You are on page 1of 18

Fluid mechanics is the branch of physics concerned with the mechanics of fluids (liquids, gases,

and plasmas) and the forces on them. It has applications in a wide range of disciplines,
including mechanical, civil, chemicaland biomedical
engineering, geophysics, oceanography, meteorology, astrophysics, and biology.
Fluid Mechanics can also be defined as the science which deals with the study of behaviour of fluids
either at rest or in motion.
It can be divided into fluid statics, the study of fluids at rest; and fluid dynamics, the study of the
effect of forces on fluid motion. It is a branch of continuum mechanics, a subject which models
matter without using the information that it is made out of atoms; that is, it models matter from
a macroscopic viewpoint rather than from microscopic. Fluid mechanics, especially fluid dynamics, is
an active field of research, typically mathematically complex. Many problems are partly or wholly
unsolved, and are best addressed by numerical methods, typically using computers. A modern
discipline, called computational fluid dynamics (CFD), is devoted to this approach.Particle image
velocimetry, an experimental method for visualizing and analyzing fluid flow, also takes advantage of
the highly visual nature of fluid flow.

Contents

 1Brief history
 2Main branches
o 2.1Fluid statics
o 2.2Fluid dynamics
 3Relationship to continuum mechanics
 4Assumptions
 5Navier–Stokes equations
 6Inviscid and viscous fluids
 7Newtonian versus non-Newtonian fluids
o 7.1Equations for a Newtonian fluid
 8See also
 9Notes
 10References
 11Further reading
 12External links

Brief history[edit]
Main article: History of fluid mechanics
The study of fluid mechanics goes back at least to the days of ancient Greece,
when Archimedes investigated fluid statics and buoyancy and formulated his famous law known now
as the Archimedes' principle, which was published in his work On Floating Bodies—generally
considered to be the first major work on fluid mechanics. Rapid advancement in fluid mechanics
began with Leonardo da Vinci (observations and experiments), Evangelista Torricelli (invented
the barometer), Isaac Newton (investigated viscosity) and Blaise Pascal (researched hydrostatics,
formulated Pascal's law), and was continued by Daniel Bernoulli with the introduction of
mathematical fluid dynamics in Hydrodynamica (1739).
Inviscid flow was further analyzed by various mathematicians Jean le Rond d'Alembert, Joseph
Louis Lagrange, Pierre-Simon Laplace, Siméon Denis Poisson) and viscous flow was explored by a
multitude of engineers including Jean Léonard Marie Poiseuille and Gotthilf Hagen. Further
mathematical justification was provided by Claude-Louis Navier and George Gabriel Stokes in
the Navier–Stokes equations, and boundary layers were investigated (Ludwig Prandtl, Theodore von
Kármán), while various scientists such as Osborne Reynolds, Andrey Kolmogorov, and Geoffrey
Ingram Taylor advanced the understanding of fluid viscosity and turbulence.

Main branches[edit]
Fluid statics[edit]
Main article: Fluid statics
Fluid statics or hydrostatics is the branch of fluid mechanics that studies fluids at rest. It embraces
the study of the conditions under which fluids are at rest in stable equilibrium; and is contrasted
with fluid dynamics, the study of fluids in motion. Hydrostatics offers physical explanations for many
phenomena of everyday life, such as why atmospheric pressure changes with altitude, why wood
and oil float on water, and why the surface of water is always level whatever the shape of its
container. Hydrostatics is fundamental to hydraulics, the engineering of equipment for storing,
transporting and using fluids. It is also relevant to some aspects of geophysics and astrophysics (for
example, in understanding plate tectonics and anomalies in the Earth's gravitational field),
to meteorology, to medicine (in the context of blood pressure), and many other fields.

Fluid dynamics[edit]
Main article: Fluid dynamics
Fluid dynamics is a subdiscipline of fluid mechanics that deals with fluid flow—the science of liquids
and gases in motion. Fluid dynamics offers a systematic structure—which underlies these practical
disciplines—that embraces empirical and semi-empirical laws derived from flow measurement and
used to solve practical problems. The solution to a fluid dynamics problem typically involves
calculating various properties of the fluid, such as velocity, pressure, density, and temperature, as
functions of space and time. It has several subdisciplines itself, including aerodynamics (the study
of air and other gases in motion) and hydrodynamics (the study of liquids in motion). Fluid
dynamics has a wide range of applications, including calculating forces and movements on aircraft,
determining the mass flow rate of petroleum through pipelines, predicting evolving weather patterns,
understanding nebulae in interstellar space and modeling explosions. Some fluid-dynamical
principles are used in traffic engineering and crowd dynamics.

