Dynamique Vortex
Dynamique Vortex
Editorial Board
R. Beig, Wien, Austria
J. Ehlers, Potsdam, Germany
U. Frisch, Nice, France
K. Hepp, Zürich, Switzerland
W. Hillebrandt, Garching, Germany
D. Imboden, Zürich, Switzerland
R. L. Jaffe, Cambridge, MA, USA
R. Kippenhahn, Göttingen, Germany
R. Lipowsky, Golm, Germany
H. v. Löhneysen, Karlsruhe, Germany
I. Ojima, Kyoto, Japan
H. A. Weidenmüller, Heidelberg, Germany
J. Wess, München, Germany
J. Zittartz, Köln, Germany
3
Berlin
Heidelberg
New York
Barcelona
Hong Kong
London
Milan
Paris
Singapore
Tokyo
Editorial Policy
The series Lecture Notes in Physics (LNP), founded in 1969, reports new developments in
physics research and teaching -- quickly, informally but with a high quality. Manuscripts
to be considered for publication are topical volumes consisting of a limited number of
contributions, carefully edited and closely related to each other. Each contribution should
contain at least partly original and previously unpublished material, be written in a clear,
pedagogical style and aimed at a broader readership, especially graduate students and
nonspecialist researchers wishing to familiarize themselves with the topic concerned. For
this reason, traditional proceedings cannot be considered for this series though volumes
to appear in this series are often based on material presented at conferences, workshops
and schools (in exceptional cases the original papers and/or those not included in the
printed book may be added on an accompanying CD ROM, together with the abstracts
of posters and other material suitable for publication, e.g. large tables, colour pictures,
program codes, etc.).
Acceptance
A project can only be accepted tentatively for publication, by both the editorial board and the
publisher, following thorough examination of the material submitted. The book proposal
sent to the publisher should consist at least of a preliminary table of contents outlining the
structure of the book together with abstracts of all contributions to be included.
Final acceptance is issued by the series editor in charge, in consultation with the publisher,
only after receiving the complete manuscript. Final acceptance, possibly requiring minor
corrections, usually follows the tentative acceptance unless the final manuscript differs
significantly from expectations (project outline). In particular, the series editors are entitled
to reject individual contributions if they do not meet the high quality standards of this
series. The final manuscript must be camera-ready, and should include both an informative
introduction and a sufficiently detailed subject index.
Contractual Aspects
Publication in LNP is free of charge. There is no formal contract, no royalties are paid,
and no bulk orders are required, although special discounts are offered in this case. The
volume editors receive jointly 30 free copies for their personal use and are entitled, as are the
contributing authors, to purchase Springer books at a reduced rate. The publisher secures
the copyright for each volume. As a rule, no reprints of individual contributions can be
supplied.
Manuscript Submission
The manuscript in its final and approved version must be submitted in camera-ready form.
The corresponding electronic source files are also required for the production process, in
particular the online version. Technical assistance in compiling the final manuscript can be
provided by the publisher’s production editor(s), especially with regard to the publisher’s
own Latex macro package which has been specially designed for this series.
Quantized
Vortex Dynamics
and
Superfluid Turbulence
13
Editors
C.F. Barenghi
University of Newcastle
Mathematics Department
Newcastle NE1 7RU, United Kingdom
R.J. Donnelly
University of Oregon
Physics Department
Eugene, OR 97403, USA
W.F. Vinen
University of Birmingham
Physics Department
Birmingham B15 2TT, United Kingdom
This work is subject to copyright. All rights are reserved, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, reuse of illustra-
tions, recitation, broadcasting, reproduction on microfilm or in any other way, and
storage in data banks. Duplication of this publication or parts thereof is permitted only
under the provisions of the German Copyright Law of September 9, 1965, in its current
version, and permission for use must always be obtained from Springer-Verlag. Violations
are liable for prosecution under the German Copyright Law. Springer-Verlag Berlin Hei-
delberg New York
a member of BertelsmannSpringer Science+Business Media GmbH [Link]
c Springer-Verlag Berlin Heidelberg 2001
Printed in Germany The use of general descriptive names, registered names, trademarks,
etc. in this publication does not imply, even in the absence of a specific statement, that such
names are exempt from the relevant protective laws and regulations and therefore free for
general use.
Typesetting: Data conversion by Steingraeber GmbH, Heidelberg
Cover design: design & production, Heidelberg
Printed on acid-free paper
SPIN: 10792065 55/3141/du - 5 4 3 2 1 0
Preface
This book springs from the programme Quantized Vortex Dynamics and Super-
fluid Turbulence held at the Isaac Newton Institute for Mathematical Sciences
(University of Cambridge) in August 2000. What motivated the programme was
the recognition that two recent developments have moved the study of quan-
tized vorticity, traditionally carried out within the low-temperature physics and
condensed-matter physics communities, into a new era.
The first development is the increasing contact with classical fluid dynamics
and its ideas and methods. For example, some current experiments with he-
lium II now deal with very classical issues, such as the measurement of velocity
spectra and turbulence decay rates. The evidence from these experiments and
many others is that superfluid turbulence and classical turbulence share many
features. The challenge is now to explain these similarities and explore the time
scales and length scales over which they hold true. The observed classical aspects
have also attracted attention to the role played by the flow of the normal fluid,
which was somewhat neglected in the past because of the lack of direct flow
visualization. Increased computing power is also making it possible to study the
coupled motion of superfluid vortices and normal fluids. Another contact with
classical physics arises through the interest in the study of superfluid vortex re-
connections. Reconnections have been studied for some time in the contexts of
classical fluid dynamics and magneto-hydrodynamics (MHD), and it is useful to
learn from the experience acquired in other fields.
The second development arises from atomic physics and is the discovery of
Bose–Einstein condensation in confined clouds of alkali atoms. The study of
superfluidity and quantized vorticity is now possible in a wide range of other
systems besides helium II. The rapid progress in this area has given momentum
to the use of the Gross–Pitaevskii Equation or Nonlinear Schroedinger Equation
(NLSE). Researchers have become more aware of the approximations and limi-
tations involved in the NLSE model, but also of its range of validity and great
power of prediction. The use of the NLSE has become more established, and
the NLSE is proving to be a powerful tool for modeling problems such as vortex
nucleation, reconnections and even turbulence.
A further development arises from the results of preliminary theory and ex-
periments in turbulent Helium 3 which suggest that there are significant differ-
ences with turbulence in Helium 4 and these are likely to be explored in the
future.
VI
It is apparent from this background that the contributions to this book come
from investigators with a wide range of backgrounds and expertise: condensed-
matter physics and low-temperature physics, classical fluid dynamics and applied
mathematics, MHD, atomic physics, and engineering (for the applications of
helium II as a cryogenic coolant).
The book is divided into topical chapters. Each chapter begins with one or
two introductory review articles, which are suitable for students and new inves-
tigators interested in entering the field. The introductory articles are followed
by shorter, more specialized papers.
Chapter 1 introduces us to the problem of quantized vorticity and super-
fluid turbulence, and it summarizes the key aspects and problems which are
currently studied. Chapter 2 is devoted to turbulence experiments. Chapter 3
considers the fundamental problem of friction and vortex dynamics. The theory
of superfluid turbulence and the interpretation of the experimental results is the
subject of Chap. 4. Chapter 5 is devoted to the application of the NLSE model
to superfluidity and vortices. Chapter 6 moves away from helium and considers
Bose–Einstein Condensation and vortices in the context of alkali atoms. Chap-
ter 7 is concerned with some aspects of classical turbulence and MHD which
are relevant in the study of superfluid turbulence. Finally, Chap. 8 deals with
Helium 3 and other systems.
We are grateful for the support and encouragement of Professor Keith Mof-
fatt, Director of the Newton Institute, and we would like to thank Tracey Andrew
who helped in the preparation of the manuscripts for publication.
Part I Introduction
5.3
Challenges for Understanding Periodic
Boundary Layer Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5.4 Instrumentation to Detect Vortices Below 1 K . . . . . . . . . . . . . . . . . 33
5.5 The Normal Fluid and the Vortex Tangle . . . . . . . . . . . . . . . . . . . . . 33
5.6 Flow over Blunt Objects, Testing Models such as Submarines . . . 34
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Grid-Generated He II Turbulence
in a Finite Channel – Experiment
J.J. Niemela, L. Skrbek, S.R. Stalp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Intermittent Switching Between Turbulent and Potential Flow
Around a Sphere in He II at mK Temperatures
M. Niemetz, H. Kerscher, W. Schoepe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
1 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
2.1 Stable Turbulent Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
2.2 Intermittent Switching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
2.3 Turbulent Phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2.4 Laminar Phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Transition to Dissipation
in Two- and Three-Dimensional Superflows
C. Huepe, C. Nore, M.-E. Brachet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
2 Definition of the System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
3 Numerical Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
4 Bifurcation Diagram and Scaling in 2D . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
5 Subcriticality and Vortex-Stretching in 3D . . . . . . . . . . . . . . . . . . . . . . . . 301
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Current-Sheet Formation
near a Hyperbolic Magnetic Neutral Line
B.K. Shivamoggi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
2 Current-Sheet Formation at a Hyperbolic Magnetic Neutral Line
in a Stagnation-Point Plasma Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
3 Effect of a Uniform Shear–Strain in the Plasma Flow . . . . . . . . . . . . . . . 385
4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
Nonlocality in Turbulence
A. Tsinober . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
1 Introduction and Simple Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
2 Different Aspects of Nonlocality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
2.1 Direct Coupling Between Large and Small Scales . . . . . . . . . . . . . . 392
3 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
Contents XVII
C.F. Barenghi
University of Newcastle L. Eaves
Mathematics Department School of Physics and Astronomy
Newcastle upon Tyne, NE1 7RU University of Nottingham
UK Nottingham NG7 2RD, UK
[Link]@[Link] [Link]@[Link]
N. Berloff
Mathematics Department A. Fetter
University of California Geballe laboratory for Advanced
Los Angeles, CA 90095-1555, USA Materials
nberloff@[Link] Stanford University
Stanford, CA 94305-4045, USA
R. Blossey fetter@[Link]
Department of Physics
University of Essen
45117 Essen, Germany K.L. Henderson
blossey@[Link] Faculty of Computer Studies and
Mathematics
A. Brandenburg University of the West of England
NORDITA Bristol BS16 1QY, UK
Blegdamsvej 17 [Link]@[Link]
2100 Copenaghen, Denmark
brandenb@[Link]
D.D. Holm
H.P. Buechler Theoretical Division
Theoretical Physics Mail Stop B284
ETH Los Alamos National Laboratory
8093 Zuerich, Switzerland Los Alamos NM 87545, USA
buechler@[Link] holm@[Link]
XX List of Contributors
G. Hornig M. Leadbeater
Dept. Theoretical Physics IV Physics Department
Ruhr-Universitaet Bochum University of Durham
44780 Bochum, Germany Durham DH1 3LE, UK
gh@[Link] [Link]@[Link]
C. Huepe T. Lipniacki
James Frank Institute Institute of Fundamental Technologi-
University of Chicago cal Research
5640 S. Ellis Avenue Świȩtokrzyska St. 21
Chicago, IL 60637, USA 00-049 Warsaw, Poland
cristian@[Link] tlipnia@[Link]
D. Kivotides
S. Nazarenko
Mathematics Department
Mathematics Institute
University of Newcastle
University of Warwick
Newcastle NE1 7RU, UK
Coventry CV4 7AL, UK
[Link]@[Link]
snazar@[Link]
J. Koplik
Levich Institute, T-1M S.K. Nemirovskii
City College of New York Institute of Thermosphysics
New York, NY 10031, USA 630090 Novosibirsk, Russia
koplik@[Link] nem@[Link]
koplik@[Link]
J.J. Niemela
M. Krusius Physics Department
Low Temperature Laboratory University of Oregon
Helsinki University of Technology Eugene, OR 97403, USA
02015 HUT joe@[Link]
Finland
krusius@[Link] M. Niemetz
Institut für Experimentelle
H. Kuratsuji und Angewandte Physik
Department of Physics Universität Regensburg
Ritsumeikan University 93040 Regensburg, Germany
Kusatsu City 525-8577, Japan [Link]@
kra@[Link] [Link]
List of Contributors XXI
Carlo F. Barenghi
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 3–14, 2001.
c Springer-Verlag Berlin Heidelberg 2001
4 C.F. Barenghi
where C is a circular path around the axis of the vortex. The ratio Γ = h/m of
Plank’s constant and helium’s mass is called the quantum of circulation and has
value Γ = 9.97 × 10−4 cm2 /sec.
The simplest way to create superfluid vortex lines is to rotate a cylinder
filled with Helium II at constant angular velocity Ω. Provided that Ω is large
enough, superfluid vortex line appear and form on ordered array of areal density
n = 2Ω/Γ , all vortex lines being aligned along the axis of rotation. In this way
the superfluid mimics the vorticity 2Ω of the solid body rotating normal fluid,
each vortex line contributing one quantum to the total circulation.
Superfluid Vortices and Turbulence 5
∂ψ 2 2
i =− ∇ ψ − mEψ + V0 ψ|ψ|2 , (8)
∂t 2m
Here E is the energy per unit mass and = h/2π. The wavefunction ψ can be
written as ψ = AeiΦ in terms of an amplitude A and a phase Φ. In this way
one can define the condensate’s density ρBEC = mA2 and velocity vBEC =
(/m)∇Φ; the last relation confirms that the superfluid is irrotational as en-
visaged by Landau. This transformation establishes the hydrodynamics of the
model: equation (8) is equivalent to a continuity equation and an Euler equation
(modified by the so called quantum pressure term).
The NLSE has a vortex solution: if Φ is the azimuthal angle φ then we have
vBEC,φ = Γ/(2πr) which is the Onsager - Feynman vortex. Substitution into
the NLSE yields a differential equation for ρBEC . One finds that ρBEC tends
to the bulk value m2 E/V0 for r → ∞, and that ρBEC → 0 for r → 0. The
characteristic distance over which ρBEC changes from its bulk value to zero is
a0 ≈ 10−8 cm. This distance is called the vortex core parameter. We conclude
that the superfluid vortex line is hollow at the core. Geometrically, a vortex
line transforms the volume occupied by the superfluid into a multiply connected
region.
Hereafter we identify ρBEC with ρs at absolute zero and vBEC with vs . It
must be noted that this identification is convenient from the point of view of
having a simple hydrodynamics model but is not entirely correct. The reason is
that Helium II is a dense fluid, not the weakly interacting Bose gas described
by the NLSE, so the condensate is not the same as the superfluid component.
One should compare the case of Helium II (in which the NLSE is a rather
approximate model) with the case of BEC in clouds of trapped alkali atoms (in
which the NLSE is a better model because the bosons’ interaction is weaker).
Another drawback of the NLSE is that it fails to describe the observed dispersion
relation of Helium II at high momenta. If one studies small oscillations of the
uniform solution of the NLSE and interprets them as thermal excitations, one
finds a dispersion relation E = E(p) in which the energy E is proportional
to the momentum p at small p (phonons) and then becomes quadratic in p at
high p (free particles). The spectrum of excitations observed in Helium II is
different, because the phonon part is followed by the rotons’ minimum. Despite
these shortcomings, of which one must be aware, the NLSE is much used as a
convenient hydrodynamical model of Helium II at T = 0.
6 C.F. Barenghi
vs,tot = vs + vi , (3)
The self induced velocity vi describes the motion which a vortex line induces
onto itself because of its own curvature and is determined by the Biot - Savart
(BS) law
Γ (z − s) × dz
vi (s) = , (4)
4π |z − s|3
The Biot - Savart law is sometimes replaced by the Local Induction Approxima-
tion (LIA), which is
vi ≈ βs × s , (5)
where β = Γ/(4π)log(1/(|s |a0 )). Since |s | = 1/R where R is the local radius
of curvature, we have
Γ R
vi ≈ ln( )b̂, (6)
4π a0
where b̂ is the binormal.
The drag force fD arises from the mutual friction between the superfluid
vortex lines and the normal fluid [4]. Normal fluid flowing with velocity vn past
Superfluid Vortices and Turbulence 7
a vortex core exerts a frictional force fD per unit length on the superfluid in the
neighborhood of the core given by
The dimensionless parameters α and α are temperature dependent and are of-
ten written in terms of mutual friction coefficients B and B defined by α =
ρn B/(2ρ) and α = ρn B /(2ρ). The values of B and B are known from experi-
ments. Samuels and Donnelly [5] showed that, at least in the high temperature
range in which most experiments are performed, the friction arises from the
scattering of rotons from the velocity field of a vortex line. The calculation of
the mutual friction parameters over the entire temperature range is still an open
question and has subtle aspects, as explained in the article by Sonin. An impor-
tant effect of the mutual friction is that it modifies the propagation of second
sound. By measuring the second sound attenuation one can determine the su-
perfluid vortex line density L0 , defined as the length of vortex line per unit
volume.
Now that fM and fD are identified we can make use of the fact that the sum
of all forces is zero as the line’s inertia is negligible:
fD + fM = 0, (8)
ds
= vs + vi + αs × (vn − vs − vi ) + α (vn − vs − vi ), (9)
dt
An algorithm to numerically simulate the time evolution of any arbitrary
configuration of vortex lines can be developed on the basis of Schwarz’s equa-
tion and is described in the article of Samuels. Here it suffices to say that an
initial vortex configuration is discretized into N points. The time evolution of
each point is calculated using (9), given externally applied fields vs and vn and
given the temperature T , which determines the friction coefficients α and α .
The transverse part of the mutual friction, proportional to α , is smaller and is
sometimes neglected. The number of points N and the time step must be allowed
to vary during the evolution to take into account the appearance of regions of
high or low curvature. Note that, if one uses the BS law, the computational time
is proportional to N 2 , while, if one uses the LIA, this time is only proportional to
N . Numerical simulations of vortex tangles based on the BS are therefore compu-
tationally expensive. However the use of the LIA can give misleading results [6].
Finally the numerical simulation must be able to perform vortex reconnections
when two vortex lines come sufficiently close to each others. This process is an
arbitrary assumption in the context of the dynamics of vortex filaments, but is
justified by a microscopic calculation [7] performed using the NLSE, as explained
in the article by Koplik.
8 C.F. Barenghi
Besides the NLSE model (in which the vortex core is visible) and the vortex
dynamics model (in which the core is not visible but the vortex line is) there
is a third macroscopic model in which the individual vortex lines are not visi-
ble and Helium II is considered as a continuous vortex flow. The third model,
called the HVBK model[8] [9], is useful to describe laminar flows in which the
vortex line are spatially organized. Examples are solid body rotation, flows in
an rotating annulus or cavity, and Taylor - Couette flow. The HVBK model is
a generalization of Landau’s equations to include the presence of vortices. The
fluid particles of the model are assumed to be large enough to be threaded by
many vortex lines which are aligned in the same direction. In this way the indi-
vidual vortex lines are not visible, the superfluid is treated as a continuum and
we can define a macroscopic, nonzero superfluid vorticity ωs , despite the fact
that, microscopically, the superfluid velocity field obeys ∇ × vs = 0. Clearly the
HVBK equations are valid only if the vortex lines are organized spatially and not
randomly oriented, and if the length scales of the flow under consideration are
much bigger than the average separation between the vortex lines. An example
is the simple case of Helium II inside a rotating cylinder, for which ωs = 2Ωẑ.
The incompressible HVBK equations are
∂vn 1 ρs ρs
+ (vn · ∇)vn = − ∇P − S∇T + νn ∇2 vn + F, (10)
∂t ρ ρn ρ
∂vs 1 ρn
+ (vs · ∇)vs = − ∇P + S∇T + T − F, (11)
∂t ρ ρ
where we have defined
ωs = ∇ × vs , (12)
B B
F= ω
s × [ωs × (vn − vs − νs ∇ × ω
s )] + ωs × (vn − vs − νs ∇ × ω
s ), (13)
2 2
ω
s = ωs /|ωs |, (14)
T = −νs ωs × (∇ × ω
s ), (15)
νs = (Γ/4π) log(b0 /a0 ), (16)
The quantities F, T and νs are respectively the friction force, the tension
force and the vortex tension parameter, and b0 = (2ωs /Γ )−1/2 is the intervor-
tex spacing. Note that νs has the same dimension of a kinematic viscosity, but
physically it is very different: it is related to the ability of a superfluid fluid
particle to oscillate because of the vortex waves which can be excited along the
vortex lines threading the fluid particle itself. Note that without F and T the
HVBK equations are formally the same as the original two - fluid equations of
Landau. The HVBK equations have interesting limits. If T → Tλ then ρs → 0 so
the normal fluid equation (10) becomes the classical Navier - Stokes equation. If
T → 0 then ρn → 0 so the superfluid equation (11) describes a pure superflow;
Superfluid Vortices and Turbulence 9
by setting Plank’s constant equal to zero we have then νs = 0 and the pure
superflow equation becomes the classical Euler equation.
The HVBK model has been used with success to study the transition from
Couette flow to Taylor vortex flow [10]: Barenghi’s predictions [11] of the critical
Reynolds number of the transition and its temperature dependence were con-
firmed by the experiments [12]. Taylor - Couette flow has also been studied in
the nonlinear Taylor vortex flow regime [13] [14] and the results are in agreement
with the observations, providing a further test of the theory. These results are
described in the article by Henderson.
4 Turbulence
Superfluid turbulence manifests itself as a tangle of vortex lines and can be
generated in many ways. Turbulent thermal counterflow was the first turbulent
flow which was studied in detail in a series of pioneering papers by Vinen [15].
Since this flow has no classical analogy it deserves a separate discussion. Other
ways to generate turbulence are more classical in character, and we refer to them
as turbulent coflows.
together, the general trend is that the slow, laminar flow of Helium II, with or
without vortices, tends to be rather different from the flow of a classical fluid,
but when Helium II moves fast and is driven turbulent it seems to behave like a
classical turbulent flow.
To characterize the turbulence we use the Reynolds number Re = U L/ν
where L is the length scale, U the velocity scale and ν the kinematic viscosity.
Examples of the observed classical features of Helium II turbulence are the fol-
lowing. Mass flow rates and pressure drops at Re ≈ 106 can be well described
by using classical relations for high Reynolds number classical flows [18]. Ex-
periments on large scale turbulent vortex rings at Re ≈ 4 × 104 detect normal
fluid vorticity and superfluid vorticity moving together as a single structure [19]
Experiments on turbulent Taylor - Couette flow at Re ≈ 4×103 show the typical
structures of classical turbulent Taylor - Couette flow [20]. Experiments on the
decay of superfluid vorticity created by towing a grid show that the decay in
time obeys the same laws as of the decay of classical turbulence [21]. Moreover
the decay appears to be independent of temperatures in the explored range,
from the lambda region down to 1.4 K, where the normal fluid fraction is only
7.5 percent. Experiments on turbulence created by rotating blades [22] show the
classical Kolmogorov −5/3 power spectrum in the temperature range explored,
from the lambda region down to T = 1.4 K again. Finally experiments on the
drag on a moving sphere (Re ≈ 105 ) show the same drag crisis observed in a
classical fluid [23].
The temperature independence of these observations is interesting. The nor-
mal fluid must be responsible for these classical aspects, but the dynamical im-
portance of the normal fluid should be related to the fraction ρn /ρ, so it should
be negligible at temperatures as low as 1.4 K. Since a large number of quantized
vortex lines must be present in these turbulent flow, it is speculated that they
are able to lock together the two fluid components of Helium II into a single
fluid which behaves somewhat like a classical turbulent fluid. This is a topic of
much current interest‘[24] [25] and is discussed in the article of Vinen.
tion and never changed, neglecting the back reaction of the superfluid vortices.
Not surprising, the vortex tangles calculatedusing different driving fields looked
different from each others. The success of the original calculation of Schwarz in
reproducing a vortex line density L0 consistent with the experiments in the T-2
state was probably due to the fact that the uniform vn − vs profile used by
Schwarz modelled well the average, flattened turbulence profile in the channel
flows under consideration.
Despite this limitation, kinematic calculations are clearly useful and can shed
light on important physical mechanisms. A particularly interesting mechanism
which is relevant to turbulence is the Ostermeier- Glaberson instability. This
is an instability of Kelvin vortex waves which takes place if the component of
the normal fluid velocity in the direction parallel to the vortex lines exceeds a
critical value. The instability was first observed by Cheng, Cromar and Don-
nelly [30], but is was Ostermeier and Glaberson [31] who explained it and it
was Samuels [28] who realized its importance in turbulence. Another physical
mechanism, which is apparent in the kinematic numerical calculations [28] [29]
is vorticity matching: once the Ostermeier - Glaberson instability has generated
superfluid vortex lines by extracting energy from the normal fluid, then the vor-
tex lines become attracted to the regions of concentrated normal fluid vorticity.
Therefore, although the local superfluid velocity pattern in the bundles is very
complicated, the averaged vorticity ωs is similar to the vorticity of the driving
normal fluid. A further interesting application of the Ostermeier - Glaberson
instability is that it creates a damping length scale for superfluid turbulence:
superfluid structures at length scale smaller than will lose energy to the normal
fluid and be dissipated [25].
measurements of the transition from the weak T-1 state (in which the superfluid
is turbulent but the normal fluid is not) to the strong T-2 state (in which both
fluids are turbulent).
