctions satisfying conditions (22) are constructed as a
combination of the downstream and the upstream waveguide modes
(15).
The eigenfrequencies % are determined by

(23)

where n is an integer. The increments of oscillations co| are calculated by


the formula
-1/2
*f ( k2 ^ 2 ,2
co,1 = — [G x;-^- dx + lnlK.K,2 Ja' (l-M r R x ; -
2 2
dx
2 M co? ' ' (24)
336 I. M. RUTKEVICH

This formula is valid if the following inequality

J « cor min(l,<o2 ~ *) (25)

holds. The function G(x) is the sum of the coefficients of amplification for
the upstream and downstream waves. The spectrum {cor + 10^}
determined in Eqs. (23) and (24) is discrete and dependent on the three
integers n, m, and s. If the reflection of waves from the end cross
sections is nonideal (with losses of acoustic energy) then InJK^j) < 0. In
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

this case, the sufficient condition for the acoustic stability of the subsonic
flow is
*2

JG(x;k 2 /co 2 )dx<0 (26)

V. Localization of Natural Oscillations in a Nonuniform Flow

The formulas for the acoustic spectrum [Eqs. (23) and (24)] cannot be
used in the presence of the cutoff cross sections. They describe only the
part of the full acoustic spectrum when the quantity n / Jm 2 + s 2 h 2 / h 2
is sufficiently large. The two-dimensional perturbations in the plane
orthogonal to the magnetic field are considered below (s=0, Zs-l). If the
inequality

^n-1/2 l x f _! 2 ! i————— ^a2-u2


m 2 h
*i y (27)

holds, then the region of flow satisfying the condition Rr<0 exists.
Therefore, the eigenfunctions are localized in the region where Rr>0.
The most interesting situation is the case of double-sided localization
corresponding to the function \|/(x) with local minimum at the point
x=xoe(xi,xx). In this case, the eigenmodes withrj < j\ are localized in the
inner resonator between two critical cross sections x = x^ < x0 and
x = x^ > x 0 . For the spectrum of two-dimensional localized oscillations,
the notations co r =co nm and coi=^ nm are used subsequently. The
following law or similarity is true for the spectrum. For a given
stationary flow, the quantities £nm = conm / m, X^ and the coordinates
of the critical cross sections x^, x^ depend on the parameter ^ only.
Since V(x^ m ) = \(f(Xn m ) so the quantity x^ is a determined function on
x
im/ i-ev xiL = gUi™) • The dependence x^m(ii) is determined by
g(Xnra)
(28)
INSTABILITIES OF NONUNIFORM FLOWS 337

If the quantities x^Cn) are found, then the frequencies conm are
calculated by the formula conm/=m^nm/ where ^nmOl) *s the function
determined by

i J a'1 (1 - M2 )-' V^L - X|/(x)dx = TI (29)


Xnm(Tl)

The increments of localized oscillations Xnm are calculated in terms


Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

of given dependencies £nmCn) andxjucn) by the formula

1 X""TI °m 2 Tr2™^
2
= -7 jG(x;n)dx/ ja-'(l-M )-' 1-^f - dx (30)
'

For some types of nonuniformities, the frequencies of oscillations


and the length of the resonator Lnm = x^ - x^ can be calculated
analytically. Thus, for the dependence \|/(x) = \|/0 + 1 / 2\|/0(x - x 0 ) 2 under
the conditions \|/0 > 0, \v'0 > 0, and a / hj = a0 / h^ = const, the solutions
of Eqs. (28) and (29) are17

2Q, )

It should be noted that Eq. (29), which determines the


eigenfrequencies, can be written in the equivalent form

(32)

where K^ and K^ are determined in Eq. (14). For the fixed value of
K y = 7on/2h y/ Eq. (32) is analogous tot he quasi classical Bohr-
Sommerfeld quantization condition for a particle in a potential hold.18 In
the quantum-mechanical problem, the integrand in Eq. (32) is equal to
T ( K ^ - K ~ ) = p / ^ , where p(x) is the momentum. The cutoff cross
sections x = x^ are analogous to the turning points of the quantum
wave function. However, the quantum-mechanical analogy is not
complete because in the acoustic problem K~ * -K* and the eigenvalues
are not real (coi * 0).
338 I. M. RUTKEVICH

a)
V
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 1 a) Typical dependence \|f(x) = (a2 - u2) / h 2 for the case of


double-sided localization of acoustic perturbations, b) Schematic v e w
of eigenfunction for localized perturbation.