Relationship to continuum mechanics[edit]


Fluid mechanics is a subdiscipline of continuum mechanics, as illustrated in the following table.

Elasticity
Describes materials that return to their rest shape after
Continuum applied stresses are removed.
mechanics Solid mechanics
The study of the The study of the physics of
physics of continuous materials with a
continuous defined rest shape. Plasticity Rheology
materials Describes materials that The study of materials
permanently deform after a with both solid and fluid
sufficient applied stress. characteristics.
Non-Newtonian fluids do
not undergo strain rates
Fluid mechanics proportional to the applied
The study of the physics of shear stress.
continuous materials which
deform when subjected to a
force.
Newtonian fluids undergo strain rates proportional
to the applied shear stress.

In a mechanical view, a fluid is a substance that does not support shear stress; that is why a fluid at
rest has the shape of its containing vessel. A fluid at rest has no shear stress.

Assumptions[edit]

Balance for some integrated fluid quantity in a control volume enclosed by a control surface.

The assumptions inherent to a fluid mechanical treatment of a physical system can be expressed in
terms of mathematical equations. Fundamentally, every fluid mechanical system is assumed to
obey:

 Conservation of mass
 Conservation of energy
 Conservation of momentum
 The continuum assumption
For example, the assumption that mass is conserved means that for any fixed control volume (for
example, a spherical volume)—enclosed by a control surface—the rate of change of the mass
contained in that volume is equal to the rate at which mass is passing through the surface
from outside to inside, minus the rate at which mass is passing from inside to outside. This can be
expressed as an equation in integral form over the control volume.[1]
The continuum assumption is an idealization of continuum mechanics under which fluids can be
treated as continuous, even though, on a microscopic scale, they are composed of molecules. Under
the continuum assumption, macroscopic (observed/measurable) properties such as density,
pressure, temperature, and bulk velocity are taken to be well-defined at "infinitesimal" volume
elements—small in comparison to the characteristic length scale of the system, but large in
comparison to molecular length scale. Fluid properties can vary continuously from one volume
element to another and are average values of the molecular properties. The continuum hypothesis
can lead to inaccurate results in applications like supersonic speed flows, or molecular flows on
nano scale. Those problems for which the continuum hypothesis fails, can be solved using statistical
mechanics. To determine whether or not the continuum hypothesis applies, the Knudsen number,
defined as the ratio of the molecular mean free path to the characteristic length scale, is evaluated.
Problems with Knudsen numbers below 0.1 can be evaluated using the continuum hypothesis, but
molecular approach (statistical mechanics) can be applied for all ranges of Knudsen numbers.

Navier–Stokes equations[edit]
Main article: Navier–Stokes equations
The Navier–Stokes equations (named after Claude-Louis Navier and George Gabriel Stokes)
are differential equations that describe the force balance at a given point within a fluid. For

an incompressible fluid with vector velocity field , the Navier–Stokes equations are

.
These differential equations are the analogues for deformable materials to Newton's equations
of motion for particles – the Navier–Stokes equations describe changes in momentum (force) in

response to pressure and viscosity, parameterized by the kinematic viscosity here.


Occasionally, body forces, such as the gravitational force or Lorentz force are added to the
equations.
Solutions of the Navier–Stokes equations for a given physical problem must be sought with the
help of calculus. In practical terms only the simplest cases can be solved exactly in this way.
These cases generally involve non-turbulent, steady flow in which the Reynolds number is small.
For more complex cases, especially those involving turbulence, such as global weather systems,
aerodynamics, hydrodynamics and many more, solutions of the Navier–Stokes equations can
currently only be found with the help of computers. This branch of science is
called computational fluid dynamics.