The same kinematic approach has been used to study how the superfluid
vortices affect the stability of normal fluid in channel flows. This modified Orr -
Sommerfeld problem[32] is described in the article by Godfrey.
ds
= vs + h1 vsi + h2 s × (vn − vi ) − h3 s × s × vn , (22)
dt
where vi is given by (4). In the absence of friction we have h2 = h3 = 0 and
h1 = 1. The normal fluid is determined by a modified Navier - Stokes equation
like (20), but now F is obtained numerically by considering the friction force on
the normal fluid per unit length of superfluid vortex line
where D and Dt are mutual friction coefficients, and summing the contribution
of each segment of s that falls within the computational grid cell of the normal
fluid.
The drawback of the fully - coupled approch is the computational cost. A
useful compromize is to implement it in two dimensions x, y neglecting the z
dependence: in this way the vortex lines becomes vortex points (the intersection
of vortex lines with the plane z = 0) and the normal fluid can be obtained
easily using the stream function - vorticity formulation. It is found that a single
superfluid vortex point induces an elongates normal fluid jet [34] whose intensity
depends on the temperature, hence creating a dipolar vorticity structure in the
normal fluid. A similar calculation in three dimensions performed by Kivotides,
Barenghi and Samuels [35] showed that a superfluid vortex ring creates around
itself a normal fluid structure which consists of two coaxial vortex vortex rings
of opposite polarity.
Since the normal fluid is accelerated by the superfluid vortices by friction
which depends on the difference between vn and vL , the normal fluid speed
cannot exceed the vortex lines’ speed. If one considers the typical vortex line
Superfluid Vortices and Turbulence 13
8 Discussion
References
1. L.D. Landau and E.M. Lifshitz: Fluid Mechanics, 2nd edn. (Pergamon Press, Lon-
don 1987)
2. R.J. Donnelly:Quantized Vortices in Helium II, Cambridge University Press, Cam-
bridge (1991).
3. K.W. Schwarz: Phys. Rev. Lett. 49, 283 (1982); Phys. Rev. B 31 5782 (1985); Phys.
Rev. B 38 2398 (1988)
4. C.F. Barenghi, R,J, Donnelly and W.F. Vinen: J. Low Temp. Phys. 52 189 (1983)
5. D.C. Samuels and R.J. Donnelly: Phys. Rev. Lett. 65 187 (1990).
6. R.L. Ricca, D.C. Samuels and C.F. Barenghi: J. Fluid Mech. 391 29 (1999)
7. J. Koplik and H. Levine: Phys. Rev. Lett. 71, 1375 (1993).
8. H.E. Hall and W.F. Vinen: Proc. Roy. Soc. London A 238, 215 (1956).
9. R.N. Hills and P.H. Roberts: Arch. Rat. Mech. Anal. 66, 43 (1977).
10. C.F. Barenghi and C.A. Jones: J. Fluid Mech. 197 551 (1988)
11. C.F. Barenghi: Phys. Rev. B 45, 2290 (1992)
12. C.J. Swanson and R.J. Donnelly: Phys. Rev. Lett. 67 1578 (1991)
13. K.L. Henderson, C.F. Barenghi and C.A. Jones: J. Fluid Mech. 283 329 (1995).
14. K.L. Henderson and C.F. Barenghi: Phys. Lett. A 191 438 (1994).
15. W.F. Vinen: Proc. Roy. Soc. A 240, 114 (1957); ibidem 240, 128 (1957); ibidem
242, 493 (1957); ibidem 243 400 (1957).
16. J.T. Tough: ‘Superfluid turbulence’. In: Progress in Low Temperature Physics, vol.
VIII, ed. by D.F. Brewer (North Holland, Amsterdam 1987) p. 133.
17. D.J. Melotte and C.F. Barenghi: Phys. Rev. Lett. 80 4181 (1998).
14 C.F. Barenghi
18. P.L. Walstrom, J.G. Weisend, J.R. Maddocks and S.V. VanSciver: Cryogenics 28
101 (1988).
19. H. Borner, T. Schmeling and D.W. Schmidt: Phys. Fluids 26 1410 (1983)
20. F. Bielert and G. Stamm: Cryogenics 33 938 (1993)
21. M.R. Smith, R.J. Donnelly, N. Goldenfeld and W.F. Vinen: Phys. Rev. Lett. 71
2583 (1993)
22. J. Maurer and P. Tabeling: Europhysics Lett. 43 29 (1998)
23. M.R. Smith, D.K. Hilton and S. V. VanSciver: Phys. Fluids. 11 751 (1999)
24. W.F. Vinen: Phys. Rev. B 61 1410 (2000)
25. D.C. Samuels and D. Kivotides: Phys. Rev. Lett. 83 5306 (1999)
26. R.G.K.M. Aarts and A.T.A.M. deWaele: Phys. Rev. B 50 10069 (1994)
27. D.C. Samuels: Phys. Rev. B 46 11714 (1992)
28. D.C. Samuels: Phys. Rev. B 47 1107 (1993)
29. C.F. Barenghi, G. Bauer, D.C. Samuels and R.J. Donnelly: Phys. Fluids 9 2631
(1997)
30. D.K. Cheng, M.W. Cromar and R.J. Donnelly: Phys. Rev. Lett. 31 433 (1973)
31. R.M. Ostermeier and W.I. Glaberson: J. Low Temp. Phys. 21 191 (1975).
32. S.P. Godfrey, C.F. Barenghi and D.C. Samuels: Physica B 284 66 (2000); Phys.
Fluids 13 983 (2001).
33. O.C. Idowu, D. Kivotides, C.F. Barenghi and D.C. Samuels: J. Low Temp. Physics,
120 269 (2000).
34. O.C. Idowu, A. Willis, D.C. Samuels and C.F. Barenghi: Phys. Rev. B 62 3409
(2000)
35. D. Kivotides, C.F. Barenghi and D.C. Samuels: Science 290 777 (2000).
36. C.F. Barenghi, D. Kivotides, O. Idowu and D.C. Samuels: J. Low Temp. Physics,
121 377 (2000).
37. M. Leadbeater, T. Winiecki, D.C. Samuels, C.F. barenghi and C.S. Adams: Phys.
Rev. Lett. 86 1410 (2001).
38. D. Kivotides, C. Vassilicos, C.F. Barenghi and D.C. Samuels: Phys. Rev. Lett. 86
3080 (2001).
39. D. Kivotides, C.F. Barenghi and D.C. Samuels: to be published in Europhys. Lett.
(2001).
An Introduction to Experiments
on Superfluid Turbulence
Russell J. Donnelly
1 Introduction
A description of the experimental background of superfluid turbulence was as-
signed to me for this lecture. Superfluid turbulence, or as some call it, quantum
turbulence, has been an active field of physics since the 1950’s. The field was
pioneered experimentally and theoretically by Joe Vinen and as such is approach-
ing a half century in age. It is safe to say that with few exceptions the results
are unknown to those investigators who are interested in classical turbulence,
that is the kind of investigation which has been pioneered by Taylor, Landau,
Kolmogorov and others.
It has only recently been realized that liquid helium I, liquid helium II and
cryogenic (critical) helium gas are attractive candidates for investigating classical
turbulence problems, and in the process many have decided to look at the kinds
of challenges encountered in using helium II, that phase of liquid helium which
exhibits superfluidity.
In preparing this talk I had hoped to cover, however briefly, the entire corpus
of experimental work on the subject. It soon became evident that there was
too much material by far than could be covered in a single one-hour lecture.
Fortunately, I had recently written a review article in honor of Joe Vinen’s
retirement from the University of Birmingham [1] and Skrbek, Niemela and
myself had written a second paper for the same occasion [2]. These papers contain
virtually all the known results on cryogenic fluid mechanics. This talk will instead
concentrate mostly on future directions, and results obtained since the articles
cited above were written.
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 17–35, 2001.
c Springer-Verlag Berlin Heidelberg 2001
18 R.J. Donnelly
Push Rod
Bellows 10 cm
Pump LHe
~
~
P
{
Smooth 25 cm
Pipe
Fig. 1. Bellows driven flow apparatus for measuring the friction factor in a smooth pipe.
The advantage of cryogenics can be appreciated by comparing the size of this apparatus
with the famous apparatus of Nikuradze built in the early 1930’s and weighing many
tons. The highest Reynolds numbers achieved are about the same in both Nikuradze’s
and our apparatus.
unit time. Flow velocity is limited by the power of the drive motor. A sensitive
capacitance manometer has been developed to measure the pressure gradient
in all working fluids and is located 38 diameters down stream of the inlet. The
in-situ differential pressure gauge was specifically designed to avoid mechanical
strains at low temperature, which can overwhelm the 0.05 Pascal resolution of
the device [3].
This experiment, primarily the effort of Chris Swanson of the University of
Oregon, with help from Gary Ihas, visiting from the University of Florida, has
revealed several subtleties in such measurements, including the need for high
purity of helium, fine stability of temperature and pressure, and care to maintain
a time-independent state. Agreement with data at room temperature using air
and water as a fluid is quite good.
The primary advantage of this apparatus is that it accommodates any non-
reactive fluid at a temperature between 1 K and 300 K, allowing unprecedented
ranges of Reynolds numbers to be obtained. For example, we have used room
temperature SF6 , air, and helium gas, as well as cryogenic helium I at 4.2 K
to measure the friction factor from Reynolds number 10 to 5,000,000, nearly
six decades of Reynolds numbers. The fact that the results agree with standard
engineering data from many sources attests to the utility of this new approach.
An Introduction to Experiments on Superfluid Turbulence 19
Motor Control
and Counter
16 bit ADC
200MHz Pentium
GPIB, Labview
G
P
I
B
Function Vacuum Tight Sliding Seal
B Generator
U Vacuum
Pump
S
Lock-in
Amplifier
Amplifier
60 mF Grid Generated
Turbulence
Grid
Second Sound
Bias Voltage Receiver
100 Volts DC
Second Sound
Transmitter
Second Sound
LR-110 Standing Wave
Resistance
Bridge
Germanium
Thermometer 1 cm x 1 cm x 29 cm
Channel
Heater
LR-130
Temperature
Controller
Fig. 2. Schematic diagram of the University of Oregon towed grid apparatus in helium
II. Turbulence is generated by sweeping the grid upward, and is observed by measur-
ing the attenuation of second sound in the presence of the turbulence which can be
interpreted to give the vorticity in the superfluid.
ε = ν ω 2 = ν κ2 L2 (1)
where ω 2 is the mean square vorticity in the coupled fluids, and the identifi-
cation of ω 2 with κ2 L2 is discussed in [9]. Therefore according to the model
second sound measures ω 2 , as a function of time. After an initial time interval
the turbulent energy spectrum has the classical Kolmogorov form
-4
2 .0 x 1 0
-4
1 .5 x 1 0
/s )
2
ν ' (c m
-4
1 .0 x 1 0
ν (T )
-5
5 .0 x 1 0
1 .0 0 1 .2 5 1 .5 0 1 .7 5 2 .0 0 2 .2 5
T e m p e ra tu re (K )
decay rate increases. The length of line is obtained from second sound resonance
measurements using the relation
16Δ0 A0
L(t) ≈ ( − 1) (4)
κB A(t)
where Δ0 is the full-width at half maximum for the second-sound (power) res-
onance curve in the absence of vortex lines (at the 50th harmonic), B is the
mutual friction constant obtained from measurements on uniformly rotating he-
lium [10], and A(t) and A0 are respectively the peak second sound amplitude
with and without vortex
lines present. Using (4) and (5) we see that the ex-
periment measures C 3 /ν . Note that Maurer and Tabeling [11] performed an
experiment with rotating disks in liquid helium in which they were able to mea-
sure the Kolmogorov spectrum by means of pressure fluctuations using a probe
inserted at a selected location in the flow and connected to a quartz pressure
transducer. Their results showed that the Kolmogorov spectrum could be ob-
served in both helium I and helium II (2.3K, 2.0K and 1.4K) and reached the
important conclusion that the Kolmogorov constant C is the same above and
below the lambda transition. With that information, we see that our experiment
determines ν providing we know C. The Kolmogorov constant is taken to be
1.5 at all temperatures, which is the accepted classical value [12]. The resulting
values of ν are shown as a function of temperature in Fig. 3. The error bars
An Introduction to Experiments on Superfluid Turbulence 23
Fig. 4. The decaying helium II vorticity measured at T=1.3 K for the indicated ReM .
helium II vorticity in the temperature range 1.2K < T < 2K. The four regimes
switch successively as the energy containing and dissipative Kolmogorov length
scales gradually grow during the decay, finally both being saturated by the size of
the channel. In Fig.4 each curve represents an average of three individual decays.
As the decay curves tend to collapse on the universal curve, we shifted them for
clarity by a factor of two downwards, the uppermost remaining unchanged. The
early part of the vorticity decay displays a power law with exponent -11/10 (see
left inset, showing normalized data forReM = 104 ) and later −5/6 (see right
inset, showing normalized data forReM = 1.5 × 105 , 2.5 × 104 and 5 × 103 ). After
saturation, typically several orders of magnitude of decaying vorticity closely
follow the power law with exponent −3/2, represented by the thick solid line.
For the first time we report the fourth and last regime - a late exponential
decay - not shown separately in Fig.4. The nature of this last decay is not
understood at present.
This experiment explores over 8 orders of magnitude of decaying turbulent
energy, an impossible task for a wind tunnel, which would need to have a 1000
km test section to observe the same thing.
So far we have not had a continuous flow facility to work on. About two years ago
we decided to build a small wind tunnel using critical helium gas as the working
fluid. The advantage of critical helium gas is the enormous range of properties
which can be reached by adjusting the pressure and temperature. Flow velocities
available depend on the density. At low densities flow velocities up to 1m/s are
possible, at high densities velocities to 30cm/s are available. One of the optimal
operating points for high Re flow will be at 4 bar pressure and 6 K at which
the kinematic viscosity is 3.1 × 10−8 m2 /s. Mesh Reynolds numbers of 30,000 to
100,000 will be available, and corresponding microscale Reynolds numbers will
range from 150 to 280.
We have recently successfully operated this tunnel at 6.5 K and mesh Reynolds
numbers around 1500. The grid generated turbulence is probed with 10 micron
diameter cryogenic hot wire anemometer also developed in our laboratory. These
hot wires are observed to obey King’s law relating velocity and voltage, and pre-
liminary velocity time series show standard statistical features. Note that this is
the first cryogenic tunnel to operate below liquid nitrogen temperature.
We have given considerable attention to the conceptual design of larger wind tun-
nels designed to be useful for model testing. These have been discussed recently
in [2] and do not need repeating in this article. The importance for superfluid
turbulence of these devices is the possibility of operating them in helium II.
An Introduction to Experiments on Superfluid Turbulence 25
Liquid helium offers much promise for tow tank design, a subject also reported
in some detail in [2]. Again the importance for superfluid turbulence is the pos-
sibility of operating with helium II as the working fluid.
26 R.J. Donnelly
• Hot wires, Laser Doppler Velocimeters (LDV) and Particle Image Velocime-
ters (PIV) should all work in principle in helium gas, helium I and helium
II.
• RMS vorticity can be measured in helium II by second sound attention in
both open flows and counterflows. There is some speculation that chemical
potential probes would give local information on vorticity fluctuations, but
this has yet to be implemented [17].
• Temperature gradients can be measured by means of standard germanium
thermometry and pressure gradients, for example, by means of capacitance
manometers [3].
• Lift and drag on models can best be measured using Magnetic Suspension
and Balance Systems (MSBS), discussed in [18] by Britcher. Of course in a
cryogenic environment the magnets can be superconducting.
• Wall stress gages can be fabricated for work in helium gas and helium I.
• Ion trapping can be used to measure vorticity in helium II [25].
Hot wire anemometers are resistive self-heating devices which balance heat lost
to flow, which depends on the fluid velocity. Calibration provides the correlation
between fluid velocity and electrical power supplied. Standard hot wires have
d ∼ 5 microns and L ∼ 1000 microns. Standard materials (e.g. platinum) used at
room temperature are insensitive at low temperature. Our cryogenic hot wires are
made on a quartz fiber of diameter 10 microns. They consist of an evaporated Au-
Ge film of thickness of order 3000Å. A small sensitive region in middle of the fiber
is defined by masking the fiber and evaporating a metallic film over it as shown
in Fig.7. Masking is achieved by laying a small diameter fiber perpendicular to
the first fiber. The metal film provides electrical contact to the sensitive region.
Figure 8 shows how the small sensors are mounted. Sensor support dimen-
sions follow the “rule of 10”: they are placed ten times (more or less) their
characteristic dimension away from the sensitive region. Electrical contact to
the fiber is made through stainless steel wires, which are isolated from the brass
ring by epoxy. The fiber is epoxied to the wires using an electrically conducting
epoxy.
We have recently tested a few prototype 10 micron sensors in the wind tunnel
of Fig.5. A helium gas flow at 7 K was generated and we applied a relatively large
sinusoidal voltage to overheat one of the sensors and measured the temperature
28 R.J. Donnelly
Fig. 7. Principle of the cryogenic hot wire anemometer developed at the University of
Oregon.
of a nearby sensor. The characteristic heat signature at twice the input frequency
was clearly seen on the monitoring sensor. Furthermore, we found that the signal
to noise ratio to be the same as that predicted by our noise analysis. Thus we
feel confident that the sensors are working according to our expectations.
An Introduction to Experiments on Superfluid Turbulence 29
The first critical velocity. What accounts for the first critical velocity in
counterflow and its temperature dependence? This phenomenon is likely a case
of extrinsic nucleation, but lacks any quantitative explanation.
30 R.J. Donnelly
Fig. 9. An illustration of the dramatic differences in decay of two turbulent flows of the
same initial vorticity produced by a counterflow (upper curve) and towed grid (lower
curve).
The Tough classification. Why do large and small aspect geometries in coun-
terflow exhibit such different behavior?
Combined rotation and heat flow. Combined rotation and heat flow is a
relatively new area of investigation. Prior to the investigation discussed in this
An Introduction to Experiments on Superfluid Turbulence 31
section it was assumed (from earlier experiments) that the ordered array of vor-
tex lines produced by steady rotation and the disordered tangle produced by
counterflow preserved their identities in a combined experiment. Measurements
at Oregon with improved sensitivity by Barenghi, Swanson and Donnelly [20]
showed that the picture just described is far from true. The observations con-
sisted of measuring the amount of vortex line present owing to counterflow or
rotation alone using second sound attenuation, and comparing the observed line
density with what would be expected if the two sources of vorticity simply added.
The results are complicated, but appear to be relatively simple in two limits:
(i) Limit of large line densityLH due to heat, slow rotation. Here the effect of
rotation is not simply to add line densityLR = 2Ω/κ. Instead the tangle appears
to be polarized to accomplish the rotation. The effective polarization increases
with rotation Ω by analogy to a gas of magnetic dipoles in a magnetic field.
The results scale with LR /LH by analogy to μH/kT . Thus rotation appears to
produce alignment in the tangle, as does a magnetic field for dipoles, and LH
appears to play the role of disordering heat bath in the statistical mechanics of
superfluid turbulence. Indeed, it takes very little polarization of a dense tangle
to produce rotation at the relatively small angular velocities of the apparatus.
(ii) Limit of fast rotation and small axial heat flux. Any rotation eliminates
the critical velocity vc . In this limit two critical counterflow velocities appear,vc1
1
and vc2 , which scale as Ω 2 .
We might speculate here that the first critical velocity appears to correspond
to the Donnelly-Glaberson instability, excitation of helical waves by the coun-
terflow on the vortex lines induced by rotation [1]. The second appears to be
a transition to turbulence, with the rotation-induced array becoming a vortex
tangle. A more formal investigation of these effects is likely to be rewarding.
Fig. 10. The variation with amplitude of the damping of gravity oscillations of liquid
helium in a U-tube at a period of 0.94 sec. After Donnelly and Penrose [22] and Donnelly
and Hollis Hallett [21].
Fig. 11. Apparatus constructed by Davis, Hendry and McClintock [14] to measure the
decay of superfluid turbulence at 70 mK.
the self-consistent interaction between the normal fluid and quantized vortices
will likely lead to the need to develop measurement techniques which will give
more information than just the RMS line density L.
Acknowledgements
First of all I am grateful to Keith Moffatt, Director of the Newton Institute, for
the opportunity to have this remarkable gathering here. I am indebted to Joe Vi-
nen, Steve Stalp, Ladislav Skrbek, Carlo Barenghi, David Samuels, Renzo Ricca
and Peter McClintock for many useful discussions. My research is supported by
the National Science Foundation under grant DMR-9529609.
References
1. R. J. Donnelly, J. Phys Condensed Matter, 11, 7783-7834 (1999).
2. L. Skrbek, J. J. Niemela and R. J. Donnelly, J. Phys Condensed Matter, 11, 7761-
7782 (1999).
3. Chris J. Swanson, Kris Johnson and Russell J. Donnelly, Cryogenics 38, 673-677
(1998).
4. P. L. Walstrom, J. G. Weisend II, J. R. Maddocks and S. W. Van Sciver, Cryogenics
28, 101 (1988).
5. M.R. Smith, R. J. Donnelly, N. Goldenfeld and W.F. Vinen, Physical Review
Letters 71, 2583 (1993).
6. M.R. Smith “Evolution and Propagation of Turbulence in Helium II”. PhD Dis-
sertation, University of Oregon (1992).
7. S. Stalp, “Decay of Grid Turbulence in Superfluid Helium”. PhD Dissertation,
Physics, University of Oregon (1998).
8. S.R. Stalp, L. Skrbek, R.J. Donnelly, Phys. Rev. Lett., 82, 4831 (1999)
9. W. F. Vinen, Phys. Rev. B61, 1410 (2000)
10. W. F. Vinen “An Introduction to the Theory of Superfluid Turbulence” Paper in
this volume
11. J. Maurer and P. Tabeling, Europhys. Lett., 43, 29 (1998).
12. K.R. Sreenivasan, Phys. Fluids, 7, 2778 (1995).
13. C.F. Barenghi, R.J. Donnelly, W.F. Vinen, J. Low Temp. Phys., 52, 189 (1983).
14. S. I. Davis, P. C. Hendry and P. V. E. McClintock, Physica B , 280, 43 (2000).
15. L. Skrbek, J. J. Niemela and R. J. Donnelly, Phys. Rev. Lett., 85, 2973 (2000).
16. H. Schlichting, Boundary Layer Theory, 7th ed, McGraw-Hill (1979).
An Introduction to Experiments on Superfluid Turbulence 35
17. C. F. Barenghi C. E., Swanson, and R. J. Donnelly, J. Low Temp. Physics, 100,
385 (1995).
18. R. J. Donnelly (Editor), Liquid and Gaseous Helium as Test FluidsSpringer-
Verlag, 1991). In High Reynolds Number Flows Using Liquid and Gaseous He-
lium(Springer-Verlag, 1991). Also R. J. Donnelly and K. R. Sreenivasan (Editors)
Flow at Ultra-High Reynolds and Rayleigh Numbers(Springer-Verlag, 1998)
19. K. W. Schwarz, and J. R. Rozen, Phys. Rev. B , 44, 7563 (1991).
20. C. F. Barenghi C. E., Swanson, and R. J. Donnelly, Phys. Rev. Lett. , 50, 190
(1983).
21. R. J. Donnelly and A. C. Hollis Hallett, Annals of Physics, 3,320 (1958).
22. R. J. Donnelly and O. Penrose, Phys. Rev. , 103, 1137 (1956).
23. R. M. Ostermeier and W. I. Glaberson, Phys. Lett. 49A, 223 (1974).
24. D. Samuels and C. Barenghi, Phys. Rev. Lett. 81, 4381,1998.
25. R. J. Donnelly,Quantized Vortices in Helium II,(Cambridge University Press,
1991).
26. S. W. Van Sciver, “The Drag Crisis on a Sphere in Flowing HeII” Paper in this
volume.
The Experimental Evidence for Vortex
Nucleation in 4He
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 36–50, 2001.
c Springer-Verlag Berlin Heidelberg 2001
The Experimental Evidence for Vortex Nucleation 37
Fig. 1. Critical velocity, normalised to the zero temperature linear extrapolation value
v0 , versus T , in Kelvin: () [5], () [1], for ultra-pure 4 He. The curves are computed
from the half-ring model for a0 = 2.2, 3.2, 4.5, 6.0 Å and are normalised to match the
experimental value at 0.5 K. The inset shows the influence of 3 He impurities on vc , (◦)
3 ppb, () 45 ppb.
Eq.(4) stems from an asymptotic evaluation of the integral at the saddle point
t = 0. The accuracy of the asymptotic evaluation (4) becomes questionable as
T → 0 where the energy barrier vanishes. It has been checked by direct numerical
integration for typical cases and found to be quite satisfactory.
The critical velocity vc is defined as the velocity for which p = 1/2. This
definition is independent of the experimental setup, except for the occurrence in
(4) of the natural frequency of the Helmholtz resonator ω. The implicit equation
for vc reads:
ω0 −2πkB T Ea (P, T, vc )
exp − = ln 2 . (5)
2πω vc ∂Ea / ∂v|vc kB T
23 1
2
3 kB T
vc = vc0 1− γ , (6)
2 EJ
Ea = Ee (1 − v/vc0 ) , (7)
Fig. 2. Vortex nucleation rate and phase slip probability in terms of the mean velocity
in the micro-aperture expressed as phase winding numbers [16] in 4 He at 12 mK, 0.6
bar, with 100 ppb of 3 He impurities: frame (a) top, time spent, in seconds (, left
scale), and number of phase slips (◦, right scale) per velocity bin of size 0.1; frame
(a) bottom, slip rate in s−1 as a function of flow velocity, in units of 2π; frame (b),
cumulative probability obtained from the histogram of the slip velocities. The plain
curve is a least square fit to the data with the functional form corresponding to (4).