The examples of two-dimensional cutoff instability in a subsonic


series MHD generator with varying cross sections are given in Ref 16.
For ti = (n--)/m<0.5, the values 0)i~50-60 s"1, cor/m~3-103 s"1 and
L < 1 were otrtained.
A cutoff instability can arise not only in MHD channels but also in
subsonic plasmatrons, and in subsonic flows of a nonconducting gas
with heat generation and heat losses. The following two conditions are
sufficient for the onset of the two-dimensional cutoff instability.
1) The function \|/ = (a2 - u 2 )h* has a local minimum at the point
X=XQ (see Fig. 1). Specifically, for a*/hy=const, the distribution of the
Mach number M(x) ha a local maximum at X=XQ.
2) The inequality G(x;T]=0) >0 holds at X=XQ, as is shown in Fig. 2
(curve 1). In this case, the localized modes with sufficiently small values
of T] will grow. If G(XQ;TI=O) <0 (see curve 2 in Fig. 2), then modes with
sufficiently small values of TJ will be damped. In this situation, the
growing modes can occur for the values of t| bounded from below. It
will take place, if the integral of G(X,TI) between the limits
is positive.
INSTABILITIES OF NONUNIFORM FLOWS 339
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Fig. 2 Possible behavior of the function G(x,T) = 0) 1, instability


arises for small values of T\; 2, instability is absent for small TJ .

The cutoff instability of the nonuniform subsonic flows is important


because the conditions for its occurrence do no depend on the boundary
conditions at the ends of a gasdynamic line. Investigations of the
nonlinear stage of the development of cutoff instability are in progress.17

VI. Instabilities of Supersonic Flows

In contrast to classical gasdynamics, for the supersonic plasma flow


in channels a spectrum of eigenperturbations can exist. In the supersonic
case, a spectrum arises exclusively due to a nonlocal dependence of the
electrical field perturbation on the gasdynamic perturbations. Therefore,
the WKB expression (11) cannot be used for the construction of
eigenfunctions, and the effect of an external electric circuit should be
taken into account.
A discrete spectrum of decaying eigenmodes for a supersonic flow
in a Faraday MHD channel with an open circuit is considered in Ref. 19.
It can be shown 20 that, for an arbitrary quasi-one-dimensional
supersonic plasma flow in a channel of finite length, the total number of
growing modes is always finite. The examples of the growing acoustic
and thermal eigenmodes in a supersonic plasmatron with longitudinal
current are given in Ref 20. Consideration of thermal instability in a
supersonic Faraday MHD generator is made in Ref. 21. In the
experiment 8 with a supersonic series MHD generator using the
equilibrium plasma, an excitation of acoustic oscillations has been
observed.

References

, E. P., "Hall Instability of Current-Carrying Slightly Ionized Plasma/'


First International Symposium on MHD Electrical Power Generation, 1962.
340 I. M. RUTKEVICH

^McCune, J. E., "Wave Growth and Instability in Partially Ionized Gases/'


Second International Symposium on MHD Electrical Power Generation, Vol.2, 1964,
pp. 533-538.
^Wright, J. K., "A temperature instability in the magnetohydrodynamic flow/'
Proceedings of the Physical Society Vol. 81, No. 521,1963, pp. 498-505.
4
Velikhov, E. P. and Dykhne, A. M., "Plasma Turbulence Due to the lonization
Instability in a Strong Magnetic Field," International Conference on Phenomena of
Ionized Gases Vo. 2,1965, p.47.
^Kerrebrock, J. L., "Nonequilibrium lonization Due to Electron Heating.. I:
Theory," AIAA Journal, Vol. 2, No. 6,1964, pp. 1072-1080.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

