Vous êtes sur la page 1sur 8

VORTEX PHYSICS IN

HIGH-TEMPERATURE
SUPERCONDUCTORS
T

he discovery of high-temother factor, that will deterOur understanding of vortex matter in mine
perature superconducthe success of these
superconductors has grown dramatically technological applications in
tors1 has stimulated dramatic
growth
in
our
in the last decade, creating new horizons the long term. Many of the
understanding of the physics
potential applications dein fundamental science and potential
of quantized vortex lines.
mand high current densities
commercial applications.
These superconductors exwith minimal energy dissiclude magnetic fields weaker
pation, a requirement that
than a lower critical field
necessitates the immobiGeorge W. Crabtree and David R. Nelson lization, or pinning, of vorHcl < 10"2 tesla. Stronger
fields penetrate as an array
tices against the driving
of vortices, each consisting of
Lorentz forces induced by
exactly one quantum of flux
current flow. In earlier
(<o = hc/2e) surrounded in the plane perpendicular to the days, the knowledge needed to improve the performance
field by circulating supercurrents that extend radially a
of superconductors was generated primarily by materials
few hundred nanometers. The behavior of vortices domiscientists and engineers working empirically on specific
nates many physical properties of high-temperature suproblems. Now, a much more fundamental and compreperconductors up to the upper critical field Hc2 ~ 102 tesla,
hensive picture of vortex physics has emerged, to the point
where superconductivity gives way to normal metallic
where general principles developed in the last few years
behavior and magnetic fields penetrate uniformly.
can be used to understand and control vortex behavior.
Encompassing both fundamental science and commerWe know, for example, that vortex pinning is inticial technology, the behavior of vortices promises to permately linked to the layered structure of high-Tc cuprates
manently alter the way we think about superconductivity
(illustrated in the upper right of figure 2). The double
and its potential applications. Of all the new physics
copper oxide (CuO2) planes that contain most of the
generated by high-temperature superconductors, vortex
superelectrons and define the vortex properties are sepabehavior with its richness and complexity is perhaps the
rated by "blocking layers" of normal metallic or insulating
most unexpected and the most rapidly advancing.
material, which act as potential barriers against the tunVortices in superconductors provide a highly accessineling of superelectrons between planes. If the coupling
ble example of flexible one-dimensional objects interacting
between layers is weak compared to thermal energies, the
through a simple potential in the presence of controlled
vortex lines behave more like sheets of "pancake vortices"
disorder. These wandering lines of magnetic flux display
that interact strongly within each layer but only weakly with
complex equilibrium behavior reflecting liquid, crystalline
neighboring layers. Pancake vortices are difficult to pin
and glassy phases. In addition, they can be subjected to
because mobile pancakes in one layer can slip past pinned
an external driving force through the application of a
pancakes in adjacent layers. Thus, a key element in
transport current, allowing steady states of driven vortex
improving superconducting performance is strengthening
motion to be established and explored. Like the equilibthe vortex coupling between adjacent CuO2 double layers.
rium phases, these dynamic phases display remarkable
Ways of implementing this strategy include reducing the
complexity and subtlety, including several types of plastic
CuO2 layer separation, metallizing the blocking layers by
and elastic motion separated by novel dynamical phase
chemical substitution and introducing extended "columnar
transitions.
defects" that pin vortex lines across many superconducting
Technology has embraced the potential of high-templanes.
perature superconductors, with more than 15 companies
actively involved in developing products that range from
Equilibrium vortex phases
high-current bulk wires and tapes (see figure 1) to thin
The richly detailed equilibrium behavior of vortex phases
film devices. It is materials performance, more than any
in high-temperature superconductors arises from the competition of four energies:
D> Thermal energy favors a vortex liquid of lines or panGEORGE CRABTREE is a senior scientist in the materials science
cakes.
division ofArgonne National Laboratory in Argonne, Illinois,
D> Vortex interaction energy favors a perfect vortex lattice.
and theme leader for vortices and critical currents in the NSF
O
Pinning energy favors an amorphous or glassy solid.
Science and Technology Center for Superconductivity.
O Coupling energy between layers controls the formation
DAVID NELSON is Mallinckrodt Professor ofPlrysics and a
of vortex lines from weakly interacting pancake vortices
professor of applied physics at Harvard University in
in adjacent layers.
Cambridge, Massachusetts.
Remarkably, all four energies can be the same order
38