The value p = 1/2, shown by the arrow, defines the critical velocity (here, vc = 59.68).
At low temperatures and large critical velocities, the quantity in curly brack-
ets in the right hand side of (8) is small with respect to the last term so that the
−1
width is simply expressed as Δvc = −(2/ln 2) kB T ∂Ea /∂v|vc . Thus, the
statistical width is an approximate measure of the inverse of the slope of Ea in
terms of v.
This quantity is derived from p, itself obtained by integrating the histogram
of the number of nucleation events ordered in velocity bins. This procedure is
illustrated in Fig. 2: p shows an asymmetric-S shape characteristic of the double
exponential dependence of p on v, (4), a consequence of Arrhenius’ law, (3),
being plugged into a Poisson probability distribution. The observation of this
asymmetric-S probability distribution constitutes another experimental clue for
the existence of a nucleation process.
The critical transition of vc displays a measurable width, shown in Fig. 3,
which implies (as does the T -dependence of vc ) that the energy barrier is neither
very high compared to kB T nor very steep in terms of v. Fig. 3 contains data from
several groups. These data are somewhat scattered, especially above 0.5 K, but
they do show overall agreement within this scatter. They provide experimental
input on ∂Ea /∂v|vc and put tight limits on the theoretical models discussed
below.
It is also possible to measure the nucleation rate Γ directly as a function
of v. This quantity is the ratio, for a given velocity bin, of the number of slips
which have occurred at that velocity to the total time spent by the system at
the same velocity. The procedure is illustrated in Fig. 2. The slope of Γ (v) yields
∂Ea /∂v|vc ; the value of Γ at vc gives a combination of ω0 and Ea (vc ). Non-linear
40 É. Varoquaux et al.
Fig. 3. Statistical width of the critical velocity transition, normalised to the linear
extrapolation limit at T = 0, v0 , in terms of temperature: () [5], () [1], (×) [17] .
fits of p to (4) give estimates of ω0 , EJ and vc /vc0 which are independent of the
specific features of a given model. These estimates compare well with those of
the more precise analysis described below.
The 1/2 –ring model. Described in full in [5], this model has a long history,
following the work of Volovik [22], itself based on the theory of homogeneous
nucleation of vortices by Iordanski, and by Langer and Fischer [23]. It has been
very successful in accounting for vortex nucleation by ions [24]. In essence, the
model is based on the hydrodynamics of an Eulerian fluid assumed valid to scales
of the order of the vortex core diameter a0 , that is down to atomic size in 4 He.
Nucleation is assumed to occur on an asperity on the walls of the micro-aperture
where i) the flow velocity is largest, ii) the superfluid density is depressed. The
perturbed volume over which nucleation takes place is, on heuristic grounds,
of the order of a30 . The radius of the nucleated half-ring when it escapes to
the bulk fluid turns out of the order of 15 Å. As a consequence, the size and
shape of the asperity can be neglected in a first approximation; the asperity
only serves the purpose of breaking translational invariance. The bases of the
1
/2 –ring model have been re-examined critically by Sonin [25] and by Fischer
[26]. It can also be mentioned at this point that Nore et al. [27] have shown by
numerical simulations of 3D flows past an obstacle in the Gross-Pitaevskii model
that vortex filaments do evolve spontaneously into one, or possibly a few, half-
rings which thus constitute the preferred configuration of small vortices close to
walls.
The 1/2 –ring model is amenable to an expansion of the form (2) but the full
expression of Ea can straightforwardly be evaluated numerically and substituted
into the expressions of vc and Δvc , (5) and (8) [5]. The energy barrier impeding
vortex nucleation is found to vanish at
κ
vc0 = 0.432 . (10)
2πa0
This value, obtained for an Eulerian fluid, is to be compared with that obtained
by Rica for the Gross–Pitaevskii model, vc0 = 0.262κ/2πa0 [28], comparison
which points up quantitative inadequacies in the 1/2 –ring model outlined above.
42 É. Varoquaux et al.
Discussion of the 1/2 –ring model. There have been several attempts to
reformulate the 1/2 –ring model, both to put it on firmer theoretical grounds and
to possibly obtain a better description of the experimental results [1,3,25,26,34].
The 1/2 –ring model main inadequacies lie at the high velocity (small vortex)
The Experimental Evidence for Vortex Nucleation 43
end. Burkhart et al. [3] have introduced corrections to the vortex energy and
momentum due to the proximity of a boundary as computed with the Gross-
Pitaevskii equation. These corrections go in the right direction but cannot be
considered as a full reformulation of the small-vortex-at-a-wall problem.
A reassessment of the model, due to Sonin [25] and aiming in particular at
a quasiclassical reformulation of vortex quantum tunnelling, yields the following
elegant analytical form for the energy barrier
2
κ3 ρs κ
Ea = ln
32 π v 4π va0
and the 1/2 –ring model can be said to give a semi-quantitative description of
vortex nucleation when single phase slips are involved.
The two types of large slips. Examples of multiple slips can be seen in Fig. 4
which shows the peak amplitude chart of a two-aperture resonator at 12.5 mK,
24 bars, in a 100 ppb 3 He in 4 He sample. The very large amplitude drop shown
in Fig. 4 and in the insert is rare (one in 104 to 105 slips) under the conditions of
this particular experiment. This type of events, called in [38] ‘singular’ collapses
and discussed further below, may occur at velocities much below the vortex
nucleation threshold (down to at least a third of vc ).
Besides the usual single slip pattern, there appears in Fig. 4 occasional double
slips (i.e. involving phase changes by 4π) and infrequent triple slips. Raising the
temperature to 80 mK, again for this particular cooldown, causes these multiple
slips to occur much more frequently and to involve more circulation quanta
on the mean. Lowering the pressure to 0 bar results in an almost complete
disappearance of multiple slips at all temperatures. These features are described
in detail in [38].
Some degree of understanding of the formation of multiple slips can be gained
by plotting the mean value of the phase slip sizes, expressed in number of quanta,
against the flow velocity at which the slips take place [41]. This flow velocity is
close to the critical velocity for single phase slips, i.e. the nucleation velocity; it is
varied by changing the temperature,
the pressure, the resonator drive level. Such
a plot is shown in Fig. 5 for n+ , i.e. in flow direction conventionally chosen as
the + direction. Slips in the opposite (−) direction behave qualitatively in the
same manner but the phenomenon displays a clear quantitative asymmetry. As
The Experimental Evidence for Vortex Nucleation 45
can be seen in Fig. 5, the mean slip size decreases on either side of the quantum
plateau, as does the nucleation velocity, but increases with pressure, contrarily
to the nucleation velocity which decreases with increasing pressure. It appears
clearly that the magnitude of the superflow velocity does not directly control,
by itself, the occurrence of multiple slips. This implies, as will be discussed fur-
ther below, that the phenomenon under study is not purely hydrodynamical in
the bulk of the fluid but involves some complex interplay with the boundaries.
As shown in Fig. 5, the velocity threshold for the appearance of multiple slips
depends
on hydrostatic pressure; in fact, the P -dependence of the upturn of
n+ vs v exactly tracks that of the critical velocity for single phase slip nucle-
ation. This indicates that multiple slips appear because of an alteration of the
nucleation process itself.
The pattern of formation of multiple slips changes from cooldown of the cell
from room temperature to cooldown but remains stable for each given cooldown.
46 É. Varoquaux et al.
Fig. 5. Mean size of (positive) multiple slips vs velocity in phase winding number in
nominal purity 4 He (100 ppb 3 He): () pressure sweep from 0.4 to 24 bars at 81.5
mK (all even values of P , and 0.4, 1, 3, 5, 7 bars) - () temperature sweep at 16
bars - (◦) temperature sweep at 24 bars - (∗) drive level sweep at 24 bars, 81.5 mK
- () temperature sweep at 0 bar. For the temperature sweeps, from 14 to 200 mK
approximately, v first increases, reaches the quantum plateau and then decreases, as
shown in the insert of Fig. 1. Lines connect successive data points in the temperature
and pressure sweeps.
It seems to depend on the degree of contamination of the cell, degree which can-
not easily be controlled experimentally. The detailed microscopic configuration
of the aperture wall where nucleation takes place probably plays an major rôle
in multiple slip formation. Multiple slips are different from ‘singular’ collapses
and the underlying mechanisms responsible for both phenomena are bound to
be different as will be discussed below.
Remanent vorticity and vortex mills. Remanent vorticity in 4 He, which had
been long assumed, has been shown directly to exist by Awschalom and Schwarz
[42]. This trapped vorticity, according to Adams et al [43], either is quite loosely
bound to the substrate and disappears rapidly, or is strongly pinned. To account
for these observations, Schwarz has proposed the following formula, based on
numerical experiments, for the velocity at which vortices unpin,
κ b
vu ln( ) ,
2πD a0
D being the size of the pinned vortex and b being a characteristic size of the
pinning asperity. Thus, vortices pinned on microscopic defects at the cell walls
can in principle exist under a wide range of superflow velocities.
The Experimental Evidence for Vortex Nucleation 47
It has been suggested by Glaberson and Donnelly [10], in connection with the
critical velocity problem in an aperture, that imposing a flow to a vortex pinned
between opposite lips of the aperture would induce deformations such that the
vortex would twist on itself, undergo self-reconnections, and mill out free vortex
loops. As shown by numerical simulations of 3D flows involving few vortices
only [44], vortex loops and filaments are stable even against large deformations.
Vortices are not prone to twist on themselves and foster loops. It takes the
complex flow fields associated with fully developed vortex tangles to produce
small rings, as discussed at the Workshop [45,46]. And it takes some quite special
vortex pinning geometry to set up a vortex mill that actually works; Schwarz
has demonstrated by numerical simulations that a vortex pinned at one end
and floating along the flow streamlines with its other end free moving on the
wall develops a helical motion, a sort of driven Kelvin wave, and reconnects
sporadically to the wall when the amplitude of the helical motion grows large
enough [47]. This helical mill does churn out fresh vortices.
The above remarks on the stability of vortex loops or half-loops in their course
make it unlikely that the multiple slips be due to the production of small rings
by a vortex after having left the vicinity of the nucleation centre. Furthermore,
such a purely hydrodynamical process would depend on the velocity of the flow
only, contrary to the results shown in Fig. 5. What appears more likely is that
multiple slips are produced by a transient vortex mill of the helical type suggested
by Schwarz operating very close to the nucleation site. The pinning of the mill
vortex and its subsequent release would take place immediately after nucleation
when the velocity of the vortex relative to the boundary is still small and the
capture by a pinning site easy. This process depends on the precise details of the
pinning site configuration and of the primordial vortex trajectory, factors which
allow for the variableness of multiple slips on contamination and pressure (i.e.
single phase slip nucleation velocity) [48]. That pinning does take place close to
sites where vortices are nucleated is shown below.
References
1. W. Zimmermann Jr., C.A. Lindensmith, J.A. Flaten, J. Low Temp. Phys. 110, 497
(1998), and references therein
2. E. Varoquaux, in: Topological Defects and the Non-Equilibrium Dynamics of Sym-
metry Breaking Phase Transitions, ed. by Y.M Bunkov, H. Godfrin (Kluwer Aca-
demic, The Netherlands 2000) p. 303 and references therein
3. S. Burkhart, M. Bernard, O. Avenel, E. Varoquaux: Phys. Rev. Lett. 72, 380 (1994)
4. W. Zimmermann, Jr: J. Low Temp. Phys. 93, 1003 (1993)
5. O. Avenel, G.G. Ihas, E. Varoquaux: J. Low Temp. Phys. 93 1031 (1993)
6. E. Varoquaux, O. Avenel: Physica B 197, 306 (1994) and references therein
7. W. Zimmermann, Jr.: Contemporary Phys. 37, 219 (1996)
8. C. Josserand, Y. Pomeau: Europhys. Lett. 30, 43 (1995)
9. C. Josserand, Y. Pomeau, S. Rica: Phys. Rev. Lett. 75, 3150 (1995)
10. W.I. Glaberson, R.J. Donnelly: Phys. Rev. 141, 208 (1966)
The Experimental Evidence for Vortex Nucleation 49
41. E. Varoquaux, O. Avenel, M. Bernard, S. Burkhart: J. Low Temp. Phys. 101, 821
(1995)
42. D.D. Awschalom, K.W. Schwarz: Phys. Rev. Lett. 52, 49 (1984)
43. P.W. Adams, M. Cieplak, W.J. Glaberson: Phys. Rev. Lett. 78, 3602 (1985)
44. K.W. Schwarz: private communication to E.V.; the 3D simulations presented at
the QVD Workshop by C. Adams also show that vortex loops can undergo severe
deformations and not break apart (to appear in Euro. Phys. Lett.)
45. B.V. Svistunov: Phys. Rev. 52, 3647 (1995)
46. M. Tsubota, T. Araki, S.K. Nemirovskii: Phys. Rev. B (in press) and this Workshop
47. K.W. Schwarz: Phys. Rev. Lett. 64, 1130 (1990)
48. W. Zimmermann: Jr., J. Low T. Phys. 91, 219 (1993) J. Flaten: PhD Thesis,
Univ. of Minnesota (1997, unpublished). The large dissipation events reported by
these authors apparently fall in yet another category than the multiple slips and
the singular collapses discussed here as they persist over a number of resonator
periods.
49. P. Hakonen, O. Avenel, E. Varoquaux: Phys. Rev. Lett. 81, 3451 (1998)
50. E. Varoquaux, O. Avenel, P. Hakonen,Yu. Mukharsky: Physica B 255, 55 (1998)
Applications of Superfluid Helium
in Large-Scale Superconducting Systems
1 Introduction
Liquid helium has a wide range of uses in low temperature technology today. The
properties of liquid helium that make it particularly valuable are its low temper-
ature, persistence of the liquid state to the lowest achievable temperatures and
the existence of the superfluid state with its associate unique transport phenom-
ena. These properties combined with an ever-improving cryogenic engineering
infrastructure have allowed the development of a number of large-scale systems
that use liquid helium cooling.
There are primarily two classes of large-scale applications for liquid helium:
1. Large superconducting systems such as magnets or RF cavities for high-
energy physics accelerators, fusion systems and other magnet facilities.
2. Large space-based instruments for infrared astronomy and other fundamen-
tal studies.
Each of these applications has a unique set of requirements that in turn place
requirement on the coolant, which determines the preferred state for the liquid
helium. In many cases, the best choice for the coolant is superfluid helium (or
He II) typically in the temperature range 1.4 K to 2.1 K. The decision to select
He II as a coolant is usually driven by a combination of the lower temperature
and beneficial heat transport characteristics.
The present paper begins with an overview of the basic design features and
operating characteristics employed in several large-scale He II cooled applica-
tions. The motivation for He II cooling, the unique design features, and the
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 51–65, 2001.
c Springer-Verlag Berlin Heidelberg 2001
52 S.W. Van Sciver
1. The dipole magnet system for the Large Hadron Collider (LHC) which
is under construction at the European Organization for Nuclear Research
(CERN) laboratory in Geneva, Switzerland.
2. The recently completed outsert superconducting magnet of the 45-T Hybrid
Magnet at the National High Magnetic Field Laboratory in Tallahassee,
Florida.
3. The RF cavity system for the Teravolt Electron Synchrotron Linear Ac-
celerator (TESLA) under design at the Deutsches Elektronen-Synchrotron
(DESY) laboratory in Hamburg, Germany.
Since He II can transport large heat fluxes (compared to that of He I), its sta-
bilizing effect on magnet systems can be very beneficial. The understanding of
the relevant thermal processes in He II has been a subject of numerous recent
experimental investigations.
The principal disadvantage to the use of He II cooling in large superconduct-
ing systems originates from the increased complexity and cost of the cryogenic
system. In very simple terms, the cost of refrigeration scales with the Carnot
factor, which at low temperatures goes as T−1 . Therefore, a unit of power de-
posited in a He II cryogenic system operating at 1.8 K will require a refrigeration
system 2.3 times larger than the equivalent system operating at 4.2 K. Since the
cost of refrigeration systems scales monotonically with power, the advantages of
superconductor performance and stability must outweigh this cost issue. Thus,
the selection of a He II cooling system is only warranted in cases where the
overall system design is optimized by this approach. What follows are three such
examples.
The LHC accelerator currently under construction at CERN will, when complete,
operate the world’s largest He II cryogenic system [1]. In this case, all the main
ring magnets (dipole and quadrupoles) are made with NbTi superconductor and
cooled to 1.9 K in a He II bath pressurized to atmospheric pressure. The dipole
magnets, shown in cross section in Fig. 1, operate at 8.3 T, which is higher than
previously built accelerator dipoles such as for the SSC and therefore require the
lower temperature and improved thermal stability that He II operation affords.
There are eight refrigeration plants around the 26.7 km circumference of the
accelerator, each cooling one 45◦ sector. The 1.8 K refrigeration capacity for
each sector is between 2.1 and 2.4 kW.
The system uses pressurized He II to cool the magnets, which reduces the pos-
sibility of trapped vapor within the windings of the coil. The use of pressurized
He II to cool magnets was first demonstrated on a large scale in a development
program for the Tore Supra tokamak plasma experiment in France [2]. The ap-
proach requires the use of a low pressure of He II heat exchanger as part of
the installation. In the case of LHC, the heat exchanger is a corrugated pipe
within the magnet vessel and immersed in the He II reservoir. A schematic of
this system is shown in Fig. 2. The heat exchanger is partially filled with satu-
rated He II. The two phase He II heat exchanger provides a much more efficient
heat removal method than can be obtained through counterflow heat transport
in the bulk fluid. Although heat transport is very efficient in bulk He II, the long
distances between refrigeration stations would demand very large He II cross
sections to minimize the temperature gradient.
54 S.W. Van Sciver
Fig. 2. Cooling system for LHC dipole magnets with two-phase heat exchanger
path. The NHMFL hybrid outsert is the first CICC magnet that employs He II
cooling. As in the case of LHC, He II cooling is applied to achieve the highest
current densities in the conductor, while providing a good thermal stabilizing
environment.
The TESLA electron collider is one of the proposed major accelerator systems
that will follow the completion of the LHC [5]. This system is a linear electron
collider as compared to a circular LHC machine, which accelerates protons. Lin-
ear machines do not require the large number of dipole bending magnets that
are the signature of the hadron colliders. Rather the electron colliders depend
on high gradient RF cavities for accelerating the beam and use a relatively small
number of magnets for beam steering and focusing. The RF cavities can be either
56 S.W. Van Sciver
principal technical issue associated with the design of the He II system pertains
to the behavior of two-phase He II/vapor in near horizontal channels of relatively
large diameter.
A H e a t T ra n s fe rre d to H e II
o w e r (W )
B
I2R
H e a t G e n e ra tio n
T im e
Fig. 7. Stability criterion for He II cooled conductor
where the pressure gradient terms take into account the frictional losses and
change of internatl energy with pressure. As before, this relationship has success-
fully modeled the behavior of heat transport in He II for a variety of experimental
systems. An interesting observation about this formulation is the fact that He II
will display the Joule Thomson effect when experiencing a pressure drop through
an insulated tube [12]. The result will be an increase in temperature, which will
generally have a negative impact on applications.
An example of heat transport in forced flow He II is displayed in Fig. 8.
The flow is from left to right and a heat pulse is deposited at x = 0. The
plot displays calculated and experimental time-dependent temperature profiles
at different locations along a channel containing He II. Note that the peak in
the temperature profile broadens due to thermal diffusion as time progresses
and the pulse propagates. Also, the location of the peak moves at approximately
the speed of the fluid, suggesting a rough method of measuring fluid velocity in
such systems. At these high velocities, the Joule Thomson effect contributes by
producing a gradual increase of the fluid temperature as seen by the increase in
the baseline temperature as the fluid moves through the tube.
From the application viewpoint, the steady-state and transient heat transport
characteristics a fully turbulent He II can be understood in terms of the He II
diffusion and energy equations. There are limits to this representation and these
limits need to be explored. Work is continuing on the development of turbulence
and heat transport at very high velocities. This seems to be an area where some
fundamental work would be able to contribute.
was whether the He II pressure drop and associated fluid dynamic properties
would be unique. Since that time, a considerable body of research has come to
support the notion that the dynamic behavior of forced flow He II is essentially
similar to that of classical fluids. Thus, He II when flowing through tubing dis-
plays a pressure drop that can be described by classical correlation [13]. This
result has been supported by theoretical investigations that suggest the two flu-
ids are coupled through turbulent interactions and thus flow together. Figure 9
displays recent measurements of the friction factor for He II flowing in a 10 mm
ID tube at Re>107 (u> 10 m/s). These data continue to support the previous
observations.
Forced flow He II at high Reynolds number has more recently become inter-
esting as a test fluid for basic fluid dynamic investigations [14]. The fluid has a
very small kinematic viscosity allowing high Reynolds numbers to be achieved
in sub-sonic flows. A question that continues to be raised is to what extent can
one ignore superfluid effects. This is still an open issue and a subject for further
investigation. In a companion paper at this conference, we report on drag coeffi-
cient measurements for a sphere in flowing He II [15]. The measurements suggest
a temperature dependence to CD not seen in normal fluids and this dependence
appears to correlate with the normal fluid density, ρn . Clearly, this topic needs
further experiment supported by theoretical analysis.
62 S.W. Van Sciver
4 .1 m s t r a ig h t
4 .6 m lo o p
f 2 m c o p p e r lo o p
1 .2 m s t r a ig h t
V o n K a r m a n - N ik u r a d s e
C o le b ro o k
0 .0 1
0 .0 0 1
5 6 7 8
1 0 1 0 1 0 1 0
R e
Fig. 9. He II friction factor for pipe flow at high Reynolds number
An area that has received recent investigation driven by the needs of large accel-
erators such as LHC or TESLA is two phase He II/vapor flow. In many of these
large systems, there are tubing sections partially filled with He II and in near
horizontal configuration. This operating condition leads to some interesting phe-
nomena due to the existence of the free surface. As is in the case of single phase
He II systems, the heat transfer and fluid dynamics behavior are of interest to
designers. Further, because of the free surface and relative velocity between the
liquid and vapor phases, there are a variety of flow stability issues that need to
be addressed. These are topics of current investigation [16,17,18]. Modeling these
systems is complex due to the interaction between the two phases [19,20]. To
date, most studies have treated the He II as a classical fluid with heat transport
character governed by the He II energy equation. Clearly, this fluid system is
complex and further experimental and theoretical work is needed.
ence between the He II bath and exciting vapor provides the thermomechanical
pressure head to hold back the liquid.
Fig. 10. Schematic of method to use porous plugs to manage He II fluid level
4 Conclusions
Superfluid helium has become an engineering fluid for a number of technical
applications in superconductivity and space-based instrumentation. A consider-
able volume of practical data has been accumulated through the development of
these systems. Steady-state and slowly varying transient thermal processes can
be described to be a diffusive process much like conduction. Forced flow at rela-
tively high velocities appears to obey clssical correlations. There are a number of
areas where further study is needed. We do not have adequate understanding of
the development of turbulence particularly at high heat fluxes. Also, two-phase
64 S.W. Van Sciver
flow is an entire subject that has only recently been under investigation. Future
applications will no doubt require additional investigation.
Acknowledgements
The National High Magnetic Field Laboratory is jointly funded by the National
Science Foundation and the State of Florida.
References
1. P. Lebrun, Cryogenics for the Large Hadron Collider. IEEE Trans. On Applied
Super., Vol 10, 1500 (2000)
2. G. Claudet and R. Aymar, Tore Supra and He II Cooling of Large High Field
Magnets, Adv. Cryog. Engn. 35A, 55 (1990)
3. J.R. Miller, et al., An Overview of the 45-T Hybrid Magnet System for the
NHMFL, IEEE Trans. On Magnetics, Vol 30, 1563 (1994)
4. S.W. Van Sciver, et al., Design, Development and Testing of the Cryogenic System
for the 45-T Hybrid, Adv. Cryog. Engn. Vol 41, 1273 (1996)
5. R. Brinkmann, et al., Conceptual Design of a 500 Gev e+e- Linear Collider with
Integrated X-Ray Laser Facility, DESY 1997-048 ECFA 1997-182, May 1997
6. G. Horlitz, et al., The TESLA 500 Cryogenic System and He II Two Phase Flow,
Cryogenics Vol 37, 719 (1997)
7. S.K. Nemirovskii and A.N. Tsoi, Transient Thermal and Hydrodynamic Processes
in Superfluid Helium, Cryogenics Vol 29, 985 (1989)
8. T. Shimazaki, M, Murakami and T. Iida, Temperature measurement in Transient
Heat Transport Phenomena Though a Thermal Boundary Layer in High Vortex
Density, Adv. Cryog. Engn. Vol 41, 265 (1996)
9. S.W. Van Sciver, Chap. 10: Helium II (Superfluid Helium), in: Handbook of Cryo-
genic Engineering, J.G. Weisend II (ed.). Taylor & Francis (1998)
10. P. Seyfert, Practical Results on Heat Transfer in Superfluid Helium, in: Stability
of Superconductors in He I and He II, IRR Commission A 1/2 (1981), pp. 53-62
11. B. Rousset, Pressure Drop and Transient Heat Transport in Forced Flow Single
Phase He II at High Reynolds Number, Cryogenics Vol 34 supplement, 317 (1994)
12. P.L. Walstrom, Joule Thomson Effect and Internal convection Heat Transfer in
Turbulent He II Flow, Cryogenic Vol 28, 151 (1988)
13. P.L. Walstrom, et al., Turbulent Flow Pressure Drop in Various He II Transfer
System Components, Cryogenics Vol 28, 101 (1988)
14. R.J. Donnelly, Ultra High Reynolds Number Flows Using Cryogenic Helium: An
Overview, in: FLow at High Reynolds and Rayleigh Numbers, R.J. Donnelly and
K. Sreenivasan (eds.), Springer (1998)
15. M.R. Smith, Y.S. Choi and S.W. Van Sciver, The Temperature Dependent Drag
Crisis on a Sphere in Flowing He II, (this publication)
16. P. Lebrun, et al., Cooling Strings of Superconducting Devices Below 2 K: The
He II Bayonet Heat Exchanger, Adv. Cryog. Engn., Vol 43, 419 (1998)
17. B. Rousset, et al. Behavior of He II in Stratified Counter-Current Two Phase
FLow, in: Proceedings of the ICEC17, Institute of Physics Publishing (1998) pp.