Inviscid and viscous fluids[edit]


An inviscid fluid has no viscosity, . In practice, an inviscid flow is an idealization, one that
facilitates mathematical treatment. In fact, purely inviscid flows are only known to be realized in
the case of superfluidity. Otherwise, fluids are generally viscous, a property that is often most
important within a boundary layer near a solid surface,[2] where the flow must match onto the no-
slip condition at the solid. In some cases, the mathematics of a fluid mechanical system can be
treated by assuming that the fluid outside of boundary layers is inviscid, and then matching its
solution onto that for a thin laminar boundary layer.
For fluid flow over a porous boundary, the fluid velocity can be discontinuous between the free
fluid and the fluid in the porous media (this is related to the Beavers and Joseph condition).
Further, it is useful at low subsonic speeds to assume that a gas is incompressible—that is, the
density of the gas does not change even though the speed and static pressure change.

Newtonian versus non-Newtonian fluids[edit]


A Newtonian fluid (named after Isaac Newton) is defined to be a fluid whose shear stress is
linearly proportional to the velocity gradient in the direction perpendicular to the plane of shear.
This definition means regardless of the forces acting on a fluid, it continues to flow. For example,
water is a Newtonian fluid, because it continues to display fluid properties no matter how much it
is stirred or mixed. A slightly less rigorous definition is that the drag of a small object being
moved slowly through the fluid is proportional to the force applied to the object.
(Compare friction). Important fluids, like water as well as most gases, behave—to good
approximation—as a Newtonian fluid under normal conditions on Earth.[3]
By contrast, stirring a non-Newtonian fluid can leave a "hole" behind. This will gradually fill up
over time—this behaviour is seen in materials such as pudding, oobleck, or sand (although sand
isn't strictly a fluid). Alternatively, stirring a non-Newtonian fluid can cause the viscosity to
decrease, so the fluid appears "thinner" (this is seen in non-drip paints). There are many types of
non-Newtonian fluids, as they are defined to be something that fails to obey a particular
property—for example, most fluids with long molecular chains can react in a non-Newtonian
manner.[3]

Equations for a Newtonian fluid[edit]


Main article: Newtonian fluid
The constant of proportionality between the viscous stress tensor and the velocity gradient is
known as the viscosity. A simple equation to describe incompressible Newtonian fluid behaviour
is

where

is the shear stress exerted by the fluid ("drag")

is the fluid viscosity—a constant of proportionality

is the velocity gradient perpendicular to the direction of shear.


For a Newtonian fluid, the viscosity, by definition, depends only
on temperature and pressure, not on the forces acting upon it. If the fluid
is incompressible the equation governing the viscous stress (in Cartesian
coordinates) is

where

is the shear stress on the face of a fluid element in the direction

is the velocity in the direction

is the direction coordinate.


If the fluid is not incompressible the general form for the
viscous stress in a Newtonian fluid is

where is the second viscosity coefficient (or bulk


viscosity). If a fluid does not obey this relation, it is termed
a non-Newtonian fluid, of which there are several types.
Non-Newtonian fluids can be either plastic, Bingham
plastic, pseudoplastic, dilatant, thixotropic, rheopectic,
viscoelastic.
In some applications another rough broad division among
fluids is made: ideal and non-ideal fluids. An Ideal fluid is
non-viscous and offers no resistance whatsoever to a
shearing force. An ideal fluid really does not exist, but in
some calculations, the assumption is justifiable. One
example of this is the flow far from solid surfaces. In many
cases the viscous effects are concentrated near the solid
boundaries (such as in boundary layers) while in regions of
the flow field far away from the boundaries the viscous
effects can be neglected and the fluid there is treated as it
were inviscid (ideal flow). When the viscosity is neglected,

the term containing the viscous stress tensor in the


Navier–Stokes equation vanishes. The equation reduced in
this form is called the Euler equation.

See also[edit]

 Physics portal

 Aerodynamics
 Applied mechanics
 Bernoulli's principle
 Communicating vessels
 Computational fluid dynamics
 Corrected fuel flow
 Secondary flow
 Different types of boundary conditions in fluid
dynamics

Notes[edit]
1. ^ Batchelor (1967), p. 74.
2. ^ Kundu, P.K., Cohen, I.M., & Hu, H.H., Fluid
Mechanics, Chapter 10, sub-chapter 1
3. ^ Jump up to:a b Batchelor (1967), p. 145.