APRIL 1997

PHYSICS TODAY

1997 American Institute of Physics, S-0031-922S-9704-030-1

FIGURE 1. FLEXIBLE MULTIFILAMENT SUPERCONDUCTING WIRE made from Bi2Sr2CaCu2O8 (BSCCO) superconductor in silver
sheathing. The wire is made by the powder-in-tube process, in which superconducting powder is put into an array of holes in
a silver rod (shown in the center of the picture), which is then rolled and heat treated to produce a flat, flexible wire with hundreds or thousands of filaments. (Photo courtesy of American Superconductor Corp.)
of magnitude in high-temperature superconductors, leading to an unprecedented variety of liquid and solid phases
and transitions among them. The term "vortex matter"
emphasizes that vortices are comparable in complexity
and diversity to conventional atomic or molecular matter.
Vortex matter provides an excellent laboratory for
exploring general phase-transformation behavior. Experimentally, all the relevant parameters can be varied over
wide ranges: the areal vortex density over many orders
of magnitude by changing the magnetic field; thermal
fluctuations through the temperature; the pinning disorder through controlled irradiation; and the coupling energy
between CuO2 planes through the choice of material.
Furthermore, vortex interactions with each other and with
an external driving current are known rather precisely,
allowing one to do analytical theory and numerical simulations with essentially no uncontrolled approximations.
This combination of accessibility, diversity of phases and
simplicity of interaction is difficult to find in conventional
matter composed of atoms and molecules. Qualitatively
new physics arises, moreover, because the basic degrees
of freedom can be lines instead of points.
Figure 2 is a generic phase diagram for single-crystal
cuprate materials in the absence of pinning. The basic
layered structure of the cuprates is illustrated in the upper
right. For simplicity, we assume a field oriented along
the c-axis, perpendicular to the CuO2 planes, which contain the a and b directions. The magnetic induction B,
which is proportional to the vortex density in the ab plane,
is plotted because demagnetizing corrections in the usual
platelet crystals ensure that B (rather than the magnetic
field H) is held fixed in most experiments. The famous
Abrikosov vortex lattice, which exists at all temperatures

below the upper critical field Bc2(T) in mean field theory,


appears here only at much lower temperatures, below a
line of melting transitions Tm(B). In the presence of strong
thermal fluctuations, the red curve Bc2(T) marks the onset
of enhanced diamagnetism but is not expected to be a
sharp phase transition.
These two curves bound a novel 'Vortex liquid1' regime
of flexible, entangled lines. Although the long-wavelength
properties in this region are qualitatively indistinguishable from those of a normal metal (there is a linear ohmic
resistivity, for example), the quantization and entanglement of vortices lead nevertheless to many important
experimental consequences.
The phase diagram of figure 2 is further subdivided
by a "decoupling line" Td(B)y above which the discreteness
of the copper oxide planes becomes significant.2 Above
Td(B), thermal fluctuations overcome the interplanar coupling, causing the vortex segment between adjacent planes
to wander a sideways distance comparable to the average
interline separation. The linelike nature of the vortices
is then less important: These fluxons behave more like
weakly coupled sheets of pointlike pancake vortices, although conservation of the total magnetic flux requires
conservation of the number of pancakes in adjacent layers.
Although there are important qualitative changes at Td(B),
there is no symmetry reason in the vortex liquid phase
why the curve Td{B) must be a locus of sharp phase
transitions, just as there need not be a sharp phase
transition line separating the liquid and gas phases of
ordinary matter above the critical point.
Remarkably, the extension of Td(B) into the crystalline
phase is in fact associated with a sharp phase transition
to a "supersolid" phase due to defect proliferation.3 The
APRIL 1997