671-674
18. J. Panek and S.W. Van Sciver, Heat Transfer in a Horizontal Channel Containing
Two Phase He II, Cryogenics Vol 39 (1999)
Applications of Superfluid Helium 65
19. L. Grimaud, et al., Stratified Two Phase Superfluid Helium Flow, Cryogenics Vol
37 (1997)
20. Y. Xiang, et al. Numerical Study of Two Phase He II Stratified Channel with
Inclination, IEEE Trans on Applied Super. Vol 10, 1530 (2000)
21. M. DiPirro and P.J. Shirron, The SHOOT Orbital Operations, Cryogenics Vol 32
(1992)
22. J. Panek, Y Zhao and S.W. Van Sciver, Liquid Level Control Using a Porous Plug
in a Two Phase He II System, Adv. Cryog. Engn., Vol 43, 1401 (1998)
The Temperature Dependent Drag Crisis
on a Sphere in Flowing Helium II
1 Introduction
The low kinematic viscosity (ν = η/ρ, where η and ρ are dynamic viscosity
and total density, respectively) of liquid helium makes it an attractive fluid
for modern dynamical similarity studies, where one wishes high Reynolds num-
bers (Re=Ud/ν, where U and d are the characteristic velocity and dimension
of the flow field) without transonic effects. While research suggests that helium
above 2.176 K (He I) behaves as a classical fluid, He II (the liquid phase below
2.176 K) is a quantum fluid with a wide range of non-classical macroscopic prop-
erties. Still, studies have found that classically generated turbulence in He II may
behave classically in certain experiments [1,2,3]. Uncertainties about the micro-
scopic character of classically generated turbulence in He I and He II motivated
the previous work [4,5] in which the form drag on a sphere in flowing He I and
He II was calculated from the observed pressure distribution over the surface.
If the critical Reynolds number for the drag crisis in He II is temperature
dependent, then the dimensionless equations of motion for the two fluid system
must scale with other dimensionless parameters, in addition to or instead of the
Reynolds number. In the classical fluid dynamics of ordinary fluids (including
He I), dynamical similarity and scaling arguments for expressing experimental
data in terms of Reynolds number, coefficients of drag, lift and so on, spring
rigorously from non-dimensionalizing the Navier–Stokes equations. Expressing
He II data in terms of an effective Reynolds number however, has been more
of an empirical convenience. Perhaps one reason for this is the empirical nature
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 66–72, 2001.
c Springer-Verlag Berlin Heidelberg 2001
The Temperature Dependent Drag Crisis 67
2
dνn ρn dP dT ρs ρn dνns d 2 νn
ρn =− − ρs S − + Fns + η 2 (2)
dt ρ dx dx 2ρ dx dx
2
Fns = − αkρs Lνns (3)
3
Now, we define non-dimensional quantities in terms of representative veloci-
ties, distances and temperatures as follows;
x0 = x/D t0 = U · t/D
L0 = L · D2 T0 = (Tλ − T )/Tλ
νn0 = νn /U S0 = Tλ S0 /U 2
νs0 = νs /U P0 = P/ρU 2
where the zero subscript refers to the dimensionless quantity and Tλ = 2.176 K is
the lambda point. Upon substituting these quantities into (1-3), and rearranging
terms a bit, we arrive at;
2 2
dνs0 dP0 dT0 αk 1 − δ dνns0
=− − S0 + 3 · L0 νns0 + (4)
dt0 dx0 dx0 UD 2 dx0
2 2
dνn0 dP0 δ dT0 3 αk δ dνns0
=− + · S0 − · L0 vns0 − (5)
dt0 dx0 1−δ dx0 UD 2 dx0
1 1 d2 νn0
+ ·
1 − δ Re dx20
The terms involving the temperature gradient drive counterflow, in which the
two fluids flow in opposite directions, and heat is transported within the system.
68 Y.S. Choi, M.R. Smith, and S.W. Van Sciver
Note that the effect of a thermal gradient upon the normal fluid component
(Eq. 6) becomes larger than 1 − δ → 0 where δ ≡ ρs /ρ. This is comforting,
since it agrees well with our expectations. Furthermore, the classical equations
of motion are recovered in the limit δ → 0 . From a more general perspective,
Eqs. (4) and (6) describe a wide spectrum of physics, which goes beyond fully
developed co-flowing turbulent fields. The driving terms for such non-classical
phenomena as counterflow however (we include for consideration at this point
terms involving vns0 ), must affect the dynamics of all flows at suitably small
length scales. Since the co-flowing condition (vns0 = 0) generally depends upon
mutual friction, it must depend upon vns0 = 0 across some range of length scales
within the flow.
In addition to the many factors of (1-δ) in (4) and (6), which underscore
the strong role of temperature in the dynamics of He II, there is one additional
dimensionless quantity, UD/(2/3)αk, which is tied into the mutual friction. Al-
though we will be discussing co-flowing turbulence, where vns0 may zero, L0 may
still be quite large. Thus, the term on the whole may fluctuate dramatically on
a local scale.
For the sake of the argument at hand however, consider the case where vns0
is strictly zero (fully coupled superfluid turbulence), with zero temperature gra-
dient. Then at suitably large length scales, we are left with
dvs0 dP0
=− (6)
dt0 dx0
the drag crisis itself changes character and magnitude as the normal component
vanishes (ρn → 0). Ultimately in this limit, one expects to recover some sort of
potential flow.
To test the two fluid equation scaling arguments, we plot in Fig. 4 the drag
coefficients versus Re (1-δ), as suggested by [5]. The open symbols correspond
to He I data and the closed symbols are for He II.
Comparing Fig. 4 to Fig. 2, we note that the scaling shifts the minimum in
the drag coefficient to approximately the same value of the modified Reynolds
number for each temperature. That value, Re (1-δ) ≈ 2.3 × 105 , is in reasonable
agreement with the minimum in the classical drag coefficient curve at approxi-
mately Re = 3 × 105 .
4 Conclusion
We have measured the pressure distribution over the surface of a sphere in flowing
He II as a function of Reynolds number. The He II data shows clear evidence
of a drag crisis at approximately the same Reynolds number. The coefficients of
drag have a minimum value at same modified Reynolds number based on scaling
the He II two fluid equations, Re ×(1 − δ) ≈ 2.3 × 105 . The variability in the
He II data confirms a temperature dependence to the turbulent transition within
boundary layer.
72 Y.S. Choi, M.R. Smith, and S.W. Van Sciver
Acknowledgments
We wish to thank the National High Magnetic Field Laboratory and the De-
partment of Energy, Division of High Energy Physics for their financial support.
Thanks to David K. Hilton for useful conversation and Scott Maier for technical
assistance.
References
1. P.L. Walstrom, J.G. Weisend II, J.R. Maddocks and S.W. Van Sciver, Turbulent
pressure drop in various He II transfer system components, Cryogenics 28, 101,
1988.
2. D.C. Samuels, Velocity matching and Poiseuille pipe flow of superfluid helium,
Phys. Rev. B 46, 11714, 1992.
3. C.F. Barenghi, D.C. Samuels, G.H. Bauer and R.J. Donnelly, Numerical evidence
for vortex-coupled superfluidity: Quantized vortex lines in an ABC model of tur-
bulence, Phys. Fluids 9, 2631, 1997.
4. M.R. Smith and S.W. Van Sciver, Measurement of the pressure distribution and
drag on a sphere in flowing He I and He II, Advances in Cryogenic Engineering,
Vol 43, 1473, 1998.
5. M.R. Smith, D.K. Hilton and S.W. Van Sciver, Observed drag crisis on a sphere in
flowing He I and He II, Physics of Fluids, Vol. 11, No.4, 1999.
6. A similar analysis was originally presented in Evolution and Propagation of Tur-
bulence in Helium II, Ph.D. Thesis, M.R. Smith, 1992, University of Oregon.
7. R.J. Donnelly, Quantized Vortices in Helium II, Cambridge, New York, 1991.
Experiments on Quantized Turbulence
at mK Temperatures
1 Background
The renaissance of interest in the turbulent hydrodynamics of HeII has led to
the realisation that, in many respects, it exhibits unexpected similarities to anal-
ogous flows in classical fluids at high Reynolds number [1,2,3]. Unlike a classical
fluid, HeII is well described by a two-fluid model, with a normal (dissipative)
component mutually interpenetrating with a superfluid (inviscid) component
with quantized circulation
h
κ = vs .dl = n (1)
m4
where the integral is taken around a loop enclosing the vortex, vs is the superfluid
velocity, m4 is the 4 He atomic mass and the quantum number n is an integer.
The flow properties of HeII in an open geometry are dominated by singly-
quantized vortex lines [4], linear singularities around which the superfluid flows
at tangential velocity vs . Unless velocities are kept extremely small, the liquid
flowing through a tube becomes filled with a tangled mass of such vortex lines.
Because of their quantization, they represent a particularly simple form of tur-
bulence. In that the vortex cores can be considered as part of the normal fluid
component, but the encircling superflow field in accordance with (1) is of super-
fluid component, vortices provide a weak coupling (mutual friction) between the
two components. So it is not at all clear, at first sight, why this complex liquid
system should ever behave like a single-component classical fluid. The question
has recently been discussed in considerable detail by Vinen [3]. One of the aims
of the present project is to establish the properties of the turbulent liquid when
it really does consist of just a single component, i.e. in the low temperature limit
where the normal fluid density is negligible.
In the conventional scenario at higher temperatures 1 < T < Tλ , a vortex
tangle can be maintained by the work done by the driving force, which could be
e.g. a pressure or temperature gradient for bulk flow, or for thermal counterflow
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 73–79, 2001.
c Springer-Verlag Berlin Heidelberg 2001
74 S.I. Davis et al.
(in which the normal and superfluid components move in opposite directions),
respectively. On removal of the driving force, the tangle decays according to the
Vinen [5] equation
dL 2
= −χ2 L (2)
dt m4
where L is the length of vortex line per unit volume and χ2 is a (weakly
temperature-dependent) dimensionless constant. The physical mechanism driv-
ing the decay has been discussed by Schwarz [6], who concluded that it involves
crossings and consequent reconnections of lines. The rapid self-induced motion
of the resultant sharp cusps through the normal fluid is dissipative, and causes
the rounding off of cusps with consequent line shrinkage. The presence of the
normal fluid component is thus a key component of the decay mechanism.
Details of the decay process, e.g. the existence of two distinct decay rates
[5,7], are not at all well understood, partly because it sometimes seems to in-
volve turbulence in the normal fluid component as well as in the superfluid. The
experiment that we describe below avoids this complication. Vortex decay is in-
vestigated at sufficiently low temperatures (in the mK range) that normal fluid
component as such is absent. Under these conditions, it is far from obvious how
the tangle will decay, but it appears that there are two possibilities –
1. Essentially the same decay process occurs as above 1K, but now driven by the
dilute phonon gas rather than by normal fluid, or
2. A new decay mechanism comes into play.
F ie ld
e m is s io n tip
T o p e le c tro d e
O s c illa tin g
H V g rid
B o tto m e le c tro d e
Fig. 1. The experimental arrangement (schematic). Some of the ions created by the
field emission tip may get trapped on vortices created previously by the oscillating grid,
thereby reducing the signal arriving at the collector. The perforated top and bottom
electrodes complete the double capacitor needed to excite oscillations of the grid, and
the Frisch grid screens the collector from the approaching charge.
3 The Experiment
The electrode structure used for the experiment is shown schematically in Fig. 1.
The operating procedure was performed in two stages. First, a high constant
voltage (usually 500 V) was applied to the vortex-generating grid and a periodic
driving voltage of ±270 V was applied to one of the adjacent plates. The drive
76 S.I. Davis et al.
4 Preliminary Results
Figure 2 shows a typical sequence of ion signals. The first signal is a reference,
recorded before the grid had been vibrated, and the others show how the signal
gradually recovered after the grid vibration had been completed. It is evident
that there is significant attenuation as a result of the grid oscillation – demonstat-
ing immediately that the technique described in the preceding sections enables
us both to create and to detect vorticity in the mK temperature range.
4 0 0 0
3 0 0 0
2 0 0 0
1 0 0 0
0 1 0 0 0 2 0 0 0 3 0 0 0 4 0 0 0
Fig. 2. A set of collector signals. The duration of each of them is ∼ 200 μs, and a
period of 1.5 s in real time separates each signal from its neighbour. The first signal is
for reference, recorded before the grid was vibrated.
In Fig. 3, the signal heights are plotted as a function of time for several
temperatures T . It appears that for T < 70 mK nothing changes, within ex-
perimental error: the decay mechanism is apparently temperature-independent
within this range.
5 Discussion
Our preliminary data are too scattered for us to be able to draw definite con-
clusions about the form of the decay and there is, in any case, no theoretical
Quantized Turbulence at mK Temperatures 77
4 6 8 1 0 1 2 1 4 1 6 1 8
Fig. 3. Signal amplitudes, showing their evolutions as a function of time t, for several
temperatures.
1 2
1 0
0
4 6 8 1 0
Fig. 4. Plot of a typical set of data to test the applicability of equation (3). S and S0
are respectively the amplitudes of a signal and of the reference signal.
of the strongly kinked lines that are produced, resulting [3] in Kelvin waves and
a cascade of energy towards smaller and smaller length scales until it is radiated
as sound.
A difficulty in interpreting the results is that the absolute vortex line densities
are unknown, because the ion-vortex trapping cross-section is unknown under
the conditions of the experiment. We can obtain a very approximate estimate
from the measured linewith of the grid resonance which, for typical electrode
potentials and driving amplitudes implies that the energy dissipation of the grid
is (0.4±0.2)μW; insertion of this value in (2), on the assumptions that all of the
dissipation goes into vortex creation, χ takes the same value as above 1K and
that the vortex tangle remains mostly between the electrodes, yields a steady
state line density of ∼1010 m−2 . Any or all of the assumptions could be in error,
however, and the only unambiguous way to clarify the situation will be through
direct measurement of the ion-line trapping cross-section in a rotating cryostat
where the line density is known precisely.
6 Conclusions
In conclusion, we would emphasize the preliminary character of these results,
and the large number of unknowns. As pointed out above, we know nothing
about the spatial distribution of the vorticity, although we imagine that it stays
mostly between the top and bottom electrodes of the triple capacitor in Fig. 1.
Quantized Turbulence at mK Temperatures 79
Acknowledgements
References
1. S.R. Stalp, L. Skrbek, R.J. Donnelly: Phys. Rev. Lett. 82, 4831 (1999)
2. C.F. Barenghi: J. Phys.: Condens. Matter 11, 7751 (1999)
3. W.F. Vinen: Phys. Rev. B 61, 1410 (2000)
4. R.J. Donnelly: Quantized Vortices in He II (Cambridge University Press, Cam-
bridge 1991)
5. W.F. Vinen: Proc. R. Soc. A 242, 493 (1957)
6. K.W. Schwarz: Phys. Rev. B 18, 245 (1978); 31, 5782 (1985); 38, 2398 (1988)
7. K.W. Schwarz, J.R. Rozen: Phys. Rev. Lett. 66, 1898 (1991)
8. A.L. Fetter: Phys. Rev. A 136, 1488 (1964)
9. M.I. Morell, M. Sahraoui-Tahar, P.V.E. McClintock: J. Phys. E: Sci. Instrum. 13,
350 (1980)
10. L. Meyer, F. Reif: Phys. Rev. 123, 727 (1961)
11. R.M. Bowley, P.V.E. McClintock, F.E. Moss, P.C.E. Stamp: Phys. Rev. Lett. 44
161 (1980)
12. R.M. Bowley, P.V.E. McClintock, F.E. Moss, G.G. Nancolas, P.C.E. Stamp: Phil.
Trans. R. Soc. Lond. A 307, 201 (1982)
13. D.R. Allum, P.V.E. McClintock, A. Phillips, R.M. Bowley: Phil. Trans. R. Soc.
Lond. A 284, 179 (1977)
14. R.J. Donnelly, P.H. Roberts: Proc. R. Soc. Lond. A 312, 519 (1969)
15. M. Tsubota, T. Araki, S.K. Nemirovskii, J. Low Temperature Phys. 119, 337
(2000)
Grid-Generated He II Turbulence
in a Finite Channel – Experiment
1 Introduction
Superfluid turbulence has long been an area of study, with an emphasis largely on
flows created by applying a heat current in He II; i.e. on thermal counterflow[1].
It is an advantage that, experimentally, counterflow turbulence requires no mov-
ing parts for its generation. The connection of this type of flow with classical
turbulence, however, is not obvious. More recently, it has been of some interest to
explore this connection by generating turbulent flows in He II in a similar manner
as for classical fluids. In particular, turbulence created in the wake of a grid can
create nearly homogeneous and isotropic turbulence (HIT)[2], and application
of this procedure to He II has led to new insights and a few surprises[3,4,5]. In
particular, a deep similarity appears to exist between grid turbulence in classical
fluids and in He II, a quantum fluid.
In this article, we focus on the experimental apparatus and techniques used
to generate grid turbulence in He II, and discuss the observed decay of the vor-
tex line density behind a towed grid. Measurements of second sound attenuation
allow detection of up to six orders of magnitude of decaying vortex line density
L, which can be converted into roughly eight orders of magnitude of decaying
turbulent energy[6] - at present hardly a feasible goal for any laboratory experi-
ment on classical turbulence. In a companion article in this book, we will further
interpret the data in terms of classical hydrodynamics. The underlying quantum
nature of this experiment has been recently discussed by Vinen [7].
2 Experimental Setup
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 80–86, 2001.
c Springer-Verlag Berlin Heidelberg 2001
Grid-Generated He II Turbulence 81
Motor Control
and Counter
16 bit ADC
200MHz Pentium
GPIB, Labview
G
P
I
B
Function Vacuum Tight Sliding Seal
B Generator
U Vacuum
Pump
S
Lock-in
Amplifier
Amplifier
60 mF Grid Generated
Turbulence
Grid
Second Sound
Bias Voltage Receiver
100 Volts DC
Second Sound
Transmitter
Second Sound
LR-110 Standing Wave
Resistance
Bridge
Germanium
Thermometer 1 cm x 1 cm x 29 cm
Channel
Heater
LR-130
Temperature
Controller
steel pulling rod of diameter 0.24 cm which exits the cryostat via a pair of tight
sliding seals. The space between the seals is continually evacuated to prevent the
introduction of impurities inside the cryostat during the experiment. Above the
cryostat, the rod is attached to a computer controlled linear servo motor that
positions the grid with 0.01 cm accuracy and provides the towing velocity, vg ,
up to 2.5 m/s. This enables the exploration of a wide range of mesh Reynolds
numbers 2 × 103 ≤ ReM = vg M ρ/μ ≤ 2 × 105 , where μ is the dynamic viscosity
82 J.J. Niemela, L. Skrbek, and S.R. Stalp
of the normal fluid and ρ the total density. It is worth noting that the linear
servo motors, developed at the University of Oregon, accomplish this precise
positioning without the problems associated with electrical switching noise and
mechanical resonances that are characteristic of the more commonly used stepper
motors.
The channel of square cross section is 29 cm long and has a width d = 1 cm.
It was manufactured by an electroforming process with a tolerance of 25 μm and
a surface roughness less than 0.5 μm. The channel is suspended vertically in the
helium cryostat and during the measurement is totally submersed in superfluid
helium, which enters the channel via 16 holes, 500 μm in diameter, placed near
the top end. Except for these, the channel is leak-tight. There is no free surface of
the liquid inside the channel during the experiment, and second sound coupling
to the free surface of the bath is minimized by the use of the small diameter
holes for mass transfer between the channel and the bath.
The temperature is measured and controlled via a germanium resistance
thermometer and heater placed in the He II bath. The thermometer is calibrated
against the saturated vapor pressure with a temperature accuracy better than 1
mK and a resolution within 10 μK[4]. Details of possible temperature fluctuations
inside the channel that might occur during and just after pulling the grid are
unknown, as the thermometer is located outside the channel and its response
time is about 100 ms. For maximum cooling, the main bath is pumped via a 20
cm diameter line connected to a 300 CFM vacuum pump and a 1400 CFM roots
blower. This enables bath temperatures down to 1 K to be reached easily. It
is difficult, however, to assure temperature stability below about 1.1 K, so this
marks the lowest reliable temperature for the experiment. Experiments close to
the lambda temperature (above 2 K) require special care (see below) mainly due
to the strong temperature dependence of the second sound velocity, u2 .
To probe the quantized vortex line density resulting from pulling the grid,
we excite and detect second sound using vibrating nuclepore membranes 9 mm
in diameter mounted flush on opposing walls of the channel. These 6 μm thick
polycarbonate membranes have a dense distribution of 0.1 μm holes and on
one side is evaporated an approximately 95 Å thick layer of gold, which is then
pressed against the channel wall. This gold layer forms one electrode of a ca-
pacitor transducer, with the other being a brass electrode pressed towards the
opposite side of the membrane from outside the channel. Applying an ac signal
(∼ 0.3 − 1 VP P ) in addition to a dc bias (∼ 100 V) results in an oscillatory
motion of the membrane. In He II, the normal fluid is clamped by viscosity in-
side the small holes of the membrane, while the superfluid component passes
freely, thereby exciting second sound, an entropy wave in He II. Directly across
the channel the second sound wave produces a corresponding oscillation of the
other membrane (i.e., the receiver) and the induced signal is input to a lock-in
amplifier referenced to the transmitter frequency. The channel acts as a second
sound resonator. The excitation amplitude is adjusted to be the upper half of
the linear response range and the n-th harmonic- with n about 50- of the funda-
mental frequency is used: typically 30 − 40 kHz. A Lorentzian resonance peak is
Grid-Generated He II Turbulence 83
It can be shown[4] that for arbitrary attenuation one has to use the more
general formula
8u2 1 + p2 P + 2p2 P + p4 P 2
L= ln √ (3)
πBκd 1 + P + 2P + P 2
where p = A0 /A and P = 1 − cos(2πdΔ0 /u2 ) that for small dΔ0 /u2 reduces
to (2). It is essential to use formula (3), as using the approximate formula (2)
at some experimental conditions leads to results that are more than an order of
magnitude off!
It is important to consider the time response of the measuring system with
regard to the finite velocity of second sound. Our calculations show that in most
84 J.J. Niemela, L. Skrbek, and S.R. Stalp
1 0
6
T = 1 .7 5 K
) 5
-2
1 0 p o w e r -3 /2
v o r te x lin e d e n s ity ( c m
4
1 0
3
1 0
2
1 0
1
1 0
0 .1 1 1 0 1 0 0
tim e ( s )
Fig. 2. The log-log plot of the decaying vortex line density versus time after grid passes
2 mm above the measuring volume. Each decay curve represents an average of three
identical pulls. The decay curves, in order, correspond to ReM = 2×105 (the uppermost
one), 1.5 × 105 , 105 , 5 × 104 , 2.5 × 104 , 104 , 5 × 103 and 2 × 103 . For each ReM , the
decaying vortex density displays an inertial range with power law exponent -3/2.
cases (except close to the lambda temperature) there is negligible error intro-
duced into the deduced vortex line density, for the following reasons. Immediately
after the grid is towed through the measuring volume the quality factor is very
low, of order unity, and the detecting system can be described rather as a second
sound pulse technique with the characteristic time response given by the time
of flight d/u2 ∼ 10−3 s, where u2 ∼ 20 m/s [10]. As the turbulence decays, the
characteristic time constant increases with the (temperature dependent) qual-
ity factor. Without the vortex tangle, the typical linewidth of the second sound
resonance is 20-500 Hz, the typical frequency used is 30-40 kHz, so the quality
factor reaches 60-2000 and the time response gradually rises to about 0.1-1 sec
at the very end of the decay, where it constitutes an error of less than 1%.
There are also other time restrictions than the finite velocity of second sound.