References[edit]
 Batchelor, George K. (1967), An Introduction to Fluid
Dynamics, Cambridge University Press, ISBN 0-521-
66396-2

Further reading[edit]
 Falkovich, Gregory (2011), Fluid Mechanics (A short
course for physicists) (PDF), Cambridge University
Press, ISBN 978-1-107-00575-4
 Kundu, Pijush K.; Cohen, Ira M. (2008), Fluid
Mechanics (4th revised ed.), Academic
Press, ISBN 978-0-12-373735-9
 Currie, I. G. (1974), Fundamental Mechanics of
Fluids, McGraw-Hill, Inc., ISBN 0-07-015000-1
 Massey, B.; Ward-Smith, J. (2005), Mechanics of
Fluids (8th ed.), Taylor & Francis, ISBN 978-0-415-
36206-1
 White, Frank M. (2003), Fluid Mechanics, McGraw–
Hill, ISBN 0-07-240217-2
 Nazarenko, Sergey (2014), Fluid Dynamics via
Examples and Solutions, CRC Press (Taylor & Francis
group), ISBN 978-1-43-988882-7

External links[edit]
Fluid mechanicsat Wikipedia's sister projects

 Media from Wikimedia Commons

 Quotations from Wikiquote

 Textbooks from Wikibooks

 Free Fluid Mechanics books


 Annual Review of Fluid Mechanics
 CFDWiki – the Computational Fluid Dynamics
reference wiki.
 Educational Particle Image Velocimetry – resources
and demonstrations

Superfluidity
From Wikipedia, the free encyclopedia

Jump to navigationJump to search


Helium II will "creep" along surfaces in order to find its own level—after a short while, the levels in the two
containers will equalize. The Rollin filmalso covers the interior of the larger container; if it were not sealed, the
helium II would creep out and escape.

The liquid helium is in the superfluid phase. A thin invisible film creeps up the inside wall of the cup and down
on the outside. A drop forms. It will fall off into the liquid helium below. This will repeat until the cup is empty—
provided the liquid remains superfluid.

Superfluidity is the characteristic property of a fluid with zero viscosity which therefore flows without
loss of kinetic energy. When stirred, a superfluid forms cellular vortices that continue to rotate
indefinitely. Superfluidity occurs in two isotopes of helium (helium-3 and helium-4) when they are
liquefied by cooling to cryogenic temperatures. It is also a property of various other exotic states of
matter theorized to exist in astrophysics, high-energy physics, and theories of quantum gravity.[1] The
phenomenon is related to Bose–Einstein condensation, but neither is a specific type of the other: not
all Bose–Einstein condensates can be regarded as superfluids, and not all superfluids are Bose–
Einstein condensates.[2] The theory of superfluidity was developed by Lev Landau.
Contents

 1Superfluidity of liquid helium


 2Ultracold atomic gases
 3Superfluid in astrophysics
 4In high-energy physics and quantum gravity
 5See also
 6References
 7Further reading
 8External links

Superfluidity of liquid helium[edit]


Main article: Superfluid helium-4
Superfluidity was originally discovered in liquid helium, by Pyotr Kapitsa and John F. Allen. It has
since been described through phenomenology and microscopic theories. In liquid helium-4, the
superfluidity occurs at far higher temperatures than it does in helium-3. Each atom of helium-4 is
a boson particle, by virtue of its integer spin. A helium-3 atom is a fermion particle; it can form
bosons only by pairing with itself at much lower temperatures. The discovery of superfluidity in
helium-3 was the basis for the award of the 1996 Nobel Prize in Physics.[1] This process is similar to
the electron pairing in superconductivity.

Ultracold atomic gases[edit]


Superfluidity in an ultracold fermionic gas was experimentally proven by Wolfgang Ketterle and his
team who observed quantum vortices in 6Li at a temperature of 50 nK at MIT in April 2005.[3][4] Such
vortices had previously been observed in an ultracold bosonic gas using 87Rb in 2000,[5] and more
recently in two-dimensional gases.[6] As early as 1999 Lene Hau created such a condensate using
sodium atoms[7] for the purpose of slowing light, and later stopping it completely.[8] Her team
subsequently used this system of compressed light[9] to generate the superfluid analogue of shock
waves and tornadoes:[10]
These dramatic excitations result in the formation of solitons that in turn decay into quantized
vortices—created far out of equilibrium, in pairs of opposite circulation—revealing directly the
process of superfluid breakdown in Bose-Einstein condensates. With a double light-roadblock setup,
we can generate controlled collisions between shock waves resulting in completely unexpected,
nonlinear excitations. We have observed hybrid structures consisting of vortex rings embedded in
dark solitonic shells. The vortex rings act as 'phantom propellers' leading to very rich excitation
dynamics.