"Nedospasov, A. V. and Kjait, V. D., Oscillations and Instabilitie of a Low-


Temperature Plasma, Nauka, Moscow, 1979 (in Russian).
'Rutkevich, I. M. and Sinkevich, O. A., "Waves and Instabilities in a Low-
Temperature Plasma," Progress in Science and Technology Series on Mechanics of
fluids and Gases VINITI, Moscow, Vol. 14,1981, pp. 127-204 (in Russian).
8
Glinov, A. P., Golovin, A. P., Dreizin, Yu. A., Birykin, S. P., and Shipuk, I. I,
"Alternating Current Excitation in MHD Channel Load," Ninth International
Conference on MHD Electrical Power Generation, Vol. HI, 1986, pp. 1151-1158.
^Locke, E. V. and McCune, J. E., "Growth Rates for Axial Magneto-acoustic
Waves in a Hall generator" AIAA Journal, Vol. 4, No. 10,1968, pp. 1748-1751.
10
Powers, W. L. and Dicks, J. B., "Transient Wave Growth in
Magnetogasdynamic Generators," AIAA Journal, Vol. 6, No. 6, 1968, pp. 1007-
1012.
11
Blokhintsev, D. L, Acoustics of a Nonuniform Moving Medium/' Nauka, Moscow,
1981 (in Russian).
12
Rutkevich, I. M. and Tokar, P. M., "Two-Dimensional Acoustic Oscillations in a
Nonuniform MHD Flow: Travelling Waves," High Temperature, 1984.
13
Rutkevich, I. M., "Formation of Electrodynamic Boundary Layers During
Propagation of Sound in a Confined Low-Temperature Plasma," Soviet Physics -
Doklady, Vol. 27, No. 12,1982, pp. 1035-1036.
14
Rutkevich, I. M., "Stability of a Nonuniform Subsonic Flow in a High-Power
MHD Generator Against Quasi-One-Dimensional Oscillations," High
Temperature, Vol. 24, No. 2,1984, pp. 265-277.
15
Walker, J. S., "Fluctuations in Plasma MHD Generators," Seventh International
Conference on MHD Electrical Power Generation. 1980, pp. 681-684.
l^Rutkevich, I. M. and Tokar, P. M., "Effect of the Cut-off of Two-Dimensional
Acoustic Waves on the Onset of Instability of the Subsonic MHD Flows,"
Izvestiya Academii Nauk, USSR, Mechanics of Fluids and Gases, 1980, No. 2, pp. 26-36
(in Russian).
17
Komov, V. I. and Rutkovich, I. M., "Excitation of the Nonlinear Acoustic
Oscillations in a Moving Nonuniform Plasma: Equations for the Amplitudes of
Interacting Waves," Submitted to High Temperature, (1989).
18
Landay, L. D. and Lifshitz, B. M. Quantum Mechanics, Nauka, Moscow, 1974.
INSTABILITIES OF NONUNIFORM FLOWS 341

^Rutkevich, I. M., "Appearance of Free Acoustic Oscillations in supersonic Flow


of Low-Temperature Plasma in a Magnetic Field/' Fifteenth International
Conference on Phenomena of Ionized Gases, Minsk, 1981, Vol. I, pp. 333-334.
20Rutkevich, I. M., "Spectrum of Gas-Dynamic Perturbations and Stability of a
Supersonic Flow in the Presence of Joule Heating/' Fluid Dynamics, Vol. 17, No.
6,1982, pp. 940-947.
21
Glinov, A. P. and Dreizin, Yu. A., "Analysis of Thermal Instability in MHD-
Systems with Back Connection," Teplofiz. Vys. Temperatur, Vol. 24, No. 5,1986,
pp. 974-979.
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227
Author Index

Branover, H........................ 165, 190 Panidis, Th. ................................254


Brown, S. H. ..............................165 Papailiou, D. D...........................254
Cambon, C. ................................ 131 Paschereit, C. O. ..........................53
Downloaded by Indian Institute of Technology on September 19, 2012 | http://arc.aiaa.org | DOI: 10.2514/4.866227

Cosnuau, A.................................295 Petrunin, A. A. ........................... 116


Dang Tran, K. ............................295 Politano, H. ..................................81
Fiedler, H. E.................................40 Pouquet, A. ..................................81
Galanti, B................................... 146 Roy, S......................................... 151
Galperin, B................................. 151 Rutkevich,!. M. .........................328
Gilbert, A. D. ............................146 Ruzmaikin, A. A. .......................309
Goldshtik, M. A. ..........................87 Ryan,J. ......................................295
Hameiri, E..................................317 Sergeev, K.................................. 103
Henoch, C. ................................. 190 Shcherbinin, E. V.......................210
Hoffert, M. ................................. 190 Shtern, V. N. ................87, 103, 116
Hussain, F. ..................................... 1 Sondergaard, N. A. .................... 165
Kolpakov, N. Ju. ..........................65 Staroselsky, I...................... 151, 159
Kraichnan, R. H. .......................... 17 Sukoriansky, S. ..151, 159, 165, 190
Lee, J..........................................268 Sulem,P.L...........................81, 146
Lummer, M. .................................40 Talmage, G................................. 165
MaronMoalem, D......................221 Tanny, J......................................235
Melander, M. V.............................. 1 Virk,D. ..........................................1
Mond, M. ...................................317 Vlasyuk, V. N. ...........................210
Morchoisne, Y. ..........................295 Walker, J. S................................ 165
Naot, D...............................221, 235 Wygnanski, I. J. ...........................53
Nottmeyer, K. ..............................40 Yacoub, N. .................................221
Orszag, S. A............................... 151 Zimin, V. D..................................65

342

You might also like