As it takes a time τg ∼= d/vg to tow the grid through measuring volume, we use
only that data obtained on the time scale longer than τg and also exceeding 8τLI ,
where τLI is the time constant of the lock-in amplifier used for detection of the
amplitude of the second sound signal. Another time scale restriction involves
consideration of how soon the flow can be assumed as nearly HIT [6]. It was
discussed in[11] that a physical criterion to estimate this time leads to about
1-2 turnover times of the largest eddies present in the flow, i.e., those of the size
of the channel. It follows that the minimum time needed to assume nearly HIT
conditions is about 1-2 widths downstream from the grid, or using the Taylor
Grid-Generated He II Turbulence 85
-4
2 .0 x 1 0
-4
1 .5 x 1 0
/s )
2
ν ' (c m
-4
1 .0 x 1 0
-5
5 .0 x 1 0
1 .1 1 .2 1 .3 1 .4 1 .5 1 .6 1 .7 1 .8 1 .9 2 .0 2 .1 2 .2
T e m p e ra tu re (K )
Fig. 3. The effective kinematic viscosity of the superfluid turbulence ν deduced from
vorticity decay data at various temperatures, and assuming a value of the Kolmogorov
constant consistent with classical experiments. The dashed line represents ν = η/ρ
which has a dissimilar temperature dependence indicative of quantization effects, where
η is the normal fluid viscosity and ρ is the total density of helium II.
frozen hypothesis, the time it takes the towed grid to pass 1-2 widths of the
channel.
The data acquisition process can be briefly described as follows. The cryostat
is filled with liquid helium, the bath being pumped and the temperature con-
trolled to a desired value. With the grid “parked” towards the top of the channel,
the second sound resonance curve is measured and fitted to a Lorentzian, giving
the value of Δ0 . The grid is then lowered to the bottom and after a necessary
waiting time (typically 2 minutes) pulled through the channel such that its veloc-
ity is constant and equal vg for at least 5 cm below and above the experimental
volume. It is then slowed down and smoothly ”parked” against the top again.
The data acquisition is triggered when the grid passes a predetermined position
2 mm above the measuring volume, as determined by a signal from a photodiode.
Typically 100 s of data are recorded at a rate of 100 Hz from the output signal
of the lock-in amplifier, representing the recovering second sound standing wave
amplitude, A(t). After another waiting period to ensure that the turbulence has
decayed down to a negligibly low level, the reference amplitude A0 is read. The
decaying vortex line density is then calculated using formula (3). It implicitly
assumes that there is no additional attenuation of the second sound that might
be caused by a turbulence in the normal fluid.
Since the second sound is transmitted and detected via membranes on oppos-
ing sides of the channel, we obtain information from a measuring volume of order
d3 ∼
= 1 cm3 . It is this natural integration that bypasses tedious statistical analysis
involved in any local velocity measurements in conventional turbulence, provides
86 J.J. Niemela, L. Skrbek, and S.R. Stalp
Acknowledgements
The towed grid He II experiment has been developed at the University of Oregon
over many years. We would like to acknowledge R.J. Donnelly, W.F. Vinen and
M.R. Smith for their valuable contributions to the conception and design of this
experiment. This research was supported by NSF under grant DMR-9529609.
References
1. R.J. Donnelly: Quantized vortices in helium II. Cambridge University Press (1991)
2. G. Comte-Bellot, S. Corrsin: J. Fluid Mech. 25, 657 (1966); 48, 273 (1971)
3. M.R. Smith: Evolution and propagation of turbulence in helium II. PhD Thesis,
University of Oregon, Eugene (1992)
4. S.R. Stalp: Decay of grid turbulence in superfluid helium. PhD Thesis, University
of Oregon, Eugene (1998)
5. M.R. Smith, R.J. Donnelly, N. Goldenfeld, W.F. Vinen: Phys. Rev. Lett. 71, 2583
(1993)
6. L. Skrbek, J.J. Niemela, R.J. Donnelly: Phys. Rev. Lett. 85, 2973, (2000)
7. W.F. Vinen: Phys. Rev. B 61, 1410 (2000)
8. [Link], W.F. Vinen: Proc. Roy. Soc. London A238, 204 (1954); 238,215(1954)
9. C.F. Barenghi, R.J. Donnelly, W.F. Vinen: J. Low Temp. Phys. 52, 189 (1983)
10. R. J. Donnelly, C. F. Barenghi: [Link]. Chem. Data 27, 1217 (1998).
11. L. Skrbek, S.R. Stalp: Phys. Fluids 12, 1997 (2000)
12. S.R. Stalp,L. Skrbek, R.J. Donnelly: Phys. Rev. Lett. 82, 4831(1999)
Intermittent Switching Between Turbulent
and Potential Flow Around a Sphere
in He II at mK Temperatures
1 Experiment
The experimental setup used in our investigations consists of a ferromagnetic mi-
crosphere (radius r = 124 μ m, m = 27 μ g) suspended between two niobium elec-
trodes by superconducting levitation. While cooling the horizontally arranged
electrodes forming a parallel plate capacitor (distance d = 1 mm, diameter 2 mm)
a dc-voltage of several hundred volts is applied to the capacitor giving rise to
an electric charge at the surface of the electrodes and the sphere. As the elec-
trodes become superconducting, the magnetic sphere is repelled and levitates at
an equilibrium position between the two electrodes. Horizontal stability is pro-
vided by trapped flux in the electrodes. As the sphere carries an electric charge
q vertical oscillations (f ≈ 150 Hz) can be excited by applying a resonant ac
electric field to the capacitor corresponding to a force F = q · Uac /d on the
sphere. The oscillations can be detected by measuring the current I = v · q/d
induced in the electrodes of the capacitor by the moving charge. The space be-
tween the superconducting electrodes is filled with pure 4 He (3 He concentration
≤1 ppb). This setup provides a simple geometry without disturbance by mechan-
ical suspension elements and a low background dissipation (Q-factors above 106
are achieved when the cell is evacuated). We measure the velocity amplitude of
the oscillating sphere for different driving forces at different temperatures. For
a more detailed description of the experimental technique and its applications,
see [1–4].
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 87–94, 2001.
c Springer-Verlag Berlin Heidelberg 2001
88 M. Niemetz, H. Kerscher, and W. Schoepe
2 8 3 5
3 0
2 5
2 6 2 0
s ta b le 1 5
la m in a r 1 0
2 4 flo w 5
0
v (m m /s )
in te r m itte n t 0 2 0 0 4 0 0 6 0 0 8 0 0
s w itc h in g
2 2
v c 2
2 0
v c 1
F c
s ta b le tu r b u le n c e
1 8
0 5 0 1 0 0 1 5 0
T = 3 0 0 m K F (p N )
Fig. 1. Three regimes of flow around the sphere: For low driving forces there is stable
laminar flow (•), resulting in a linear dependence v(F ) = F/λ. For large driving forces
there is stable turbulent flow () resulting in large dissipation and low oscillation
amplitudes. (The solid line is a fit of the turbulent v(F ) dependence derived in the text.)
Between these two regions there is a regime where both flow patterns are unstable and
which extends from vc1 over an interval Fc of driving forces to vc2 where turbulence
becomes stable. In this regime the system switches between laminar and turbulent
flow intermittently. The inset shows a larger range of driving forces and velocities. The
enlarged region is marked by a box
2 Results
The inset of Fig. 1 shows a typical result of the experiment. For small driving
forces the flow around the sphere remains laminar corresponding to a linear drag
force λv, which is given by ballistic scattering of phonons (left regime in Fig. 1).
For large driving forces there is stable turbulent flow instead, accompanied by
large nonlinear dissipation. Between these two regimes we find an interval of
driving forces, where neither laminar nor turbulent flow is stable but where
instead the fluid switches between these states intermittently.
T /2
T /2
1 0 0
Fig. 2. Turbulent flow around
8 0 the sphere: Data points taken
up to one hundred times larger
v (m m /s )
to find the corresponding F (v) dependence F (v) = 8γv 2 /3π. In Fig. 2 a larger
range of driving forces is shown, providing a more complete view of the turbulent
regime. The solid line is a fit, using a similar drag force as in (1) but reduced by
a constant value:
|Fd | = γ(v 2 − v02 ) = γv 2 − F0 . (3)
By applying the energy balance (2) this drag force corresponds to
8γ 3 2
F (v) = v − v0 = γ (v 2 − vc1
2 2
) (4)
3π 2
which perfectly fits the data points obtained in the superfluid (Fig. 2). A similar
drag force has been calculated for turbulent flow of a two dimensional dilute
Bose-Einstein condensate around a cylinder [5]. Even if these calculations are
not directly applicable to liquid helium, there might be a similar mechanism in
both cases for the reduced drag force. In our case turbulent drag exists only
above a critical velocity vc1 = 19.4 mm/s (s. Fig. 1).
3 0 a ) 4 7 p N
2 8
2 6
2 4
2 2
2 0
1 8
3 0 b ) 5 5 p N
2 8
v (m m /s )
2 6
2 4
2 2
2 0
1 8
3 0 c ) 7 5 p N
2 8
2 6
2 4
2 2
2 0
1 8
tim e ( 1 0 0 s / d iv is io n )
Fig. 3. Random sections of 800 s out of a 4 h time series of the oscillation velocity
amplitude for three different drives at 300 mK. Series (a) was taken at the lowest
driving force, showing the velocity increase during laminar phases and its exponential
saturation at a level given by phonon drag and driving force. Laminar phases are
interrupted intermittently by turbulent phases accompanied by a sharp drop of the
oscillation velocity amplitude to an equilibrium value given by the large turbulent drag
and driving force. Series (b) was taken using a larger driving force, resulting in a higher
saturation value (that is not reached any more) and slightly longer turbulent phases.
At the largest drive (series (c)) there is turbulent flow most of the time, interrupted
by laminar phases
the laminar phases are shorter than in series a. Taking a closer look at the life-
times of the turbulent phases shows that they in turn are longer than in series
a. Increasing the driving force further results in turbulent flow most of the time,
interrupted by short laminar phases. As the switching between both flow pat-
terns is intermittent, the results are discussed in terms of a statistical analysis
of the velocity amplitudes reached during laminar phases and their lifetimes as
well as the lifetimes of the turbulent phases. We apply reliability theory [6] for
the evaluation of the time series.
1 0 0 0
a ) b )
1 0 0
n u m b e r o f tu r b u le n t p h a s e s
1 0 0
μ (s )
1 0 1 0 3 2 m K
1 0 0 m K
T = 3 0 0 m K 2 0 0 m K
F = 5 9 p N 3 0 0 m K
4 0 3 m K
1
1
0 1 0 2 0 3 0 4 0 5 0 2 0 3 0 4 0 5 0
t (s ) F -λ v t (p N )
Fig. 4. Analysis of the lifetimes of the turbulent phases. (a) The number of turbu-
lent phases living longer than a certain time t shows an exponential decay (solid line)
corresponding to exponentially distributed lifetimes. The reciprocal slope of the distri-
bution in the plot gives the mean lifetime μ. (b) Mean lifetimes of turbulent phases for
different driving forces and temperatures. There is no temperature dependence, but a
strong dependence on the driving force. The solid line is a fit of a fourth-power law
divergence of the mean lifetime μ at a critical driving force value
in a wide range of driving forces and have always found exponentially distributed
lifetimes. A comparison of the mean lifetimes obtained is shown in Fig. 4b. The
values of μ are independent of temperature and collapse to an universal drive
dependence if the strongly temperature dependent laminar drag λv (where v = vt
is the velocity amplitude of the turbulent phase) is subtracted from the external
driving force. The mean lifetime of turbulent phases increases with the driving
force and diverges at a critical value Fc = 54 pN approximately with a fourth-
power law and stays infinite at larger drives corresponding to turbulent velocities
above vc2 , see Fig. 1. The power dissipated at Fc is 0.6 pW, corresponding to the
production of ≈ 1 mm vortex lines per half period. This is equivalent to producing
a vortex ring with a diameter of ≈ 1.4 times the diameter of the sphere.
7
2 8 m K
a ) b ) 1 0 0 m K 2 0 0 m K
3 0 0 m K 4 0 3 m K
1 0 0 6
n u m b e r o f la m in a r p h a s e s
+ 1 0 %
(m m /s )
5
4 .7 9
1 0
w
v
-1 0 %
4
T = 3 0 0 m K
F = 5 5 p N
1 3
0 2 4 6 8 1 0 1 2 0 1 0 2 0 3 0 4 0 5 0
Δ v (m m /s ) F -λ v t (p N )
Fig. 5. Analysis of velocity amplitudes reached during laminar phases. (a) The number
of laminar phases having an amplitude larger than Δ v shows a quadratic dependence
on Δ v in this semilogarithmic plot (solid line), corresponding to a Weibull probability
density of velocities. (b) Fitting parameter vw for different driving forces and temper-
atures. We find no systematic dependence on these parameters. The values vary in a
10% bandwidth around a mean value of 4.8 mm/s
of breakdown after the laminar phase has survived for a time t) is given by [6]
2
d d Δ v(t) 2
Λ(t) = − ln P (Δ v(t)) = = 2 Δ v Δ v̇ . (5)
dt dt vw vw
Because the velocity amplitude is exponentially approaching an equilibrium
value with a time constant τ = 2m/λ, this result has several implications: First,
as Δ v̇ is constant for small values of Δ v (or t τ ), the failure rate is increasing
with the velocity amplitude of the sphere. Second, with the velocity reaching the
equilibrium value (or t τ ), Δ v̇ goes to zero, and so does Λ(t). This means that
the laminar phase, having survived for many τ and having reached its stationary
velocity amplitude, will live forever although the velocity is clearly above vc1 . Ex-
perimentally, however, this is not exactly true. This can be seen in Fig. 6, where
the probability distribution for large lifetimes is shown. The decrease slows con-
siderably down for lifetimes greater than approximately 100 s (corresponding to
3τ at 300 mK), but the distribution does not approach a constant value. Instead,
a constant failure rate is found, leading to a mean lifetime of 25 min (indicated by
the straight line in Fig. 6). Obviously, there must be another mechanism causing
the breakdown of those long lived laminar phases. We can exclude mechanical
vibrations or acoustic noise to be the origin, as we tried to destroy such long
laminar phases by slamming the door, jumping on the floor, or even refilling
helium into the cryostat. But placing a small radioactive source (60 Co, 74 kBq)
outside the dewar, had a dramatic effect on the lifetimes. As can be seen from
Fig. 6, the mean lifetime changes by a factor 8.3 from 25 minutes to 3.0 minutes.
We have measured the dose rate of the source at the position of the measuring
cell inside the cryostat (taking into account a measured 20% loss in the dewar
Intermittent Switching Between Turbulent and Potential Flow 93
walls) to be 440 nGy/h (± 5%). Comparing this value with a measured dose rate
due to natural background radiation in our laboratory of 50 nGy/h (± 10%),
which is typical for our area, we obtain an increase of the dose rate due to the
source by a factor of (440 + 50)/50 = 9.8 This compares well with the ratio of
lifetimes of metastable laminar phases obatined in the experiment. Therefore, it
is obvious that natural background radioactivity limits the lifetime of metastable
laminar phases above the critical velocity vc1 . This effect may be attributed to
local vorticity generated by ions produced by radiation, inducing the breakdown
of the metastable laminar flow around the sphere.
3 Conclusion
The turbulent flow of the superfluid causes a drag force on the sphere which
is very similar to turbulence in classical fluids except for a constant offset in
analogy with a dilute Bose-Einstein condensate. Furthermore, we have found
that the transition from potential flow to turbulent flow around a sphere in
superfluid 4 He at mK temperatures occurs by intermittent switching between
both flow patterns instead of the gradual transition observed in viscous fluids. A
statistical analysis of this switching phenomenon has shown that the lifetimes of
turbulent phases diverge at a critical driving force. The probablity for breakdown
of the laminar phases has been obtained. Finally, there exist metastable laminar
phases above the critical velocity vc1 . Their lifetime is limited only by natural
background radioactivity [7].
We have repeated our experiment several times by heating the measuring cell
above Tc of niobium in order to prepare a new levitating state of the sphere. All
our observations were reproducible. The three quantities vc1 , vw and Fc which
we expect to be affected by the properties of the sphere have standard deviations
of 8%, 13% and 15%, respectively, probably due to asymmetries of the surface
of the sphere.
We understand now why above 0.5 K the intermittent switching changes into
the hysteretic behavior observed earlier [1,2]: the velocity increases Δ v become
very small at higher phonon drag which implies a very low failure rate of the
laminar phase. But if it fails (i.e. when the driving force F is largely increased)
the following turbulent phase is stable because the critical drive is exceeded.
With our present results in superfluid 4 He it appears extremely promising to
extend these experiments to superfluid 3 He in order to investigate the transition
to turbulence in this very different quantum fluid.
References
1. J. Jäger, B. Schuderer, and W. Schoepe: Physical Review Letters 74, 566 (1995).
2. J. Jäger, B. Schuderer, and W. Schoepe: Physica B 210, 201 (1995)
3. P. Eizinger, W. Schoepe, K. Gloos, J.T. Simola, and J.T. Tuoriniemi: Physica B
178, 340 (1992)
4. M. Niemetz, W. Schoepe, J.T. Simola, and J.T. Tuoriniemi: Physica B 280, 559
(2000)
5. T. Winiecki, J.F. McCann, and C.S. Adams: Physical Review Letters 82, 5186
(1999)
6. B.V. Gnedenko, Yu.K. Belayev, and A.D. Solovyev: Mathematical Methods of
Reliability Theory (Academic Press, New York, 1969)
7. M. Niemetz, H. Kerscher, and W. Schoepe: [Link]
mat/0009299 and to be published
Vortex Filament Methods for Superfluids
David C. Samuels
1 Introduction
Vortex filaments are an idealized form of rotational flow where the vorticity
is confined to a small core region, of radius a, around a one dimensional line
embedded in the three dimensional flow. Outside of this core region the flow is
potential. When the dynamics of the core size are not important these objects
may also be referred to as vortex lines.
In classical fluid mechanics, by which I mean solutions of the Navier-Stokes
or Euler equations, vortex filaments are a useful tool for understanding the
geometry and dynamics of a flow. But after an initial popularity in the early
1980’s [1] [2] [3] [4] [5], the use of vortex filament methods fell out of favor
in classical fluid mechanics. Though there has been some slight resurgence in
this method recently [6] [7] due to the rapidly increasing computational power
available and the development of new computational algorithms, direct numerical
simulations and large eddy simulations have become the methods of choice for
calculating the motion of fluids. One reason for the decreased use of vortex
filament methods is that while they give a clear and intuitive understanding of
a flow through the easy visualization of the vortex filaments, this representation
was often just a rough cartoon of the true flow. Vortex filaments are only a
convenient idealization in a classical flow. The vorticity in a realistic classical
flow rarely takes the form of clearly discrete vorticity filaments.
But in superfluids like helium II vortex filaments are real [8]. Due to the
quantization of circulation, vorticity in a helium II flow can only exist within
vortex filaments with a core size of a. Since this core size is very small in helium
II, about 1 Angstrom, the thin-core vortex filament idealization is actually a
very accurate description of the true superfluid flow.
There are a few special qualities of the superfluid vortex filament [8] that
make it even simpler than the standard model of a vortex filament in a classi-
cal fluid. The circulation around each superfluid vortex filament in helium II is
set by quantum mechanics to be an integral multiple of κ = 9.97x10−4 cm2 /sec.
Since an n quantum vortex filament contains more energy than n single quan-
tum vortices, it is generally assumed that only single quantum vortex filaments
will be commonly observed. Thus all helium II superfluid vortices have identical
circulation. The core size a is also determined by quantum mechanics and is
closely related to the concept of a healing length, the length scale required for a
wave function to change from its bulk value to zero. In the classical fluid vortex
filament the core size is an important variable and much of the complications,
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 97–113, 2001.
c Springer-Verlag Berlin Heidelberg 2001
98 D.C. Samuels
and the interesting dynamics, of classical vortex filaments are due to changes in
the core size along a filament [9]. Without a variable core size, the dynamics of
a quantized vortex filament are much simpler than those of its classical counter-
part. Of particular importance is the effect of the constant core size on vortex
stretching. When a classical vortex filament is stretched, the core size shrinks
and the local vorticity rises along the filament. When a quantum vortex filament
is stretched, it can only elongate with no change in core size. Due to this dif-
ference, many prefer to reserve the phrase “vortex stretching” for the classical
process, and will say that vortex stretching does not occur for individual quan-
tized vortex lines. One should note, however, that a bundle of parallel quantized
vortex filaments can undergo vortex stretching, when the core size is interpreted
as the radial size of the bundle of filaments.
Many of the complications of classical vortex filament calculations have to
do with vortex stretching and with other core dynamics which are assumed
negligible in helium II quantized vortex filaments. We thus have a much simpler
system to deal with, both in terms of physical behaviour and in terms of the
computation of the motion of the filaments. My discussion in this paper will be
concerned directly with the superfluid vortex filaments of helium II, and when
necessary I will take the values of constants to be those of helium II.
The first difficulties in the vortex filament method come from the definition of
the term V I in the equation of motion. This term represents the advection of
the point on the vortex filament by the velocity field due to all the superfluid
vortex filaments present in the flow. This advection velocity at the point r is
given by the Biot-Savart law.
κ (S − r) ⊗ dS
V I (r) = (2)
4π |S − r|3
The line integral is taken over the entire length of vortex filament present in
the fluid. The Biot-Savart law can be used to find the flow at any point r in
the fluid, but the immediate difficulty with this equation comes from its use
in calculating the advection velocity in (1). In this case, the point r lies on
the vortex line and the Biot-Savart law contains a singularity at that point.
To heal this singularity we must include some aspect of the core structure of
the vortex filament. Many different methods of de-singularizing the Biot-Savart
law have been used in classical fluid dynamics [10] [11] [12] [13], where the core
dynamics of the filaments can often be quite complicated and important, but in
superfluid vortex methods I have only seen one method used [14]. This method
breaks the line integral into two parts; an integral over a local neighborhood
around the point r (this contains the singularity), and the rest of the Biot-Savart
integral (this part is non-singular). Then the local integral must be replaced by
an algebraic approximation that takes into account the core size of the filament.
100 D.C. Samuels
S is the unit tangent vector to the vortex line at the J’th mesh point and S ,
the second derivative of the line with respect to arclength, is the local curvature
vector. The length scales l+ and l− are distances between the J’th mesh point
and the J + 1 and J − 1 points respectively. The parameter aef f is an effective
core size since numerical constants of order one have been absorbed into this
parameter. This choice for the local part of the Biot-Savart integral gives the
correct velocity for a planar vortex ring. Defining the velocity of a planar vortex
ring requires that you specify some details about the core structure of the vortex
(hollow core vs solid body rotation, for example). Different choices of the core
structure will alter the log term in (3), but unless you are calculating flows on
very small length scale, on the order of the core size, then these details make
negligible changes to the value of (3).
Once you prevent the singularity by splitting the Biot-Savart law into two
sections you are still left with the difficulty of calculating the integral over the
rest of the vortex line. In a simulation, we only know the positions of the discrete
mesh points along the line, so the mesh points must be joined in some numerically
convenienent manner to calculate the integral between the mesh points. Any
interpolation method through the mesh points will do the job. I prefer to use
the simplest method of integrating over piece-wise linear vortex line segments
between the mesh points. The velocity at a point r (which may be a mesh point
on the vortex line or may be any point in the volume of the fluid) due to the
Biot-Savart integral over the segment of vortex line between the J and J + 1
mesh points is
SJ+1
κ (S − r) ⊗ dS
V segment (r) = . (4)
4π SJ |S − r|3
Taking this integral over a straight line between the two mesh points gives
One of the reasons that vortex filament methods have fallen out of favor in
classical fluid dynamics is the fact that the calculation of the Biot-Savart integral
requires order N 2 operations over a mesh of N points. The calculation of this
integral is computational expensive whenever a large number of mesh points are
needed to represent the vortex lines. This calculation is often not practical, even
with the speed of today’s computers, if the amount of vortex line is large, or if
the fine detail of the vortex line shape needs to be captured using a very fine
meshing. And these are often exactly the cases that are of the most interest to
us.
One way around this difficulty is the Local Induction Approximation, (LIA)
[15]. In this method we keep the local term in the induced motion (3) and neglect
completely the non-local Biot-Savart integral. This is typically done with some
minor adjustments to the log term so that we have
κ 2R
V I,LIA = (S ⊗ S ) ln , (6)
4π aef f
where R is a length scale of the filament. R may be taken as a constant, such as
the length scale of the computational box, or it may be taken as the local radius
of curvature of the vortex line. In the first case, the log term is a constant, and
it is often absorbed into a non-dimensional timescale. In the second case, the log
term will vary along the filament and with time, but unless you are capturing a
very wide range of curvatures with your mesh, or the curvatures are only about
a factor of 10 larger than the effective core size, the log term is nearly constant.
LIA is a very convenient approximation. It is simple and easy to calculate.
The time required to calculate (6) increases only linearly with the number of
mesh points. The interpretation of (6) is simple: the induced velocity is in the
direction of the local binormal S ⊗ S and is inversely proportional to the
local radius of curvature. It is commonly used in analytic investigations of the
properties of vortex filaments. And LIA correctly describes the motion calculated
by the Biot-Savart law of simple vortex line geometries such as the planar vortex
ring and low amplitude vortex waves. But it is a very severe approximation to
the true equation of motion!