— Lene Hau, SIAM Conference on Nonlinear Waves and Coherent Structures

Superfluid in astrophysics[edit]
The idea that superfluidity exists inside neutron stars was first proposed by Arkady Migdal.[11][12] By
analogy with electrons inside superconductors forming Cooper pairs because of electron-lattice
interaction, it is expected that nucleons in a neutron star at sufficiently high density and low
temperature can also form Cooper pairs because of the long-range attractive nuclear force and lead
to superfluidity and superconductivity.[13]

In high-energy physics and quantum gravity[edit]


Main article: Superfluid vacuum theory
Superfluid vacuum theory (SVT) is an approach in theoretical physics and quantum
mechanics where the physical vacuum is viewed as superfluid.
The ultimate goal of the approach is to develop scientific models that unify quantum mechanics
(describing three of the four known fundamental interactions) with gravity. This makes SVT a
candidate for the theory of quantum gravity and an extension of the Standard Model.
It is hoped that development of such theory would unify into a single consistent model of all
fundamental interactions, and to describe all known interactions and elementary particles as different
manifestations of the same entity, superfluid vacuum.
On the macro-scale a larger similar phenomenon has been suggested as happening in
the murmurations of starlings. The rapidity of change in flight patterns mimics the phase change
leading to superfluidity in some liquid states.[14]

See also[edit]
 Boojum (superfluidity)
 Condensed matter physics
 Macroscopic quantum phenomena
 Quantum hydrodynamics
 Slow light
 Superconductivity
 Supersolid

Advanced information

Additional background material on the


Nobel Prize in Physics 1996
The Royal Swedish Academy of Sciences has decided to award the 1996 Nobel
Prize in Physicsjointly to

Professor David M. Lee, Cornell University, Ithaca, NY, USA,

Professor Douglas D. Osheroff, Stanford University, Stanford, CA, USA, and

Professor Robert C. Richardson, Cornell University, Ithaca, NY, USA,

for their discovery of superfluidity in helium-3.


This additional background material gives a short account of the discovery and
its importance and is written mainly for physicists.

1. A breakthrough in low temperature physics


The pioneering work of David Lee, Douglas Osheroff and Robert Richardson in
the beginning of the 1970’s at the low-temperature laboratory of Cornell
University has given a most valuable contribution to our current view of the
manifestations of quantum effects in bulk matter. The anisotropic superfluid
helium-3, appearing below a critical temperature of about two thousandths of a
degree above the absolute zero, is considered to be a particular kind of Bose-
Einstein condensate with a rich set of physical properties. The study of this exotic
quantum liquid has led to concepts that are of general importance and, e.g., could
become useful for the theoretical treatment of high temperature
superconductors. Recently phase transitions in helium-3 have been studied as a
model for the dynamics of the cosmological phase transitions that are thought to
have occurred a fraction of a second after the Big Bang (see Nature, July 25, and
Science, August 2, 1996). Critical points of superfluid helium-3 are used to define
temperature scales at values extremely close to the absolute zero.

2. The discovery
Superfluidity in helium-3 first manifested itself as small anomalies in the melting
curve of solid helium-3, i.e. as small structures in the diagram representing
pressure against time, when the fluid was cooled. It is always tempting to
consider small deviations as more or less inexplicable peculiarities of the
equipment, but the discoverers became convinced that there was a real effect.
They were actually not looking for superfluidity, but for an antiferromagnetic
phase in solidhelium-3, which according to predictions was to appear below 2
mK. It was thus natural that they, in their first publication 1972, interpreted the
effect as the observation of such a phase transition.

The agreement was not perfect, but by further development of their technique
and new measurements they could, just a few months later, pinpoint the effect. It
actually turned out to involve two phase transitions in the liquid phase, at 2.7
and 1.8 mK respectively.