PHYSICS TODAY

39

FIGURE 2. EQUILIBRIUM

PHASE DIAGRAM for clean


superconductors, with the
magnetic field perpendicular to
the C u O : planes. The layered
structure for BSCCO,
composed ol double
superconducting C u O : planes
and blocking layers with the
composition BijSnO^ is shown
at the upper right. The vortex
configurations in the
disentangled, entangled and
decoupled liquid phases are
sketched in the vortex liquid
region. The cage model of the
vortex lattice is illustrated at
the lower left. With B (rather
than H) as the vertical axis, the
Meissner phase and the lower
critical field are compressed
onto the B = 0 line.

z
o
y

_J

Q
Z
U

w
Z

TEMPERATURE 7

relevant defects are linelike interstitials and vacancies.


Such imperfections have energies and entropies proportional to their lengths, and their free energies are positive
at low temperatures, so they cannot extend completely
across an equilibrated macroscopic sample. This is the
case in the crystalline phase below Td(B), where vacancies
and interstitials will appear only in thermal equilibrium
as small, tightly bound loops, each composed of a vacancy
and interstitial string bound together. Above TdiB), however, the entropy of a wandering interstitial (or vacancy)
line is sufficient to cause its free energy to become negative. Such defects can then become arbitrarily long and
extend completely across a macroscopic sample. The conservation of total magnetic flux (and therefore the total
number of vortex lines) in each layer of the superconducting sample is essential for the existence of this sharp
proliferation transition of line defects, which resembles
the proliferation of the vortex lines themselves at Hcl.
The crystalline phase above Td(B) is called a supersolid
by analogy with a phase that was hypothesized, but never
observed, for quantum crystals of helium-4 in the 1960s.

Melting of the vortex crystal


Although melting of the vortex lattice is always expected
near the upper critical field on general grounds, it occurs
unobservably close to Bc2(T) in many low-temperature
type II superconductors.2 In high-temperature superconductors, however, there is sufficient thermal energy to
induce melting well below Tc. The linelike nature of the
vortices making up the solid and liquid phases brings new
features not seen in pointlike atomic solids. Many of the
salient features of line melting can be appreciated from
a simple model of line segments fluctuating independently,
40

APRIL 1997

PHYSICS TODAY

as in the classical Einstein model of phonons in an atomic


solid (see box 1 on page 41). As shown in the box, the
melting temperature depends dramatically on the flexibility of the lines, which is described by the effective mass
anisotropy mjm of the superconducting electrons. The
weak coupling between double CuO2 layers in
Bi2Sr2CaCu208 (BSCCO) leads to highly flexible vortices
with anisotropies mz/mL of 100 000 or more; these vortices
melt at much lower temperatures than the more strongly
coupled vortices in YBa2Cu307 (YBCO), which have an
anisotropy of about 50. The unification of the phase
diagrams of all layered superconductors in a single picture
that includes melting and decoupling is a major challenge
for vortex physics.
Theoretically, it is expected that freezing to a perfect
lattice would occur by way of a first-order transition in
clean systems,4 while freezing to a glassy solid would be a
more gradual process, possibly accompanied by a sharp
second-order transition in systems with quenched disorder.2
Vortex melting was first observed by Hugo Safar and
his coworkers in transport, experiments on YBCO as a
series of small reproducible jumps in the resistive transition accompanied by hysteresis in the temperature dependence.5 (See PHYSICS TODAY, October 1992, page 17.)
The most perfect crystals show dramatic melting behavior,
as illustrated in figure 3a, where a sharp drop to zero
resistivity occurs along a well-defined line in the B-T
plane.6 The drop to zero resistivity is accompanied by the
onset of hysteresis expected for a first-order transition,
and of nonlinear current-voltage curves characteristic of
pinning in the solid vortex state. At high magnetic fields,
the transition is broader and the hysteresis disappears,
suggesting that there may be a critical point where the