LIA clearly works very well for calculating the motion of single vortex lines
which do not loop around so that sections of the filament far apart in arclength
are actually close to each other. Also it probably works well in a random vortex
tangle, where the non-local part of the Biot-Savart law may tend to cancel out
when integrated over the random vortex lines. LIA will not work well for vortex
configurations which tend to have parallel, or antiparallel, sections of vortex
filament near one another (for example, vortex knots [16]). LIA will not work well
for flows which tend to develop any alignment of vorticity. Unfortunately, aligned
superfluid vorticity does tend to be formed by flows at non-zero temperatures
where the normal fluid has some local vorticity structure (the superfluid vortices
tend to align with the normal fluid vorticity [17]). In the case of turbulent flows, I
personally doubt that the LIA captures enough of the physics of the interactions
between the vortex filaments to give much useful information on these flows,
though others would debate this point.
102 D.C. Samuels
Some of the qualitative differences between the motion of vortices under LIA
and the full Biot-Savart law should be pointed out. A superfluid flow calculated
by LIA can never develop any rotational flow, aside from the rotation around a
single filament. In an LIA calculation, the energy of the flow is just an energy
per unit length of the vortex lines and when there is no mutual friction (at zero
temperature) the length of vortex line does not change with time (without further
assumptions to the model, such as phonon emission). Under the Biot-Savart law,
the kinetic energy in the flow can be quite difficult to calculate, and even at zero
temperature the length of vortex line can change while the kinetic energy remains
constant. In a LIA calculation, the physical effect of vortex stretching cannot
occur while it can occur in a Biot-Savart calculation whenever there are even
just two approximately parallel vortex lines.
With the vortex filaments represented by N mesh points, the vortex equation
of motion (1) becomes N coupled, first order ordinary differential equations. This
coupled set of ODEs can be solved by any general method. I prefer to use a fifth
order Runge-Kutta method with an adaptive stepsize. An adaptive method like
this calculates the result of one time step in two different ways (one fourth order
method and one fifth order method) and compares the results at each of the N
mesh points to estimate the error. An allowed error range is defined and if the
estimated error lies below this allowed error range for all N points, the time step
is increased, while if the estimated error is above the allowed range for any of the
N mesh points, the time step is decreased and that timestep is recalculated. This
is a sturdy method, capable of automatically handling the rapid changes in vortex
line velocities which can occur, often as a result of vortex line reconnections. It is
not a very efficient method however. Whenever any region of very high curvature
(or a close approach of two vortex filaments under the Biot-Savart law) occurs
then the timestep for the entire vortex tangle can drop significantly. Some type of
‘multi time step’ method, where the motion of fast moving sections of the vortex
line could be calculated using more time steps than the slow moving sections of
the vortex, would would be far more efficient, particularly when the equation of
motion is as expensive to evaluate as is the Biot-Savart law.
such as the vortex line density. Periodic boundary conditions are easily pro-
grammed for LIA simulations, though care must be taken at the places where
neighboring mesh points along a filament lie on opposite sides of your periodic
boundary. But these problems are easily handled. The main difficulty with peri-
odic boundary conditions comes from the Biot-Savart calculations. In principle,
periodic BCs make the Biot-Savart integral infinitely long. In practice, this inte-
gral must be cut off at some arbitrary point. The costs of periodic BC in Biot-
Savart calculations can be staggering. For a three dimensional periodic cube,
including the first layer of periodic vortices around the central cubic volume will
increase the number of vortex mesh points by a factor of 27. Including even this
level of periodicity is usually prohibitively costly. To deal with problem I will
sometimes define a buffer layer around the central cubic volumn, with a width
of 1/2 to 1/4 of the width of the computational volume, and I will only include
in the Biot-Savart calculation the periodic vortex filaments that fall within this
fairly thin buffer zone. This type of method actually only ensures that vortex
filements moving through one side of the periodic volume and re-entering on the
other side, experience a more smoothly varying velocity field during the transi-
tion. Ideally, this type of problem could be removed by an analytic method to
calculate the velocity field due to this infinitely repeating vorticity distribution.
While this has been done in two dimensions, I know of no such method in 3D.
In LIA, solid boundary conditions are quite easy to implement [14]. In this
case, for any vortex filaments which end on the solid surface we must set the
normal vector at the ends of the vortex filament equal to the local normal vector
of the solid surface. This is the only condition needed no matter how complex the
boundary geometry. In Biot-Savart calculations we must use image vortices for all
the vortex filaments in the flow. For a single flat boundary, these image vortices
are simply vortices in the mirror image positions, and with their orientation
reversed. For a single flat surface this doubles the length of the Biot-Savart
calculations, which is not a terribly onerous increase. But for multiple solid
surfaces, you must then include the images of images, ad infinitum (just as
you must do in electrostatic calculations). This naturally introduces the same
computational difficulties as in the periodic BC case, and it must be solved in
the same manner, either by cutting off the image calculations at some arbitrary
number of images or by developing an analytic summation method to collapse
the infinite images to a reasonable calculation. I know of no one who has done
the latter for a 3D vortex flow, though the tools may exist in the electrostatics
literature for us to use.
from other sections of the filament as they move too close to one another (close
in terms of arclength). The criteria that you choose for remeshing depends on
the behaviour that you are trying to capture with a specific simulation.
For this discussion, let the arclength distance between two neighboring mesh
points, J and J + 1 be represented by δJ,J+1 . The simplest remeshing criteria is
to attempt to keep an approximately uniform distance between the mesh points.
In this method, the mesh point J would be removed if δJ−1,J + δJ,J+1 < δ and a
mesh point would be added between the J and J + 1 points if δJ,J+1 > 2δ. This
will keep all the distances along the mesh roughly in the range δ/2 to 2δ. This
is the most straightforward meshing method, but it suffers from the limitation
that it can only resolve structures on the vortex filament of scale δ or larger.
This method is best if you are only interested in the vortex behaviour at one
particular length scale.
If you want to calculate the development of vortex filament structures over
a wide range of length scales then you must use a more complicated remeshing
method. We would like to have more mesh points in the parts of the vortex
filament with high curvature and fewer in regions of low curvature. This can be
done by interpreting the meshing length scale δ as a variable, proportional to the
local radius of curvature. Typically, I will take δ = 2πR/32, where R is the local
radius of curvature of the filament. This will mesh a planar ring, at any length
scale, with approxamately 32 mesh points. Of course, there must be a limit on
the range of this meshing. This is done by setting a range of R over which this
remeshing will be used, with minimum and maximum mesh lengths to be used
outside this range. With this method it is possible for the mesh to represent a
range of two or three orders of magnitude in radius of curvature without having
an unreasonably large number of mesh points. It may be possible to extend
this range by another order of magnitude by setting the meshing length δ to
decrease slower than linearly with R. Techniques like this are used in 2D vortex
simulations in the meshing of the surface of vorticity blobs [18].
However you choose to remesh, removing mesh points poses no problems but
in adding mesh points you must make a choice on the position of the new mesh
point. Once again, the simplest choice is to use a piece-wise linear interpolation,
and simply place the new point at the center of the straight line between the J
and J + 1 point. But if you do this, then you are likely to change the first and
second derivatives, S and S at the points J and J +1. Worse than this, the new
point would be introduced with zero local curvature, almost certainly causing a
rapid (and artificial) change in curvature along the filament. For adding mesh
points a better interpolation must be used, and almost any interpolation will
perform well as long as it is not piece-wise linear. I use a method that inserts
new mesh points at a position where the curvature vector, S , at the new mesh
point is the average of the curvature vectors at the two neighboring points, J
and J + 1. This prevents the intruduction of sudden jumps in the vortex line
curvature, errors which would radiate vortex waves and affect more than just
the local segment of the vortex line.
Vortex Filament Methods for Superfluids 105
For ease of explanation in this paper I have been labelling neighboring mesh
points with consecutive number, J and J + 1 for example. Any remeshing natu-
rally will destroy this order, and consecutive numbering is not practical. Instead,
a data structure should be used that has for each mesh point, the labels of the
two neighboring mesh points. Then a remeshing involves only the addition or
deletion of a mesh point and the necessary changes in this data structure for the
neighboring mesh points.
One further remeshing routine should be used. Occasionally, very small vortex
rings may develop. These may be formed by the decay through mutual friction
of larger vortex loops. Or they may be generated by the pinching off of a small
loop in a reconnection event. However they are formed, these small loops should
be removed from the calculation. By definition, the curvature of these loops will
be very large, and thus their motion will be very fast. With an adaptive stepsize
technique, the code will automatically drop the time step to a very small value
to attempt to accurately calculate the rapid motion of these small loops. If you
are not using an adaptive stepsize, then the motion of these small loops will not
be accurately calculated, possibly leading to the sudden and artificial expansion
of the small loop. In either method these small loops must be removed from the
calculation. It is a good idea to keep a record at least of the length of vortex line
removed from the simulation by this routine.
3 Reconnections of Filaments
When two classical Navier-Stokes vortex filaments cross, viscous effects will re-
connect the vortex filaments. Lacking these viscous effects, the crossing of two
classical Euler vortex filaments is believed to lead to a singularity and the break-
down of the Euler equations [19]. This singularity formation is not observed in
calculations of the crossing of superfluid vortex filaments, where vortex recon-
nection occurs without the need for viscosity [20]. In this way, superfluid vortices
are clearly not Euler vortex filaments. The difference lies in the core structure
of the superfluid vortices. At the centerline of each vortex filament, quantum
mechanics requires the density of the superfluid (given by the amplitude of the
ground state wave function) to go to zero and the vorticity of the flow is singular
along this centerline.
Studies of superfluid vortex filament reconnections must be done with a quan-
tum theory calculating the evolution of the wavefunction of the superfluid ground
state. A more detailed discussion of these calculations can be found in the articles
by Roberts and by Adams in this volume. These studies have traditionally been
done using a Non-Linear Schrodinger Equation (NLSE). It is recognized that
the NLSE is probably a poor representation of helium II superfluid because the
dispersion curve calculated from the NLSE contains only phonons and no rotons.
The NLSE does, however, seem to be a very good representation of the new Bose
condensed alkali atom superfluids. Without a fundamental microscopic theory
of helium II superfluidity, we must consider the results of NLSE calculations,
keeping in mind the possible errors of these calculations. The NLSE calculations
106 D.C. Samuels
clearly show that when two superfluid vortex lines cross, a reconnection event
occurs and the topology of the vortex lines changes [20]. This behavior must be
included as a basic assumption in the superfluid vortex filament model since it
is not a consequence of the equation of motion. This discussion will be made in
terms of the crossing to vortex filaments in the bulk of the fluid. In simulations
with boundaries, reconnections of filaments with the boundaries must also be
considered, but these reconnections may be easily detected as vortex segments
approach and eventually hit the boundaries so we need not discuss them further
here.
The reconnection assumption leaves us with the problem of detecting the
crossings of 1-D lines moving through 3-D space, when we only know the posi-
tions of a finite number of mesh points along these lines at discrete time intervals.
Though the assumption can be quite simply stated, it has proved very difficult
to develop any satisfactory algorithms to implement this. The first reconnec-
tion algorithm, due to Schwarz [21], was a simple, intuitive one: vortex lines are
reconnected whenever two mesh points come within a pre-defined distance Δ
of each other. While this algorithm is very simple to use, the objections to it
are many and strong. This algorithm will obviously trigger a reconnection event
early, since a true reconnection event does not occur until the distance between
the lines is on the order of a core radius (practically zero in filament simulations).
Sometimes, a large section of vortex filament, including many mesh points, will
move within the reconnection distance triggering a series of reconnection events
resulting in the formation of a number of small vortex loops where only one
crossing event would actually have occured. Even worse, in some cases this al-
gorithm will trigger a reconnection where no crossing of the vortex lines would
actually occur. This error happens when vortex lines pass by each other, without
intersection but within the distance Δ. This may at first seem to be a rare and
thus negligible event, but it is not. We now know that vorticity concentrations
in the normal fluid generate bundles of well-aligned superfluid vortex filaments
at the center of the normal fluid rotation [17]. These superfluid vortex filaments
may be closer to each other than the defined reconnection distance Δ but they
cross only rarely. With the Δ reconnection algorithm, these vorticity bundles
will reconnect wildly.
The Δ reconnection model also introduces a new and completely artificial
length scale to the simulation; the reconnection length scale Δ. If you set Δ too
large, you will have many spurious reconnections. If you set it too low, you will
miss many real reconnections, as vortex lines jump past one another in the finite
time step, plus you will still have some spurious reconnections of nearly aligned
vortices. Some of the problems of this algorithm can be helped (though not
eliminated) by making the reconnection length scale variable along the vortex
filament and setting it equal to the local mesh size (a natural choice). This
refinement to the algorithm still suffers from the separate problems of premature
and also spurious reconnection.
A different reconnection detection algorithm, the one that I currently use,
is based on the idea that a vortex filament reconnection event is a dissipative
Vortex Filament Methods for Superfluids 107
the fundamental averaged quantity of a vortex tangle and many of the analytic
model of superfluid vortex turbulence have been based on this quantity alone.
These models are differential equations for the evolution of the line density as a
function of the average normal fluid velocity and some temperature dependent
parameters. The most basic of these models is the Vinen equation [27]
dL χ2 κ 2
= χ1 αVns L3/2 − L , (7)
dt 2π
where χ1 and χ2 are temperature dependent parameters and Vns is the average
relative velocity between the normal fluid and the superfluid. The standard in-
terpretation of the Vinen equation is that the first term represents the growth of
L due to mutual friction and the second term represents the decay of the tangle
due to reconnection events, though other interpretations of these terms are have
been made [28].
Now let us define some measure of the isotropy of the tangle. Very little work
has been done on measures of isotropy in vortex tangles. One set of measures
defined by Schwarz [28] defines three quantities which measure the isotropy of a
tangle relative to two perpendicular unit vectors, r̂ and r̂⊥ .
1
I = [1 − (S · r̂ )2 ]dξ (8)
VL
1
I⊥ = [1 − (S · r̂⊥ )2 ]dξ (9)
VL
1
I = (S ⊗ S ) · r̂ dξ (10)
V L3/2
In flows with a unidirectional normal fluid flow the unit vectors r̂ and r̂⊥ are
usually taken to be parallel and perpendicular to the direction of the normal
fluid flow. The interpretation of these three isotropy measures is discussed in
[28]. Other isotropy measures can be defined, such as the length of line in a
projection along a vector r̂
1
J(r̂) = 1 − (S · r̂)2 dξ. (11)
VL
It is possible that more refined versions of (7) could be defined using these
isotropy measures as well as L.
The local rate of extension (or compression if negative) of the vortex line is
defined by
r(ξ) = −S · VL (12)
where VL is the vortex line velocity at the arclength position ξ. Integrating
r(ξ) over the arclength gives the instantaneous rate of change of the vortex line
density.
dL 1
= rdξ (13)
dt V
Vortex Filament Methods for Superfluids 109
Simulations show that even in a steady state vortex tangle the line length density
has strong fluctuations, so one must be careful in interpreting instantaneous
measures such as (13).
It is tempting to define a whole range of quantities averaged over the vortex
tangle. One that must be mentioned is the averaged curvature
1
C= |S |dξ, (14)
where = dξ is the total vortex filament length. This quantity can be used
to define an average length scale of the tangle, 1/C, and an average velocity
scale κC. Following the evolution of this quantity with time can detect if the
vortices are ‘crinkling’ (increasing C) or smoothing (decreasing C). But I urge
against an over-reliance on averaged quantaties. It is easy enough to measure the
full distribution of values of quantities such as the curvature in the tangle, thus
measuring the full range of behavior of the vortices and not just some average
value. Again, little has been done with this approach.
But what about the more traditional fluid mechanics measures such as kinetic
energy and linear and angular impulse? While these are quite easily determined
in a simulation where the velocity is defined on a grid extending throughout the
fluid, they are actually quite difficult to measure in a vortex filament simulation.
Purely for diagnostic purposes, one can define a 3D grid and calculating the
velocity on the grid by the Biot-Savart law (2) and then such quantities can be
easily measured. But it is important when doing this to check the convergence
of your results with grid resolution. The 1/r velocity field of the vortex lines can
make resolution of the energy and momentum difficult. It is possible to avoid
this problem by using definitions of these quantities in terms of the line integrals
over the vortex filaments [29]. Some useful integrals are: kinetic energy
T = κρs V s · (S ⊗ dS), (15)
linear impulse
1
I= κρs S ⊗ dS, (16)
2
angular impulse
1
A = κρs S ⊗ S ⊗ dS, (17)
3
and helicity
J =κ V s · dS. (18)
These integrals are limited to flows with no vorticity at infinity, and the integrals
must be taken over closed vortex loops.
While simple things such as kinetic energy can become quite difficult to mea-
sure in vortex filament methods, there are some qualities of the flow which are
simpler to define and measure in filament methods. There has been much re-
cent interest in the topology of vorticity in Euler flows, interest which is mainly
110 D.C. Samuels
5 Alternative Approaches
The computational cost of calculating the full Biot-Savart law is high, due to
its non-locality. The usefulness of local induction approximation calculations is
suspect because of its complete locality. Is there not a middle way that contains
some, but not all, of the non-locality of the Biot-Savart integral? The recent
development of fast multipole methods seems to fill this need [31] [6]. Originally
developed for simulations of large numbers of gravitationally interacting parti-
cles, these methods can lower the computational time of pairwise interactions
between particles from order N 2 to order N log(N ). In these methods the inter-
acting particles are grouped into clusters, and then into clusters of clusters, et
cetera, in a hierarchical tree structure. The interactions between the clusters are
calculated by a finite multipole series expansion, greatly decreasing the number
of calculations needed while retaining the primary non-local effects. The general-
ization from the original gravitational interaction to a vortex motion calculation
is straightforward and these new methods have fueled much of the recent resur-
gence (however slight) in vortex methods for classical fluid dynamics. These
methods work best in cases where the fluid vorticity has some strong spatial
structure. The drawback to these methods is the high overhead cost associated
with the determination of the hierarchical tree data structure, and the necessity
of rebuilding this data structure as the vortices move, and groups of vortices
break up or coalesce. For a moderate number of vortex mesh nodes, up to a few
thousand in my experience, the Biot-Savart calculations are still faster than the
fast multipole methods, though this limitation may become outdated soon as
new implementation algorithms are being developed. For simulations af actual
superfluid turbulence, where we must reproduce a wide range of spatial scales in
the superfluid vorticity (requiring a large number of mesh points) and include
the non-local vorticity interactions of turbulence (ruling out the LIA approach)
we must use some non-local approximate method in our simulations, and the
fast multipole methods are currently our best hope.
Vortex Filament Methods for Superfluids 111
Yet another approach is to drop the idea of vortex filaments completely and
use a locally averaged set of equations for the superfluid vorticity field, ωs . These
approaches use the Hall-Vinen-Bakharevich-Khalatnikov (HVBK) equations (or
modifications of these equations) for the coupled evolution of continuous vorticity
fields in the superfluid and the normal fluid. Since this approach does not use
vortex filaments it lies somewhat out of the scope of this article, but is dealt
with in the articles by Henderson and Holme in this volume.
phonon and roton emission from reconnecting vortices [36] and possibly even
from just the motion of vortices. Does this ‘zero temperature dissipation’ clearly
affect the properties of the vortex tangle? Is it enough to produce an inertial
range in the pure superfluid turbulence? Can we develop an understanding of
the energy cascade in a pure superfluid vortex tangle? I say that this is a not en-
tirely separate issue from that of the fully coupled helium II turbulence because
we need to understand the behaviour of the fully coupled turbulence in both the
high and low temperature limits. In the high temperature limit, helium turbu-
lence must surely be classical Navier-Stokes turbulence after the lambda tran-
sition temperature is passed and some experiments show no detectable change
in the measured turbulence quantities as this transition temperature is passed.
Pure superfluid turbulence is the low temperature limit of fully coupled helium
II turbulence and there is no reason not to expect this transition to also be
very smooth. Helium II turbulence experiments show little, if any, temperature
dependence down to approximately 1.4 Kelvin [37], where the normal fluid rel-
ative density and the mutual friction interaction both become quite small. An
understanding of both classical Navier-Stokes turbulence and pure superfluid
turbulence will be helpful in developing our understanding of coupled helium II
turbulence.
References
1. A. J. Chorin: SIAM J. Sci, Stat. Comput. 1, 1 (1980)
2. A. Leonard: J. Comput. Phys. 37 289, (1980)
3. B. Couet, O. Buneman, A. Leonard: J. Comput. Phys. 39 305, (1981)
4. A. Leonard: Ann. Rev. Fluid Mech. 17 523 (1985)
5. W. T. Ashurst, E. Meiburg: J. Fluid Mech. 189 87 (1988)
6. G-H. Cottet, P. D. Koumoutsakos: Vortex Methods (Cambridge University Press,
Cambridge, 2000)
7. R. Cortez: J. Comput. Phys. 160 385 (2000)
8. R. J. Donnelly: Quantized Vortices in Helium II (Cambridge University Press,
Cambridge, 1991)
9. V. M. Fernandez, N. J. Zabusky, V. M. Gryanik: J. Fluid. Mech. 299 289 (1995)
10. T. Y. Hou: Math. of Comp. 58 103 (1992)
11. R. Klein, O. Knio: J. Fluid Mech. 284 275 (1994)
12. M. F. Lough: Phys. Fluids 6 1745 (1994)
13. R. Klein, L. Ting: Appl. Math. Lett. 8 45 (1995)
14. K. W. Schwarz: Phys. Rev. B 31 5782 (1985)
15. R. L. Ricca: Fluid Dyn. Res. 18 245 (1996)
16. R. L. Ricca, D. C. Samuels, C. F. Barenghi: J. Fluid Mech. 391 29 (1999)
17. D. C. Samuels: Phys. Rev. B 47 1106 (1993)
18. D. G. Dritschel, N. J. Zabusky: Phys. of Fluids 8 1252 (1996)
19. R. B. Pelz: Phys. Rev. E 55 1617 (1997)
20. J. Koplick, H. Levine: Phys. Rev. Lett. 71 1375 (1993)
21. K. W. Schwarz: Phys. Rev. Lett. 49 283 (1982)
22. T. F. Buttke: Phys. Rev. Lett. 59 2117 (1987)
23. K. W. Schwarz: Phys. Rev. Lett. 59 2118 (1987)
Vortex Filament Methods for Superfluids 113
Darryl D. Holm
Theoretical Division and Center for Nonlinear Studies, Los Alamos National
Laboratory, MS B284, Los Alamos, NM 87545
1 HVBK Equations
Recent experiments establish the Hall-Vinen-Bekarevich-Khalatnikov (HVBK)
equations as a leading model for describing superfluid Helium turbulence. See
Nemirovskii and Fiszdon [1995] and Donnelly [1999] for authoritative reviews.
See Henderson and Barenghi [2000] for a recent fluid mechanics study of steady
cylindrical Couette flow using computer simulations of the incompressible HVBK
equations.
In the Galilean frame of the normal fluid with velocity vn , the HVBK equa-
tions may be expressed as follows, upon ignoring thermal diffusivity and viscosity,
∂t ρ = − divJ , ∂t Ji = − ∂k Tik , ∂t S = − div(Svn ) + R/T ,
1
ρs ∂t vs + ρs (vs · ∇)vs = − ρs ∇ μ − |vs − vn |2 + ρs f , (1)
2
and summing over pair of upper and lower repeated indices. One may consult,
e.g., Bekarevich and Khalatnikov (BK) [1961] and Donnelly [1999] to compare
these equations with the form they take in the reference frame of the superfluid.1
1
In making this comparison it is useful to recall the Galilean transformation of the
chemical potential, μ = μ − 12 |vs − vn |2 , where μ is evaluated in the superfluid
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 114–130, 2001.
c Springer-Verlag Berlin Heidelberg 2001
Introduction to HVBK Dynamics 115
Notation. Here ρ and ρs denote the total and superfluid mass densities, respec-
tively. The entropy density of the normal fluid is S, its temperature is denoted
T , and ρn = ρ − ρs denotes its mass density. The superfluid velocity is denoted
vs and
J = ρs vs + ρn vn
is the total momentum density. In the entropy equation R is the rate that heat is
produced by the phenomenological friction and reactive forces in f , which must
be specified to close the theory. We also denote
In the superfluid First Law, ω = curl vs is the superfluid vorticity with magnitude
|ω| = ω̂ ·ω, where ω̂ = ω/|ω| is its unit vector. BK [1961] takes the energy density
ε0 to depend on the magnitude of the superfluid vorticity, |ω|, as
ρs κ |ω| R
ε0 = ln .
4π a
This is the energy per unit length of a superfluid vortex line, ρs (κ2 /4π)ln(R/a),
with quantum of circulation κ = h/m 10−3 (cm2 /sec) and ratio R/a of mean
distance between vortices R to effective vortex radius a, times the vortex length
per unit volume, |ω|/κ. Hence, we find
∂ε0 ∂|ω| κ R
λ= = |λ| ω̂ with |λ|/ρs ≈ ln = λ0 ,
∂|ω| ∂ω 4π a
where one ignores the derivative of R κ/|ω| inside the logarithm. The ap-
pearance of λ in the stress tensor Tik shifts the pressure P , and div T introduces
an additional force − ω · ∇λ ≡ −ρs T into the motion equation. The quantity T
is called the “vortex line tension.”