The discovery became the starting point of an intense activity among low
temperature physicists. The experimental and theoretical developments went
hand-in-hand in an unusally fruitful way. The field was rapidly mapped out, but
fundamental discoveries are still being made.
3. Particle statistics and superfluidity
In quantum physics the atoms in a gas are described by a wavefunction which is
a function of all the coordinates of the atoms, but which only specifies the
probability of finding a particle in a given region at a given time. In the quantum
regime (which applies at high density or low temperature) the
indistinguishability of the atoms leads to dramatic quantum effects. In nature
there are two fundamental types of particles, fermions and bosons. Fermions
have half-integral spin and are described by wavefunctions that are
antisymmetric in the exchange of two particles, i.e. the wavefunctions change
sign when two particles change places, and they follow what is called Fermi-
Dirac statistics. Bosons have integral spin and symmetric wavefunctions, i.e. their
wavefunctions are unchanged when two particles are exchanged, and they follow
Bose-Einstein statistics. Fermions tend to avoid each other and a gas of fermions
can have at most one particle in each one-particle quantum state. Bosons, on the
other hand, are more sociable and can occupy the same quantum state. Below a
certain temperature, which depends on the particle density, the bosons tend to
gather in a Bose-Einstein condensate in the quantum state of the lowest energy
and momentum. They are then described by one and the same wavefunction.

A pure Bose-Einstein condensate of (bosonic) atoms that only interact weakly


was not experimentally produced until last year. Then a number of groups
managed to cool small samples of dilute gases to temperatures well below one
microkelvin. But a type of Bose-Einstein condensate, with atoms condensed into
the ground state, was identified already in the 1930’s, namely the superfluid
phase of helium. This quantum liquid is freely flowing (without viscosity), can
penetrate fine pores that are closed for ordinary liquids and for many gases, can
creep upwards along walls and is an excellent heat conductor.

Helium mainly consists of the isotope helium-4, which is a boson (electronic and
nuclear spins are zero). The more rare isotope helium-3, on the other hand, has
nuclear spin 1/2, is a fermion and as such cannot undergo Bose-Einstein
condensation. But in explaining the phenomenon of superconductivity in metals
in 1957, Bardeen, Cooper and Schrieffer showed that fermions (in this case
electrons) under certain conditions can make up pairs (Cooper pairs) that
behave as bosons. These pairs can then undergo condensation to a ground state.
In principle this explains the 1972 finding of the phenomenon of superfluidity in
helium-3 by Lee, Osheroff and Richardson. But the nature of the pairing and the
properties of the pairs are very different in the two cases.

In a superconducting metal it is the surrounding lattice of positive ions that


provides the mechanism making it possible to pair together electrons with
opposite momenta and spin to quasiparticles having zero orbital angular
momentum or spin (L=S=0). In the superfluid phase of helium-3 the atoms
themselves provide the pairing interaction, through magnetic interaction (the
superfluid phase is almost ferromagnetic), and the pairs are more complicated.
The atoms in the pair rotate around each other and the pair has one unit of
internal orbital angular momentum (L=1). The nuclear spin magnetic moments
tend to be oriented along a common direction (S=1). The wave function which
describes the pair is a complex valued function and has both amplitude and
phase. This means that the wavefunction of a superfluid helium-3 pair has
2(2L+1)(2S+1)=18 real components, as compared to 2 for the superconducting
electron pair. Even though some components are coupled to each other (there is
a spontaneous breaking of the symmetry in spin-orbit space) the wave function
is still quite complicated and gives rise to a rich set of orientational effects.

In the condensate, the bosonic quasiparticle pairs are coupled to each other and
can be described by a macroscopic wave function with a well defined phase. This
means that the pairs, with their spinning nuclei and partners rotating around
each other, all move coherently so that their individual nuclear spins and orbital
angular momenta are coupled to a correlated state with large spatial extension.
Some consequences of this are that a minimum energy (gap energy) is needed to
break up the condensate, that the liquid cannot rotate freely above a critical
rotational velocity, but vortices appear with quantized circulation, and that
Josephson effects appear, e.g., leading to a kind of “ringing” in the liquid after the
variation of a magnetic field over the sample. Most of the theoretical concepts
regarding the paired state and the pairing mechanism were developed already
before the experimental discovery, by, among others, Anderson and Morel (later
on also with Brinkman), by Vdovin, and by Balian and Werthamer, and others.
Experiments on superfluid helium-3 have later on helped to discriminate among
different theories.

4. The experimental technique


Helium is an inert gas that is present as a small component in ordinary air (about
one part in 200 000). But the fraction of the isotope helium-3 is about one million
times smaller and it would be too costly to extract it out of air or out of ordinary
helium gas. Instead it can be produced by irradiation of lithium by neutrons from
a nuclear reactor. After the nuclear reaction and beta decay a gas rich in helium-3
is left, which is sold at a high price.