Box 1. Cage Model of Flexible Line Melting

65

70

75
80
85
TEMPERATURE (kelvin)

90

95

[r(z)] =

b
0.3

NORMLAJLIZED RESIS

he melting of a crystal of flexible vortex lines due to


thermal agitation may be understood by using a simplified cage model. Intervortex interactions are described by a
harmonic potential, which confines each line to the interior
of a cage defined by the surrounding near-neighbor lines (see
the illustration in the lower left of figure 2). The free energy
and various other configurational averages can be estimated
by breaking the lines up into independent segments just long
enough to allow exploration of the interior of the cage.
Melting occurs when the root-mean-square line fluctuations
grow to a fixed fraction of the equilibrium lattice constant,
beyond which confinement is no longer effective.
We denote the trajectory of a vortex as it traverses a
sample with a magnetic field parallel to z by r(z) = [x(z), y(z)]
and assume that the probability of a particular configuration
is proportional to exp(E[r(z)]/k%T)> where the energy of
the line in its confining cage of length L is

Ji (**W ] i

(1)

The coupling constant g is a local tilt modulus, which


measures the energy of bending the line away from the
external magnetic
field direction. The curvature of the
g

Pre-irradiation

hharmonic
i
fii potential
il k
confining
k ~dd2V/dr

0.2 -

0.1 -

^Post-irradiation

\.
Post-anneal

S\

n
0.89

0.94

0.90
0.91
0.92
0.93
NORMALIZED TEMPERATURE T/Tc

358.9

Ur,

w h e r e aQ =

( / ) 1 / 2 is the vortex lattice constant and P(r) describes


pairwise interactions between vortices. The lines interact
through a screened Yukawa potential, V(r) = 2 e0 KQ(r/\),
where e0 is an interaction energy per unit length and A is the
London penetration depth in a plane perpendicular to z.
Although K0(p) decays exponentially to zero for p 1 , it is
approximately equal to -In p for p = atf'k < 1, which is the
case for most of the experimentally relevant field range.
Thus, k ~ eQ/al, where we neglect constants of order unity
in this back-of-the-envelope discussion. In the same range of
fields, one has g = {mJmz) o> where mjmz is the effective
mass anisotropy of the superconducting electrons. As discussed in the body of the article, mjmz <^C 1 in the highly
anisotropic high-temperature superconductors.
Assume that the vortex requires a distance z = / to sample
the interior of its cage. If the transverse deflection associated
with this motion is r, the energy of this segment is, according
to equation 1, just ~ Vi (g/l + kl)r2. Upon optimizing
over /, we find a preferred segment length /* = vg/k with
energy E* = \gk r2. Application of the equipartition theorem to one such segment then leads to

ao

358.5
43.15

43.20
43.25
43.30
TEMPERATURE (kelvin)

43.35

FIGURE 3. VORTEX LATTICE MELTING, a: Resistivity of a

crystal of strongly coupled YBa2Cu3O7 (YBCO) showing a


sharp drop at the melting transition and rounded behavior at
high fields, which suggests a critical point.6-7 b: Resistivity of
YBCO before and after electron irradiation to produce
pointlike defects, and after annealing to remove the defects.8
The sharp pre-irradiation transition at melting is replaced by a
broad transition without hysteresis in the disordered crystal.
After annealing, the sharp first-order transition (and hysteresis)
is recovered, c: Discontinuity at melting in the local
magnetic induction of a crystal of weakly coupled BSCCO.9

According to the Lindemann criterion, melting occurs when


<\r2y > cL2 a02, where the Lindemann constant cL is typically
0.10-0.15. The melting temperature predicted by this argument is thus
1/2

1/2

Note that the slope of the melting curve Tm(B) predicted by the
Lindemann formula (equation 2) is indeed negative: At higher
fields the spatial fluctuations required for melting are reduced.
The soft, long-range nature of the logarithmic repulsion between
vortex lines is crucial: For potentials of the form V(r) \/rn,
a conventional melting curve with dTm/dB > 0 (and a denser
crystal phase) appears whenever n > 2.