BK [1961] assigned the following form to the phenomenological coupling force
f appearing in the superfluid velocity equation in (1),
f = (vL − vs ) × ω , vL = vn − ρs (α s0 + β ω̂ × s0 ) , (5)
where s0 = v0 − vn , v0 = vs + ρ−1
s curl λ , (6)
The BK [1961] form of the phenomenological force f also implies the dissipative
heating rate,
R = (J − ρvn + curl λ) · ω × (vL − vn ) ,
which is Galilean invariant and positive. Substituting this form of f into the
superfluid motion equation in (1) and taking its curl provides the following
equation for the superfluid vorticity, ω = curl vs ,
∂t ω = curl (vL × ω) .
The mean velocity v̄ also figures prominantly in Hills and Roberts [1977] discus-
sion of the HVBK equations. As we shall see, the alternative expression (12) for
the auxiliary vortex slip velocity in terms of the mean velocity v̄ arises naturally
in the Hamiltonian derivation of a slightly modified set of HVBK equations.
Introduction to HVBK Dynamics 117
These equations possess the same formal conservation and circulation properties
as HVBK, modulo redefining the vortex slip velocity as s rather than s0 . The
vortex slip velocity s in (12) is defined relative to the Galilean frame of the nor-
mal fluid, which is present only at finite temperature. The HVBK s0 in (6) is the
limit of the vortex slip velocity s for zero temperature, at which no normal fluid
remains. The vortex slip velocity s in (12) is a slight modification of s0 in (6)
necessary to incorporate finite temperature effects, without changing the form of
the HVBK theory, obtained from a Hamiltonian derivation of these equations in
the normal fluid frame. In the Hamiltonian framework, the energy-momentum
conservation laws and Kelvin circulation theorem are all natual consequences.
Moreover, the velocities v and vn are identified as being dual to the momenta
given by ρvs and ρn (vn − vs ), respectively.
Outline. We shall use a Hamiltonian approach with Lie-Poisson brackets to de-
rive the expression (12) for the vortex line velocity v at finite temperature from
first principles by using the energy E in (7) as the Hamiltonian. The momenta
conjugate to the velocities v and vn shall be our basic dynamical variables. The
finite temperature vortex line velocity v and slip velocity s determined this way
turn out to be
ρ
v = ρ−1 (J + curl λ) , and s = v − vn
s
(v0 − vn ) . (13)
ρ
the axial velocity decouples from the lateral motion. Therefore, under sufficiently
rapid rotation, the superfluid vortex filaments will straighten and become parallel
to the axis of rotation as they approach a steady state.
Numerical implications. This renormalization of the vortex line element slip
velocity in the HVBK equations from s0 → s s0 ρs /ρ is sensitive to temper-
ature, but it does not affect the vortex line tension. Therefore, a temperature
sensitive trade-off arises between mutual friction and vortex line tension that
may be worth testing in numerical simulations such as those reported in Hen-
derson and Barenghi [2000]. The HVBK equations are thought to break down in
the presence of strong counterflow. However, as general conservation laws there
is no mechanism in the equations that would signal this breakdown. A rotating
Rayleigh-Besnard experiment might be useful in testing the range of validity of
the HVBK equations (Barenghi, private communication). Such an experiment
might also indicate how these equations should be modified in the presence of
strong counterflow.
Experimental implications. Temperature sensitivity of the coupling between
the superfluid vortices and the normal fluid component is an area of intense
current investigation in superfluid Helium turbulence, see Donnelly [1999]. One
would like to know whether the ρs /ρ renormalization of the mutual friction forces
relative to the vortex line tension would matter significantly in comparisons of
the predictions of the HVBK equations with modern experiments in Helium
turbulence at low, but finite temperatures.
Superfluid vortex dynamics. To begin addressing this issue, we may use the
superfluid vorticity equation for the renormalized HVBK equations obtained in
the Appendix via the Hamiltonian approach to write an explicit equation for
the dynamics of Vinen’s vortex length density L = |ω|/κ. In the superfluid
turbulence decay experiments reported by Skrbek, Niemela and Donnelly [1999]
the spatial integral of this quantity is measured as a function of time to decrease
over six decades as t−3/2 . The integrated vortex length measured in these ex-
periments is predicted by the renormalized HVBK equations to be governed by
the superfluid vorticity dynamics alone.
Upon including mutual friction, the superfluid vortex dynamics for the renor-
malized HVBK equations is expressed as, cf. equation (9),
∂t ω = curl(vL × ω) ,
in which the renormalized total vortex line velocity given by, cf. equation (5),
B ρn Bρn
vL = v − s− ω̂ × s , where s = v − vn , (15)
2ρ 2ρ
and its Hamiltonian limit is found to be
v = v̄ + ρ−1 curl λ , with v̄ = ρ−1 J and λ = λ ω̂ . (16)
Thus, relative to the Hamiltonian approach, the terms in B and B are additional
velocities introduced by phenomenology, while v is the vortex line velocity in
the absence of mutual friction.
Introduction to HVBK Dynamics 119
The HVBK superfluid vorticity equation implies the following dynamics for
the integrated vortex length measured in the turbulence decay experiments,
d 3
Ld x = ω̂ · ∂t ω/κ d 3 x
dt
!" # !" #
length vorticity dynamics
= L ṽ · (ω̂ × curl ω̂) d 3 x
!" #
transport · curvature
Bρn
− L (ω̂ × curl λ ω̂) · (ω̂ × curl ω̂) d 3 x
2ρ2
!" #
damping by curvature
Bρn
+ L n̂ × ω̂ · (ω̂ × v ) + (v − vs ) dS . (17)
2ρ
!" #
creation and destruction at the boundary
B ρn Bρn
ṽ = v − (v − vn ) − ω̂ × (v − vn ) . (18)
2ρ 2ρ
According to the last term in (17), vortex length is created or destroyed at the
boundary, unless the vortex filaments approach it in the normal direction, so
that n̂ × ω̂ = 0.
Formula (17) for the evolution of the total superfluid vortex length presents a
trade-off between the mass-weighted velocity v and the local induction velocity
(or filament curvature) ω̂ × curl ω̂, in the interior of the domain. This trade-
off in the interior competes with the process of creation and destruction at the
boundary. For example, in counterflow turbulence, the superfluid moves toward
the heater at the boundary, so the term in v would tend to be nonzero. In
contrast, for grid turbulence, v is small, so this term would tend to contribute
less.
This formula governs the dynamics of the experimentally measured quantity
L d 3 x. However, it does not yet show how to obtain the t−3/2 decrease seen in
this quantity by Skrbek, Niemela and Donnelly [1999] in their experiments on
decay of turbulence.
Suppose the main source of decay were the term labeled “damping by cur-
vature” in formula (17) and the flow were isothermal and incompressible. This
would imply
1 d L/R2 3
L = − c0 (T ) λ0 =− , (19)
L dt L 2 (t + t0 )
120 Darryl D. Holm
The measured t−3/2 decrease in L implies via formula (19) that the length-
weighted mean curvature of the vortex tangle L/R2 /L decays due to mutual
friction as t−1 . Thus, on the average as the vortex length decays, the vor-
tices tend to straighten, under the effects of mutual friction damping.
Preservation of helicity versus preservation of vortex length. The he-
licity, or linkage number for the superfluid vorticity is defined as
Λ = (vs · ω) d 3 x .
The helicity satisfies an evolution equation obtained from the superfluid vortex
dynamics,
dΛ 1
= − (n̂ · ω) μ − vn2 − vs · (vL − vn ) dS − (n̂ · vL )(vs · ω) dS .
dt 2
Therefore, even with mutual friction, helicity is created and destroyed only on
the boundary. Moreover, helicity will be preserved, provided both ω and vL are
tangential at the boundary. The former condition, however, is the opposite of
that required for the creation and destruction of vortex length at the boundary
to cease. Therefore, no equilibrium should be expected that preserves both the
helicity and the vortex length in a superfluid.
Superfluid vortex equilibria are not ABC flows. The steady equilibrium
solutions of the superfluid vorticity dynamics satisfy
curl (vL × ω) = 0 . (23)
For example, a steady equilibrium exists when ω and vL are parallel. Note that
these “super-Beltrami flows” are not eigenfunctions of the curl. Therefore, they
are not Arnold-Beltrami-Childress (ABC) flows, as occur for the Euler equations.
As we have seen, finite temperature effects renormalize the total vortex line
velocity as
ρn B B
vL = v − ω̂ × s + s , where s = v − vn , (24)
ρ 2 2
and the Hamiltonian part of the line velocity (with corrections for finite temper-
ature) is defined as
To the extent that ρ, ρs , ρn and S all may be taken as constants for a given
temperature and the heating rate R is negligible, then the velocities vn and vs
are incompressible, i.e.,
∇ · vn = 0 and ∇ · vs = 0 .
In this situation, the pressure P may be obtained from the Poisson equation,
−∇2 P + λ · ω = div ρs (vs · ∇) vs + ρn (vn · ∇) vn − ω · ∇λ , (27)
found from the divergence of the total momentum equation. Combining the su-
perfluid motion equation with total momentum conservation results in an equa-
tion for the normal fluid velocity in the incompressible case
ρs
ρn ∂t vn + ρn (vn · ∇)vn = − ∇ P − ρs μ + vs2 n − ρs (vL − vs ) × ω + ω · ∇λ ,
2
where vL is given in equation (24). We set P ≡ P + λ · ω and take it as the
total pressure. (One also could have absorbed λ · ω into P earlier, by including
it in Euler’s pressure law.) Since λ = |λ|ω̂ and ω̂ is a unit vector, we find for
constant ρs the standard relation for the vortex line tension denoted as T.
Namely,
ω · ∇λ = − λ0 ρs ω × curl ω̂ ≡ ρs T ,
where λ0 = λ/ρs = (κ/4π)ln(b/a) is a constant.
Remark. We note that the quantity T known as the vortex line tension first
appears in the normal fluid equation, as a reaction to the presence of the super-
fluid. The standard convention for introducing the mutual friction force has the
effect of shifting T into the superfluid equation. By action and reaction, though,
T could appear in either equation.
These equations of motion must be completed by providing an equation of
state relation for the quantity μ − 12 vs2 n . BK [1961] assumes a law of partial
pressures,
ρn ρs 1
Pn = P = P − Ps and Ps = P = ρs μ − ρs vs2 n .
ρ ρ 2
122 Darryl D. Holm
In this case, the renormalized HVBK motion equations for incompressible flow
reduce to
1 ρn
∂t vs + (vs · ∇)vs = − ∇P − Fn s + T , (31)
ρ ρ
1 ρs
∂t vn + (vn · ∇)vn = − ∇P + Fn s . (32)
ρ ρ
In these superfluid motion equations, the renormalized mutual friction force
Fn s is defined as the sum (with ω = curl vs )
ρ B B 0
s
Fn s = (s 0 × ω) + Fn0 s , where Fn0 s = ω̂ × s 0 + s × ω . (33)
ρ 2 2
Here Fn0 s is the HVBK mutual friction force without any finite temperature cor-
rections. To acquire these formulas, we used the relations for the incompressible
case,
ρs ρs 0 ρs 0
s = v − vn = vs + λ0 curl ω̂ − vn = v − vn = s , (34)
ρ ρ ρ
with s 0 = vs + λ0 curl ω̂ − vn , and we eliminated vL by using the relation
ρ s ρn B B
− ρs vL − vs − λ0 curl ω̂ = ρn s + ω̂ × s + s . (35)
ρ 2 2
and there is no tendency for mutual friction to cause any alignment in the
vorticities of the superfluid and its normal component. Instead, the curl ω̂ part
of Fn0 s = 0 would break any such alignment, if it were to form spontaneously.
Indeed, alignments sufficient for steady solutions are
component. It also includes the interaction with the normal component, since
v depends on the relative momentum density and contains finite temperature
effects. The superfluid Coriolis force is essential in the spin up problem in He-II,
see, e.g., Reissenegger [1993].
Superfluid Taylor-Proudman theorem. For steady, or slow motions and
rapid rotation we have
0 = curl v∗ × 2Ω .
If the rotation is uniform (∇Ω = 0) and oriented vertically (Ω = |Ω|ẑ) this
becomes
T
0 = 2|Ω| ∂z v∗ − ẑ div v∗ = 2|Ω| ∂z v∗x , ∂z v∗y , −∂x v∗x − ∂y v∗y ,
where ( )T denotes transpose of a row vector into a column vector. Thus, for
steady, or slow motions and rapid uniform rotation, we find that vortex line
motion is columnar. That is, the lateral vortex line velocity is nondivergent
and independent of the axial coordinate, and the axial velocity decouples from
the lateral motion. Therefore, under sufficiently rapid rotation, the superfluid
vortex filaments will straighten and become parallel to the axis of rotation as
they approach a steady state. However, they may still undergo nondivergent
motion in the lateral plane. This superfluid Taylor-Proudman theorem explains
why steady superfluid vortices tend to be aligned with the rotation axis under
rapid uniform rotation. The same conclusion applies, if the velocity v∗ in the
Hamiltonian formulation is replaced by the phenomenological relative velocity
∗
vL = vL − R. Similar considerations are discussed in Sonin [1987] from a more
microscopic viewpoint.
Relative total momentum is not conserved for rotating compressible
flows. Since the Hamiltonian depends explicitly on spatial position, instead of
conserving relative total momentum, we have the balance
∂ h ρ
∂t Ji∗ + ∂j Ti∗ j = − i = ∂i |R|2 ,
∂x explicit 2
where h is the Hamiltonian density in equation (40). This relative momentum
balance is the effect of centrifugal force. Here we have dropped terms proportional
to ρ − n, since ρ = n is still preserved in a rotating frame. Consequently, the
stress tensor in the relative momentum equation also keeps its form in passing to
a rotating frame, although the total relative momentum is no longer conserved
if the flow is compressible.
Acknowledgments
I am grateful to H. R. Brand, A. Brandenburg, P. Constantin, R. Donnelly, V.
V. Lebedev, F. Lund, J. E. Marsden, J. Niemela, A. Reisenegger, L. Skrbek, K.
Sreenivasan and W. F. Vinen for stimulating discussions and encouragement. I
am also grateful for hospitality at the UC Santa Barbara Institute for Theoretical
Physics where this work was initiated during their Hydrodynamic Turbulence
Introduction to HVBK Dynamics 125
program in spring 2000. This research was supported by the U.S. Department of
Energy under contracts W-7405-ENG-36 and the Applied Mathematical Sciences
Program KC-07-01-01.
where the superfluid vorticity ω is the areal density of vortices and n̂ is the unit
vector normal to the surface S whose boundary ∂S moves with the vortex line
velocity v . When ω = curl vs this is equivalent to a vortex Kelvin theorem
d
vs · dx = 0 , (47)
dt ∂S(v )
∂t vs − v × ω = ∇μ . (48)
∂t φ + vn · ∇φ = ν . (49)
The mass density ρ and the phase φ are canonically conjugate in the Hamiltonian
formulation. Therefore, one may set ν = − δh/δρ for a given Hamiltonian h and
the phase gradient u = ∇φ satisfies
δh
∂t u + vn · ∇u + (∇vn )T · u = − ∇ , (50)
δρ
is advected by the vortex line velocity v , instead of the normal velocity vn . Ab-
sorbing all gradients into u yields
∂t A + v × ω = 0 . (52)
Taking the difference of the equations for u and A then recovers equation (48)
as δh
∂t vs − v × ω = − ∇ vn · u + with vs = u − A , (53)
δρ
in which one uses regularity of the phase φ to set curl u = 0. It remains to
determine v from the Hamiltonian formulation. Including the additional degree
of freedom A associated with the vortex lines allows them to move relative to
both the normal and super components of the fluid, and thereby introduces
an additional reactive force without introducing any additional inertia. This
Hamiltonian approach thus yields renormalized HVBK equations.
Proposition: Upon splitting the superfluid velocity into vs = u − A (with u =
∇φ so that ω = − curl A) the (renormalized) HVBK equations in the Galilean
frame of the normal fluid form an invariant subsystem of a Lie-Poisson
Hamiltonian system,
∂f
= {f, h} with f, h ∈ (M, ρ, S, u, A, n) ,
∂t
and Lie-Poisson bracket given by
{f, h} =
δf δh δh δh δh
− (Mk ∂j + ∂k Mj ) + ρ ∂j + S∂j + (∂k uj − uk , j )
δMj δMk δρ δS δuk
δf δf δf δh δf δh δf δh
+ ∂k ρ + ∂k S + (uk ∂j + uj , k ) + ∂k + ∂j
δρ δS δuj δMk δρ δuk δuj δρ
δf δh Aj , k − Ak , j δh δf δh
− ∂j + − ∂k d 3x . (55)
δAj δn n δAk δn δAk
Remarks: Here M is the total momentum density, the total mass density is ρ
and the entropy density is S. We shall interpret the density n later, after we
develop the Hamiltonian equations of motion. It shall emerge that n = ρ is an
invariant condition and, hence,
M − nA = J = p + ρvn ,
where
∂h ∂h ∂h ∂h
P = Ml +ρ +S +n − h,
∂Ml ∂ρ ∂S ∂n
as in the Euler relation for pressure.
Remark. One notes many parallels and correspondences among these equations.
Note especially the expected similarities in the equations for u and N/n. Recall
that A = − N/n, so that the superfluid velocity is given by vs = u − A =
u + N/n. The evolution of the superfluid velocity is consistently composed as
the sum of these two separate dynamical pieces.
Proof of the Proposition: The following Hamiltonian h (and conserved en-
ergy) will yield the HVBK equations in the frame of the normal fluid upon using
this Lie-Poisson bracket
1 2
h = d x − ρ vn + (M − ρA) · vn + ε0 (ρ, S, vs − vn , ω) .
3
(61)
2
The variational derivatives of the Hamiltonian h are computed in this reference
frame by using the thermodynamic first law (2). Namely,
1
δh = d 3 x μ − vn2 − A · vn δρ + T δS + vn · δM + (p + curl λ) · δu
2
− p + curl λ + ρvn · δA + M − p − ρvn − ρA · δvn .
Here we have used the velocity split δvs = δu − δA and assumed the boundary
condition n̂ · ω × λ = 0 when integrating by parts. This boundary condition
is satisfied identically, since λ = λ ω̂ in the HVBK theory. Upon substituting
these variational derivatives into the Lie-Poisson bracket, Corollary #1 yields
the following equations expressed in the normal fluid reference frame,
∂t S = {S, h} = − div(Svn ) ,
∂t n = {n, h} = − div(ρvn + p + curl λ) ,
∂t ρ = {ρ, h} = − div(ρvn + p + curl λ) ,
( Hence, the condition n = ρ is preserved.)
1
∂t u = {u, h} = − ∇(μ − vn2 + vn · vs ) + vn × curl u ,
2
( Hence, curl u = 0 is preserved.)
∂t A = {A, h} = n−1 ρvn + p + curl λ × curl A ,
( Hence, v = vn + ρ−1 (p + curl λ) when n = ρ is used.)
∂t Mj − nAj = {Mj − nAj , h} = − ∂k Tjk
Remarks:
(1.) Preservation of the condition n = ρ by these equations allows the introduc-
tion of the momentum-carrying field A as an independent degree of freedom
Introduction to HVBK Dynamics 129
In the stress tensor πjk the pressure P is defined by the Euler relation,
P = − ε0 + μρ + T S ,
so that in the normal-fluid frame the pressure satisfies
dP = ρdμ + SdT − p · d(vs − vn ) − λ · d ω .
The stress tensor Tjk = πjk + τjk may be derived by using Corollary #2 for
the Hamiltonian formulation.
130 Darryl D. Holm
References
1. Bekarevich, I. L. and I. M. Khalatnikov [1961] Phenomenological derivation of the
equations of vortex motion in He II, Sov. Phys. JETP 13 643-646.
2. Donnelly, R. J. [1999] Cryrogenic fluid dynamics, J. Phys.: Condens. Matter 11
7783-7834.
3. Gorter, C. J. and J. H. Mellink [1949] Physica 15 285.
4. Henderson, K. L. and C. F. Barenghi [2000] The anomalous motion of superfluid
helium in a rotating cavity, J. Fluid Mech. 406 199-219.
5. Hills, R. N. and P. H. Roberts [1977] Superfluid mechanics for a high density of
vortex lines, Arch. Rat. Mech. Anal. 66 43-71.
6. Holm, D. D. and B. Kupershmidt [1987] Superfluid plasmas: multivelocity non-
linear hydrodynamics of superfluid solutions with charged condensates coupled
electromagnetically, Phys. Rev. A 36 3947-3956.
7. Nemirovskii, S. K. and W. Fiszdon [1995] Chaotic quantized vortices and hydro-
dynamic processes in superfluid helium, Rev. Mod. Phys. 67 37-84.
8. Putterman, S. J. [1974] Superfluid Hydrodynamics, North Holland, Amsterdam.
9. Reissenegger, A. [1993] The spin up problem in Helium-II J. Low Temp. Phys. 92
77-106.
10. Skrbek, L., J. J. Niemela and R. J. Donnelly [1999] Turbulent flows at cryogenic
temperatures: a new frontier, J. Phys.: Condens. Matter 11 7761-7783.
11. Sonin, E. B. [1987] Vortex oscillations and hydrodynamics of rotating superfluids,
Rev. Mod. Phys. 59 87-155. 1987
12. Volovik, G. E. and V. S. Dotsenko [1979] Poisson brackets and continuous dynam-
ics of the vortex lattice in rotating He-II, JETP Lett. 29 576-579.
13. Volovik, G. E. and V. S. Dotsenko [1980] Hydrodynamics of defects in condensed
media in the concrete cases of vortices in rotating Helium-II and of disclinations
in planar magnetic substances, Sov. Phys. JETP 58 65-80.
Magnus Force, Aharonov–Bohm Effect,
and Berry Phase in Superfluids
Edouard Sonin
1 Introduction
If the vortex moves with respect to a liquid, classical [1] or quantum, there
is a force on the vortex normal to the relative vortex velocity with respect to
the liquid. This is the Magnus force, which plays an important role in mod-
ern condensed-matter physics. In particular, it determines the mutual fiction in
superfluids [2,3,4] and the Hall effect in superconductors [5].
An obvious generalization of the classical Magnus force in the superfluid
seemed to be a force proportional to the superfluid density ρs [2]:
ρs [(v L − v s ) × κ] = F , (1)
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 131–137, 2001.
c Springer-Verlag Berlin Heidelberg 2001
132 Edouard Sonin
area around a moving vortex demonstrates the existence of the Iordanskii force. I
also discuss the Berry phase. According to Refs. [11,13], the Berry phase and the
Magnus forces are proportional to the total current circulation at large distances.
But the total current circulation contains a normal-fluid contribution, which is
proportional to the Iordanski force. Taking this contribution into account, the
Berry-phase analysis agrees with the momentum-balance approach.
2 Gross–Pitaevskii Theory
and Two-Fluid Hydrodynamics
In the Gross–Pitaevskii theory [17] the ground state and weakly excited states
of a Bose-gas are described by the nonlinear Schrödinger equation
∂ψ 2 2
i =− ∇ ψ + V |ψ|2 ψ (3)
∂t 2m
for the condensate wave function ψ = a exp(iφ). The nonlinear Schrödinger
equation is the Euler–Lagrange equation for the Lagrangian
i ∂ψ ∂ψ ∗ 2 V
L= ψ∗ −ψ − |∇ψ|2 − |ψ|4 . (4)
2 ∂t ∂t 2m 2
The Noether theorem yields the momentum conservation law ∂ji /∂t+∇j Πij = 0
where j = Im{ψ ∗ ∇ψ} is the mass current, and the momentum-flux tensor is
2 ∗ ∗ V 2 2 2
Πij = (∇i ψ∇j ψ + ∇i ψ ∇j ψ) + δij |ψ| −
4
∇ |ψ| . (5)
2m 2 4m
Using the Madelung transformation [3], one obtains from complex Eq. (3)
two real equations for the liquid density ρ = ma2 and the liquid velocity v =
(/m)∇φ = (κ/2π)∇φ. Far from the vortex line these equations are hydrody-
namic equations for an ideal inviscous liquid:
∂ρ
+ ∇(ρv) = 0 , (6)
∂t
∂v
+ (v · ∇)v = −∇μ . (7)
∂t
Here μ = V a2 /m is the chemical potential. Equation (5) becomes the hydrody-
namic momentum-flux tensor Πij = P δij + ρvi (r)vj (r).
A plane sound wave propagating in the liquid generates the phase variation
φ(r, t) = φ0 exp(ik · r − iωt). Then ρ(r, t) = ρ0 + ρ(1) (r, t) and v(r, t) = v 0 +
v (1) (r, t), where ρ0 and v 0 are the average density and velocity in the liquid,
whereas ρ(1) (r, t) and v (1) (r, t) = (κ/2π)∇φ are periodical variations of the
density and the velocity due to the sound wave (ρ(1) = 0, v (1) = 0). Equations
(6) and (7) linearized with respect to ρ(1) and v (1) yield the sound equation for
φ with the sound velocity cs = V a2 /m and the spectrum ω = cs k + k · v 0 .
Magnus Force, Aharonov–Bohm Effect, and Berry Phase 133
The total mass current expanded up to the terms of the second order with
respect the wave amplitude and averaged over time is
κ2 k
j = ρ0 v 0 + ρ(1) v (1) = ρ0 v 0 + ρ0 φ20 k. (8)
8π 2 cs
If there is an ensemble of phonons with the Planck distribution
−1 −1
E(p) − p · v n cs p + p · (v 0 − v n )
n0 (E, v n ) = exp −1 = exp −1 ,
T T
(9)
the total mass current linearized with respect to v 0 − v n is
1
j = ρ0 v 0 + 3 n0 (p)p d3 p = ρ0 v 0 + ρn (v n − v 0 ) . (10)
h
Here p = k is the phonon momentum, E = cs p+p·v 0 is the phonon energy, and
v n is the drift velocity of phonons. This expression is equivalent to the two-fluid
expression assuming that ρ = ρ0 = ρs + ρn , v 0 = v s , and the normal density is
given by the usual two-fluid expression ρn = −(1/3h3 ) [∂n0 (ε, 0)/∂E]p2 d3 p.