Both isotopes of He are inert and light gases, which among other things means
that their electronic dipole polarizabilities are small, thus making the van der
Waals interaction between individual atoms weak, but also that the zero point
motion is large. This implies that the condensed gas, liquid helium, does not
freeze at ordinary pressure, but remains in liquid form even at temperatures
close to the absolute zero. In this respect helium is unique among all the
elements of the periodic table. It is only under high pressure at low temperatures
that the liquid helium crystallizes and transforms into a solid phase.

Several powerful techniques for cooling were developed during the 1960’s. Lee,
Osheroff and Richardson used a method that had been proposed by
Pomeranchuk and which was put into practical use by Anufriev and later
developed by, among others, the scientists at Cornell. The method makes use of
the remarkable property of helium-3 that the liquid phase at low temperature is
more well-ordered than the solid phase. (Ordinary liquids are much more
disordered, have a higher entropy, than the corresponding crystals, with their
periodically ordered rows of atoms.) By applying a pressure to the liquid, some
parts of it are transformed into the solid phase. These parts thus transform from
a higher to a lower order, for which heat is needed (cf. the melting of an ordinary
crystal). This heat is taken from the remaining liquid, which thus is cooled
further.

Using Pomeranchuk cooling one can reach a final temperature just below 2 mK
before all the liquid has been transformed into the solid phase. The process is
hampered by not being continuous, but it has several positive properties. The
cooling power is high and the heat contact with the liquid helium-3 sample is
good, since the cooling medium is the same as the sample. At very low
temperatures it can otherwise be difficult to get a good heat contact; it can easily
happen that the cooling agent, the sample and the thermometer have different
temperatures. Different excitations (e.g., thermal motions among the atoms, spin
waves and electrons) may also not be in thermal equilibrium.

5. Discovery and properties of superfluid helium-3


The scientists at Cornell were low temperature specialists and had built their
own apparatus. But in their first measurements on helium-3 they had a problem
with their thermometer at below a few thousandths of a degree from absolute
zero. They decided to monitor the internal pressure in the sample under an
external pressure that increased uniformly with time. It was the research student
Osheroff who observed a change in the way the internal pressure varied with
time. He did not put the observation aside as being due to some feature of the
apparatus, but instead insisted that it was a real effect. He observed two
anomalies, shown in Fig. 1. They turned out to be the transition to phase A,
where the individual members of the boson pairs have parallel spins, and to the
phase B, in which they have both parallel and anti-parallel spins. (In a magnetic
field, phase A will increase at the expense of phase B, as seen in a pressure-
against-temperature diagram. Then also a new phase (A ) appears, in which the
1

pairs have atoms with parallel spins (as in phase A) and they all point in the same
direction.)

Another speciality of the group at Cornell was the nuclear magnetic resonance
technique (NMR). In an applied magnetic field the nuclear spins of the sample
atoms will rotate around the field lines. The frequency of rotation is given by the
strength of the field and by the magnetic moments of the nuclei. When the
frequency becomes equal to that of an applied radio frequency (r.f.) field,
resonance appears and the absorption of the r.f. field increases. This kind of
measurement gives valuable information on the magnetic state of the helium-3
nuclei. Lee, Osheroff and Richardson found characteristic changes of the
resonance frequency at the phase transition, changes that are dependent on the
magnetic field strength and on the temperature and are different in the A and B
phases. The theoretician Leggett could, within a few weeks, explain the
characteristic behaviour in detail. He showed that in each pair the nuclear spins
are coupled with the rotation, and pointed out the importance of the phase of the
macroscopic wave function that describes the condensate.

The fact that the new phases of helium-3 really were superfluid and could flow
without resistance was shown by two groups soon after the discovery. A group at
the University of Technology in Helsinki, led by Olli Lounasmaa, measured the
damping of a string vibrating in the liquid. They showed that the damping
diminished by a factor of about 1 000 as the liquid was cooled from above 2 mK
to 1 mK. The group led by the late John Wheatley at La Jolla detected and
measured the velocity of the so-called fourth order sound. This is not a pressure
or density wave, as in ordinary sound, but a temperature wave at constant
pressure appearing in fine pores. A persistent flow experiment in Helsinki
showed that the flow of superfluid helium-3 in a torus, with packed powder and
helium-3 inside, did not decay, at least on the scale of a few days, in the B phase
(but not in the anisotropic A phase). This implied a viscosity at least 12 orders of
magnitude smaller than the one in the normal fluid helium-3.