Expanding the momentum-flux tensor up to terms of the second-order with
respect to the sound wave amplitude one obtains:
c2s ρ(1) v(1)
2 2
Πij = P0 δij + ρ0 v0i v0j + − ρ0 δij
ρ0 2 2
+ρ(1) (v(1) )i v0j + ρ(1) (v(1) )j v0i + ρ0 (v(1) )i (v(1) )j . (11)
For the Planck distribution this yields the two-fluid momentum flux tensor
In presence of a vortex the sound equation is (see Refs. [8,14] for more details)
∂2φ ∂φ
2
− c2s ∇2 φ = −2v v (r) · ∇ , (13)
∂t ∂t
where
κ×r
v v (r) = (14)
2πr2
134 Edouard Sonin
is the circular velocity field induced by a vortex line. Here r is the position vector
in the plane xy. The sound wave produces the density variation
ρ0 κ ∂φ
ρ(1) = − 2 + v v · ∇φ(r) . (15)
cs 2π ∂t
One can calculate the differential cross-section in the Born approximation,
but since it is quadratic in κ this does not yield a transverse force [7,8,14]. Instead
we consider a quasiclassical solution of the sound equation:
iδS iκk
φ = φ0 exp −iωt + ik · r + = φ0 exp(−iωt + ik · r) 1 + θ , (16)
2πcs
r
where δS = −(k/cs ) v v · dl = θκk/2πcs is the variation of the action due
to interaction with the circular velocity around the vortex. The angle θ is an
azimuth angle for the position vector r measured from the direction opposite
to the wave vector k. This choice provides that the quasiclassical correction
vanishes for the incident wave far from the vortex. One can check directly that
Eq. (16) satisfies the sound equation (13) in the first order of the parameter
κk/cs . The velocity generated by the sound wave around the vortex is
κ κ ik
v (1) = ∇φ = φ0 exp(−iωt + ik · r) ik − v v . (17)
2π 2π cs
According to Eq. (16) the phase φ is multivalued, and one must choose a cut for
an angle θ at the direction k, where θ = ±π . The jump of the phase on the cut
line behind the vortex is a manifestation of the Aharonov–Bohm effect [9]: the
sound wave after its interaction with the vortex has different phases on the left
and on the right of the vortex line. This results in an interference [8,14]. The
width of the interference region is dint ∼ r/k.
Now we consider the momentum balance using the condition that dSj Π⊥j =
0 for a cylindrical surface around the vortex line. The subscript ⊥ points a com-
ponent normal to the wave vector k of the incident wave. The total momentum-
flux tensor can be obtained from Eq. (11) assuming v 0 (r) = v v (r) + v s :
The first two terms in this expression yield the momentum flux without phonons,
which produces the Magnus force for a liquid with the density ρ0 and the ve-
locity v s . The rest terms cancel except for the contribution from the term
ρ0 (v(1) )i (v(1) )j in the interference region where v(1)⊥ = (κ/2πr)∂φ/∂θ. The
contribution depends on the phase jump and for a single sound wave is:
1 κ2 k 1 κ3 k 2
ρ0 (v(1) )⊥ (v(1) )r rdθ = ρ0 φ20 [δS(−) − δS(+)] = ρ0 φ20 ,
8π 2 8π 2 cs
(19)
where δS(±) = ∓κk/2cs are the action variations at θ → ∓π.
Magnus Force, Aharonov–Bohm Effect, and Berry Phase 135
For the Planck phonon distribution the condition dSj Π⊥j = 0 yields:
force on the normal fluid. This is confirmed by the solution of the Navier-Stokes
equation obtained by Thouless et al. [20]. However, separation on longitudinal
and transverse components of a force should be done with respect to the normal
velocity v n∞ − v L , but not v n − v L . Using Eq. (23), Eq. (2) for the force from
the normal fluid can be rewritten (neglecting the longitudinal force ∝ D) as
1 D ln(rm /lph )
F =− 2 (v L − v n∞ ) + D [ẑ × (v L − v n∞ )] .
D ln(rm /lph ) 4πηn
1+ 4πηn
(24)
The transverse component of this force determines the normal circulation at very
large distances:
D κ
κn = dl · v n = − 2 = 2 , (25)
D ln(rm /lph ) κρn ln(rm /lph )
ρn 1 + 4πηn
1 + 4πηn
where we used the value D = −κρn for the Iordanskii force. In the limit of
a strong viscous drag κρn ln(rm /l)/4πηn 1 the transverse force and related
normal circulation are suppressed [20]. But the effect of the superfluid Magnus
force and the longitudinal force ∝ D is also suppressed in this limit.
References
1. H. Lamb, Hydrodynamics (Cambridge University Press, New York, 1975).
2. H.E. Hall and W.F. Vinen, Proc. Roy. Soc. A238, 204 (1956).
3. R.J. Donnelly, Quantized vortices in helium II (Cambridge University Press, Cam-
bridge, 1991), Sec. 2.8.3.
4. E.B. Sonin, Rev. Mod. Phys. 59, 87 (1987).
5. P. Nozières and W.F. Vinen, Phil. Mag. 14, 667 (1966).
6. E.M. Lifshitz and L.P. Pitaevskii, Zh. Eksp. Teor. Fiz. 33, 535 (1957) [Sov. Phys.-
JETP 6, 418 (1958)].
7. S.V. Iordanskii, Zh. Eksp. Teor. Fiz. 49, 225 (1965) [Sov. Phys.-JETP 22, 160
(1966)].
8. E.B. Sonin, Zh. Eksp. Teor. Fiz. 69, 921 (1975) [Sov. Phys.-JETP 42, 469 (1976)].
9. Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959).
10. A.L. Shelankov, Europhys. Lett., 43, 623 (1998).
11. P. Ao and D.J. Thouless, Phys. Rev. Lett. 70, 2158 (1993).
12. M.V. Berry, Proc. R. Soc. London A 392, 45 (1984).
13. M.R. Geller, C. Wexler, and D.J. Thouless, Phys. Rev. B 57, R8119 (1998).
14. E.B. Sonin, Phys. Rev. B 55, 485 (1997).
15. H.E. Hall and J.R. Hook, Phys. Rev. Lett., 80, 4356 (1998); E.B. Sonin, ibid. 81,
4276 (1998); C. Wexler et al., ibid. 80, 4357 (1998); 81, 4277 (1998).
16. A more detailed report on the present analysis is to be published in Proceedings
of the the workshop “Microscopic structure and dynamics of vortices in uncon-
ventional superconductors and superfluids” Dresden, Germany, March 2000, cond-
mat/0104221.
17. E.P. Gross, Nuovo Cimento 20, 454 (1961); L.P. Pitaevskii, Zh. Eksp. Teor. Fiz.
40, 646 (1961) [Sov. Phys.-JETP 13, 451 (1961)].
18. S.J. Putterman and P.H. Roberts, Physica 117 A, 369 (1983).
19. Existence of the normal circulation at large distances in the presence of the trans-
verse force on the vortex was pointed out by Pitaevskii (unpublished).
20. D.J. Thouless, M.R. Geller, W.F. Vinen, J-Y. Fortin, and S.W. Rhee, cond-
mat/0101297.
Using the HVBK Model
to Investigate the Couette Flow of Helium II
Karen L. Henderson
1 Introduction
When the temperature of liquid helium drops below the transition temperature
of Tλ = 2.172 k a phase transition occurs and it becomes a quantum liquid
called helium II. Helium II can be described macroscopically by Landau’s two-
fluid model in which it is considered to be a mixture of a viscous normal fluid
and an inviscid superfluid. In addition, vortex lines appear in the superfluid
component when helium II rotates or when it moves along a tube faster than
a small critical velocity. Feynman [1] showed that the circulation about each
individual vortex line is quantised, taking the value of Γ = 9.97 × 10−4 cm2 /sec.
The most generally accepted equations for modelling the macroscopic flow of
helium II are the Hall-Vinen-Bekharevich-Khalatnikov (HVBK) equations which
were derived by a number of people over the years [2,3,4,5,6]. These equations
extend Landau’s two-fluid model to take into account the presence of quantized
vortex lines in the flow. The derivation of the equations is based on a continuum
approximation, assuming a high density of vortex lines, all aligned roughly in
the same direction.
The incompressible isothermal HVBK equations of motion of the two fluids
are
∂vn ρs
+ (vn ·∇)vn = −∇pn + νn ∇2 vn + F , (1a)
∂t ρ
∂vs ρ n
+ (vs ·∇)vs = −∇ps − νs T − F, (1b)
∂t ρ
∇· vn = 0, ∇· vs = 0. (1c)
where vn and vs are the velocity profiles of the normal fluid and superfluid
respectively, ρn and ρs the normal fluid and superfluid densities, ρ = ρn + ρs
helium’s total density, pn and ps effective pressures and νn the normal fluid
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 138–145, 2001.
c Springer-Verlag Berlin Heidelberg 2001
Using the HVBK Model to Investigate the Couette Flow of Helium II 139
kinematic viscosity. The relative amount of normal fluid and superfluid present
in the flow depends on the temperature T of the liquid: if T → Tλ then ρs /ρ → 0
and ρn /ρ → 1; if T → 0 then ρs /ρ → 1 and ρn /ρ → 0.
The mutual friction force, F , describes the interaction between the normal
fluid and the vortex lines and is given by
F = 12 B ω
× (ωs × (vn − vs − νs ∇× ω
)) + 12 B ωs × (vn − vs − νs ∇× ω
s s s
), (2)
2 Linear Theory
In the case of flow between infinite cylinders with inner radius R1 rotating with
angular velocity Ω and stationary outer radius R2 , Couette flow, whose velocity
profile is given by:
v c = (A + C/r) er (4)
is an exact solution of the HVBK equations for both the normal fluid and su-
perfluid. This is provided that Ω is greater than a small critical value at which
vortex lines first appear in the gap. A and C are constants depending on R1 ,
R2 and Ω determined by the no-slip condition imposed on the normal fluid at
the cylinder walls. The superfluid, being inviscid is not required to satisfy such
boundary conditions. The only restriction is that there is no penetration through
the boundary. From (4) it can be seen that the superfluid vorticity is purely ax-
ial and has magnitude 2|A|, thus the vortex lines are aligned in the direction of
rotation.
Early attempts to theoretically examine the stability of Couette flow theo-
retically were made by Chandrasekhar & Donnelly [13], however the issue has
only recently been resolved fully by Barenghi & Jones [14] and Barenghi [15].
They performed a linear stability analysis on the HVBK equations. The Couette
state was linearly perturbed and they numerically calculated the critical angular
velocity, Ωc , and corresponding critical axial wavenumber at which Couette flow
becomes linearly unstable. We summarize their findings below:
For a classical fluid the critical axial wavelength at which Couette flow becomes
unstable and Taylor vortices form occurs at kc ≈ 3.1, in other words the result-
ing Taylor vortices are approximately square. This is not the case for helium II.
As the temperature decreases below the lambda temperature of Tλ = 2.172k the
Using the HVBK Model to Investigate the Couette Flow of Helium II 141
3 Nonlinear Solutions
3.1 Infinite Cylinder Assumption
Using the infinite cylinder assumption, Henderson, Barenghi & Jones [17] nu-
merically solved the HVBK equations for the first time to obtain the nonlinear
flow of helium II between infinite cylinders. The aim of the work was to ob-
tain solutions for helium II corresponding to what would correspond to Taylor
vortices in a classical flow and to investigate what happens to the vortex lines.
Axisymmetric solutions were considered since linear theory predicts that the ax-
isymmetric mode onsets first. To solve equations (1a-1c), boundary conditions
are required. The nonlinear problem is 6th order in both the normal fluid and
superfluid, however the linear problem is only 2nd order in the superfluid. Thus
two further boundary conditions for the superfluid are needed in addition to the
no penetration of the boundary used successfully in the linear stability analysis.
The extra boundary conditions employed were
Equations (5a,5b) force the superfluid vorticity to be purely axial at the cylinder
walls. This is consistent with Couette flow, in which the vortex lines are purely
axial throughout the flow and results in the mutual friction being small at the
boundaries which is an advantage numerically. The normal fluid satisfies the
standard no slip boundary conditions as for the linear model.
142 K.L. Henderson
In order to be able to compare with further experimental data, end effects need
to be included in the model. Henderson & Barenghi [21] considered helium II
contained within a cylindrical annulus of inner radius R1 , outer radius R2 , height
H where the gap between the cylinders has been chosen such that H = R2 −
R1 . Thus the Couette annulus has unit aspect ratio, in that the gap between
the cylinders is equal to the height between the endcaps. The inner cylinder
rotates with constant angular velocity Ω, whilst the outer cylinder and two end
plates are stationary. This simple flow configuration enabled us to study how
the vortex lines respond to a shear in the presence of boundaries which are both
parallel and perpendicular to the natural axial direction of the vortex lines. The
axisymmetric form of the HVBK equations (1a-1c) were solved using a finite
difference approach taking a regular grid in both the r and z direction. The
boundary conditions on the curved cylinder walls were taken to be the same as
for the infinite cylinder case. However extra boundary conditions are also needed
on the two endcaps, z = 0, H. For the normal fluid standard no slip boundary
conditions were imposed. Whilst for the superfluid the following were used
The first condition (6a) ensures that there is no penetration of the superfluid
through the boundary, whilst the last two conditions (6b,6c) correspond to per-
fect sliding of the vortex lines as discussed by Khalatnikov [5].
The main result of this investigation is the anomalous motion of helium II
when compared to the motion of a classical fluid. The velocity profile obtained
is a superposition of an azimuthal motion vφ around the inner cylinder and a
toroidal motion vr and vz in the vertical plane. The latter motion is in the form
of a pair of cells similar to a Taylor vortex pair, but being caused by boundaries
rather than a centrifugal instability, it is hereafter referred to an Ekman cell
pair. The first interesting finding is that vφs is almost z-independent, that is
the superfluid moves around the cylinders in a column-like fashion, which is
due to the tension in the vortex lines. This effect becomes more pronounced at
lower temperatures when the superfluid component is higher as is illustrated in
Fig. 1a,b. Each figure extends over the whole computational domain with the
inner cylinder and outer cylinder on the left and right respectively. In contrast,
vφn exhibits strong z-dependence due to the no-slip boundary conditions imposed
on the normal fluid at the ends and walls of the cylinders and has a similar profile
to that of a classical fluid, see Fig. 1c.
144 K.L. Henderson
Fig. 1. Azimuthal motion vφ of (a) the superfluid at T =1.8 K, (b) the superfluid at
T =2.17 K, (c) the normal fluid at T =2.17 K. Lighter regions correspond to larger
magnitude.
Fig. 2. Motion of helium II compared to a classical fluid. Classical: (a) radial veloc-
ity vr ; (b) axial velocity vz . Helium II at T =2.11 K: (c) vrn ; (d) vzn ; (e)vrs ; (f) vzs .
Lighter/darker regions correspond to positive/negative velocities.
The second interesting finding comes from looking at the Ekman cells in both
the normal fluid and superfluid. In a classical fluid the two Ekman cells form with
outflow at the centre and inflow at the ends of the cylinder, as in Fig. 2a,b. The
results for helium II are quite different. We find that the superfluid Ekman cells
Using the HVBK Model to Investigate the Couette Flow of Helium II 145
always rotate in a counter-classical way due to the mutual friction force, that is
outflow occurs at the ends of the cylinder with inflow at the centre, see Fig. 2c,d.
It is also seen that the normal fluid Ekman cells rotate in a counter-classical way
at lower temperatures, see Fig. 2e,f, but revert to a classical direction close to the
transition temperature Tλ = 2.172k as one would expect. We also investigated
the magnitude and direction of the superfluid vorticity in order to gain a picture
of how the vortex lines are situated in the flow. As for the infinite cylinder case,
the superfluid vorticity is primarily axial with small deflections in both the r
and φ direction. However the vorticity is concentrated near the inner rotating
cylinder, which is a measurable result.
4 Discussion
The excellent agreement between the linear stability analysis [15] and exper-
imental data [16] was a rigorous test of the validity of the HVBK equations
at least in the linear regime. The HVBK model has been validated further by
the good agreement between the nonlinear calculation [18] and experimental
data [19]. Obtaining nonlinear solutions for the flow of helium II between infi-
nite and finite cylinders has allowed us to gain more insight into the flow which,
because of the low temperature environment, cannot be observed directly like
a classical fluid. It has also enabled us to explore the boundary conditions for
the superfluid. A possible future direction of the work would be to extend the
aspect ratio in order to investigate the transition to Taylor cells.
References
1. R.P. Feynman: ‘Application of quantum mechanics to liquid helium.’ In P rogress
in Low Temperature Physics 1. (C.J. Gorter, North Holland 1955)
2. H.E. Hall, W.F. Vinen: Proc. Roy. Soc. London A 238, 215 (1956)
3. H.E. Hall: Phil. Mag. Suppl. 9, 89 (1960)
4. I.L. Bekharevich and I.M. Khalatnikov: Sov. Phys. JETP 13 , 643, (1961)
5. I.M. Khalatnikov: An Introduction to the Theory Superfluidity. (Benjamin 1965)
6. R.N. Hills, P.H. Roberts: Arch. Rat. Mech. Anal. 66, 43 (1977)
7. C.F. Barenghi, R.J. Donnelly, W.F. Vinen: J. Low Temp. Phys. 52, 189 (1983)
8. D.C. Samuels & R.J. Donnelly: Phys. Rev. Lett. 65, 187 (1990)
9. G.I. Taylor: Phil. Trans. Roy. Soc. Lond. A 223, 289 (1923)
10. F. Bielert, G. Stamm: Cryogenics 33, 938 (1993)
11. P.L. Kapitza: J. Phys. USSR 4, 181 (1941)
12. R.J. Donnelly, M.M. Lamar: J. Fluid Mech. 186, 163 (1988)
13. S. Chandrasekhar, R.J. Donnelly: Proc. Roy. Soc. London A 241, 9 (1957)
14. C.F. Barenghi, C.A. Jones: J. Fluid Mech. 197, 551 (1988)
15. C.F. Barenghi: Phys. Rev. B 45, 2290 (1992)
16. C.J. Swanson, R.J. Donnelly: Phys. Rev. Lett. 67, 1578 (1991)
17. K.L Henderson, C.F. Barenghi, C.A. Jones: J. Fluid Mech. 283, 329 (1995)
18. K.L. Henderson, C.F. Barenghi, J. Low Temp. Phys. 98, 351 (1995)
19. R.J. Donnelly: Phys. Rev. Lett. 3, 507 (1959)
20. K.L. Henderson, C.F. Barenghi: Phys. Lett. A 191, 438 (1994)
21. K.L. Henderson and C.F. Barenghi: J. Fluid Mech. 406, 199 (2000)
An Introduction
to the Theory of Superfluid Turbulence
W.F. Vinen
1 Introduction
In one sense superfluid turbulence is an old subject: it was mentioned as a theo-
retical possibility by Feynman in 1955[1]; and it has been known experimentally
since the early 1950s that flow of the superfluid component of helium II can
become turbulent when there is a steady counterflow of the two fluids, such as
occurs in a steady heat current[2]. The original experimental discovery was ac-
companied by the beginnings of a theory[3], and this theory has been developed
steadily, especially by Schwarz[4,5], so that many aspects of this type of turbu-
lence are now well understood. However, counterflow turbulence has no classical
analogue, and it has attracted little interest from those who study classical fluid
mechanics. Types of flow for which classical analogues do exist were observed
by low temperature physicists for many years, but the presence of the two fluids
were thought to make them very complicated, and they were not therefore stud-
ied in detail. More recently experiments have been reported on the analogue of a
rather simple case of classical turbulence, namely that produced by steady flow
through a grid[6,7]. In the classical analogue the turbulence is approximately ho-
mogeneous and isotropic, and its study has been important in the development
an understanding of classical turbulence[8]. The superfluid analogue promises to
be equally important.
In this paper I shall first describe some aspects of the theory of counterflow
turbulence. But I shall then devote most of the paper to grid turbulence, where
the theory is less well developed, although I shall make use of an important
experimental result obtained with a more complicated type of flow generated by
two counter-rotating discs[9]. My aim is to stimulate interest in the theory of
superfluid turbulence, particularly, at this stage, in the simple case of grid tur-
bulence, among both low temperature physicists and those with a background in
classical fluid mechanics. I shall focus on open questions and unsolved problems,
questions and problems that are clearly seen in grid turbulence, but which are
more widely relevant. My own background is in experimental quantum fluids,
and certainly not in theoretical fluid mechanics. I shall tend to speculate about
what I see as the physics of superfluid turbulence, and others will tell me where
my physical intuition is unreliable or, hopefully, where it can be developed along
more rigorous lines.
C.F. Barenghi, R.J. Donnelly, and W.F. Vinen (Eds.): LNP 571, pp. 149–161, 2001.
c Springer-Verlag Berlin Heidelberg 2001
150 W.F. Vinen
2 Counterflow Turbulence
In counterflow turbulence the vortex tangle is believed to be at least approxi-
mately homogeneous, provided that the average velocities of the two fluids, Vn
and Vs , are spatially uniform. The turbulence is maintained by the mutual fric-
tion. If the vortex lines move on average with the superfluid, the total average
force of mutual friction per unit volume, Fns , is equal to γL (Vn − Vs ), where L
The Theory of Superfluid Turbulence 151
is the length of line per unit volume, and where we have ignored factors of order
unity arising from the random orientation of the lines. We can illustrate one
approach to superfluid turbulence by deriving the dependence of L on (Vn − Vs )
from a principle of dynamical similarity[3,14,15].
Let us assume that the vortex tangle is characterized by a single length,
= L−1/2 , characteristic of both the vortex line spacing and the vortex radius
of curvature. Taking into account the Magnus effect and the force f , we can
easily show that the velocity with which any element of line moves is given by
where vs is the superfluid velocity at the element due to the rest of tangle,
κ̂ is the unit vector parallel to the element, and α ≈ γ/ρs κ. There are two
contributions to the magnitude of vs : vs1 ≈ κ/2π, which is due to neighbouring
lines at the distance from the element; and vs2 ≈ (κ/4π) ln (/ξ0 ), which is
due to the local curvature of the line. (We note in passing that computations
based on the“localized induction approximation”take into account only the latter
contribution.)
Suppose that we change the length scale in the vortex tangle by a factor g, so
that → g, and let us ignore the logarithmic factor in vs2 . Then vs changes by
the factor g −1 . If we change Vs and Vn by the same factor, we see from equation
(1) that vL is also changed by the same factor (formally, we change the length
scale by g and the time scale by g 2 ). Therefore the whole tangle evolves in the
same way as it would before scaling, except for the scaling factor g, suggesting
that there is a principle of dynamical scaling. Application of this principle shows
−1
easily that is proportional to (Vs − Vn ) , so that the mutual friction per
3
unit volume is proportional to (Vs − Vn ) , as is observed to be approximately
true. Taking into account the logarithmic term in vs2 introduces logarithmic
corrections, which are indeed probably observed.
As we have mentioned, the superfluid turbulence is maintained by the mutual
friction acting on individual elements of line. That the length of line can in
principle increase as a result of the mutual friction is clear from the fact[3]
that a vortex ring can grow if the self induced velocity of the ring is in the
direction of, but less in magnitude than, (Vn − Vs ); otherwise it will shrink.
Appropriately oriented parts of the tangle with low curvature can behave in a
similar way, although it is not obvious that a steady turbulent state (a finite
L) can be maintained. A detailed understanding of counterflow turbulence came
only from the simulations of Schwarz[4,5], which showed that a steady turbulent
state can be achieved through the effect of reconnections, which generate vortex
configurations that favour the growth of line.
This theory of counterflow turbulence is based on the assumption that the
flow of the normal fluid is laminar. The turbulent flow of the superfluid must
then occur on length scales not significantly larger than ; flow on larger length
scales would be damped out by mutual friction. Recently it has been suggested by
Melotte and Barenghi[16] that the laminar flow of the normal fluid in counterflow
may not always be stable, so that both fluids may become turbulent, probably
152 W.F. Vinen
on length scales significantly larger than . The theory for such a regime presents
us with a major challenge, which we mention again in Sect. 3.5.
parallel to the axis of rotation have a character that depends on the magnitude
(q) of the wavevector relative to the line spacing . If q 1 the waves are Kelvin
waves on the individual vortices; if q 1 the waves become indistinguishable
from the classical “inertia waves”found in a classical rotating liquid (see, for
example, references[19,20]. In the case of turbulent flow we must remember that
the non-linear term in the Navier-Stokes equation couples motion on different
length scales, so that the validity of our suggestion depends on the hypothesis
of the “independence of Fourier components for distant wavevectors”[8]. But we
know of no formal proof of this validity.
We emphasize that this similarity between superfluid and classical flow can-
not extend to wavenumbers of order or greater than −1 , where the discrete
nature of the vorticity cannot be ignored.