6. Research on superfluid helium-3 today


The most convincing experiments testing the coherence of a superfluid are
probably those showing the appearance of quantized vortices. When a superfluid
is set in rotation and the velocity of rotation exceeds a critical value, microscopic
vortices appear. The circulation around such a vortex cannot take on any
arbitrary value, but is quantized. This is known from “ordinary” superfluid
helium. In helium-3 the vortices can take on complicated appearances, in fact
eight different types of vortices have been seen with discontinuous or continuous
flow in the vortex cores. Each of them represents a novel topological object with
peculiar symmetry and structure. NMR, vibrating strings and other methods have
been applied to study the detailed structure of vortices. Their appearance can
even be observed directly, through an optical fibre and a cooled CCD camera, as
done by the Finnish group which looked on the surface of a rotating sample.

Another topical field of study is textures, similar to those appearing in liquid


crystals, with nuclear spins and orbital angular momenta pointing in different
directions in different domains of the liquid. The influence of boundary surfaces
on the orientation of the liquid, the nucleation and time dependence of phase
transitions are also studied.

The phase transitions in helium-3 have recently been used by two different
experimental groups (Grenoble and Helsinki) in attempts to simulate the
formation of cosmic strings in the early universe. These hypothetical strings
might have appeared as topological defects in the rapid phase transitions that are
thought to have broken the symmetry of the originally unified interaction and
given rise to the four fundamental forces as we know them today (strong,
electromagnetic, weak, gravitational). Both groups used neutron induced nuclear
reactions to heat their samples locally in such an abrupt way that the well
localised phase transitions were accompanied by vortex formation, these
vortices being the analogues of the cosmic strings. The validity of a theory
formulated by Zurek, following an idea by Kibble, thus seems to have been
confirmed. The cosmic strings are believed to be of importance, e.g., for the
formation of galaxies.

7. Summary
Superfluidity in helium-3 only appears at very low temperatures, below about 2
mK, and has found practical applications only for specialists in the extreme low
temperature techniques. Its main importance has been to develop our
understanding of the complicated behaviour of strongly interacting many-
particle quantum systems, such as quantum liquids, and for the development of
theoretical concepts in the field of macroscopic quantum phenomena. The
understanding of high temperature superconductivity, which is still not
complete, has gained from concepts developed for helium-3, giving examples of
the interactions that lead to pairing of particles in strongly interacting systems as
well as for the symmetry of the wave function for such pairs. As a practical
application, the polycritical point, where the superfluid phases A and B are in
equilibrium with the normal liquid phase, is also used as a fixed point to define
temperature scales at very low temperatures.
8. Further reading
“Superfluid Helium 3”, by N.D. Mermin and D.M. Lee, Scientific American,
December 1976, p. 56.

“Low temperature science – what remains for the physicist?”, by R.C. Richardson,
Physics Today, August 1981, p. 46.

Special Issue: He and He , Physics Today, February 1987, including among other
3 4

articles “Novel magnetic properties of solid helium-3”, by M.C. Cross and D.D.
Osheroff, p. 34.

“The He Superfluids”, by O.V. Lounasmaa and G.R. Pickett, Scientific American,


3

June 1990.

“The Superfluid Phases of Helium 3”, by D. Vollhardt and P. Wölfle, Taylor and
Francis 1990.

Figure 1. Time dependence of the internal pressure in a Pomeranchuk cell


containing a mixture of liquid and solid helium-3 under a cycle of uniform
compression and decompression. Note the change in slope of the curves at the
points A and B and the temperatures at which they appear. The curve is taken
from a paper published by D.D. Osheroff, R.C. Richardson, and D.M. Lee in
Physical Review Letters 28, 885 (1972), which gives the first description of the
new phase transition in helium-3.

To cite this section


MLA style: Advanced information. NobelPrize.org. Nobel Media AB 2019. Fri. 9 Aug 2019.
<https://www.nobelprize.org/prizes/physics/1996/advanced-information/>

WOMEN WHO CHANGED


SCIENCE
Back to top

Explore a new storytelling experience that celebrates and explores the contributions, careers and
lives of the 19 women who have been awarded Nobel Prizes for their scientific achievements.

You might also like