APRIL 1997

PHYSICS TODAY

41

when pinning dominates. It is still unclear whether point


disorder alone is sufficient to lead to a true second-order
phase transition.
Although transport measurements are effective for
locating the melting line in the B-T plane, they give no
information about the line's thermodynamic character.
Thermodynamic measurements of the magnetization
change and latent heat at melting were long considered
futile because the effects were thought to be too small to
observe. Recently, this pessimistic outlook has been dramatically revised by sophisticated magnetization and calorimetric experiments showing tiny but measurable
changes in the magnetization and entropy at melting.9"12
Figure 3c shows the jump in the local
magnetic induction B on vortex melting in BSCCO, implying a first-order
change in the vortex density at the
melting transition.9 From thermodynamic data like these, entropy jumps
in BSCCO and YBCO have been derived by using the Clausius-Clapeyron equation. Recently, the first calorimetric experiment that directly
measures the latent heat TAS on vortex melting was reported for YBCO,
finding excellent agreement with the
entropy derived from the magnetic
experiment.12
The magnetization and calorimetric data are a tour de force of experimental technique and dramatically
confirm the existence of the thermodynamic first-order melting first
suggested by theory and transport
measurements.
The thermodynamic measure0
0.88
0.92
0.96
l
ments reveal a remarkable feature in
the melting curve of YBCO and
NORMALIZED TEMPERATURE 7/7.
BSCCO: The vortex density is higher
in the liquid than in the solid. Thennodynamicalry, this
FIGURE 4. TRANSFORMER EXPERIMENT, a: Geometry of the
"icelike" melting is equivalent to a negative slope of the
experiment, in which voltage differences on the top and
bottom surfaces probe the coherence between layers. Vtop and melting curve, as is evident from the Clausius-Clapeyron
equation, dHJdT = -AS/AM, where AS and AAf are the
^bottom count the number of vortices per unit time crossing
entropy and magnetization discontinuities, respectively, at
between the voltage taps on the top and bottom surfaces,
a first-order magnetic transition. If the slope of the
respectively, b: Measured voltages on the top and bottom
melting curve is negative, then the thermodynamic relayers of crystals of YBCO in the transformer geometry.13 In
quirement of greater entropy in the higher temperature
crystals without defects, the top and bottom voltages merge
liquid phase requires a larger magnetization as well,
only at the melting temperature Tm marked by the sharp drop
equivalent to a higher density of paramagnetic vortex
in both voltages to zero, indicating that the liquid phase is
lines.
entangled over nearly its whole range. In crystals with planar
The physical origin of icelike melting is the flexibility
pinning defects (called twin boundaries), the top and bottom
of the vortex lines combined with their long-range intervoltages merge at 7th in the liquid phase, indicating a disentangled
actions (see box 1). Vortices at typical densities interact
liquid regime between 7^ and Tx where the (presumably glass}'')
with a logarithmic repulsive potential, a remarkably weak
vortex solid forms and the voltage vanishes.
spatial dependence, which ensures that the slope of the
melting curve is negative and that the melting is icelike.
line of first-order transitions terminates.' Critical points
are possible because pointlike disorder destroys translational order in the crystal at the longest length scales,
removing one sharp symmetry difference between solid
and liquid phases.2
Introducing controlled point disorder (by means of
electron irradiation, for example8) has dramatic consequences: The sharp transition and hysteresis disappear
at all fields, giving way to a broad transition to zero
resistivity, as illustrated in figure 3b. This behavior
qualitatively confirms the theoretical picture of first-order
transitions to relatively perfect vortex lattices in clean
materials and more gradual transitions to a glassy phase
42

APRIL 1997

PHYSICS TODAY

Entanglement and decoupling in liquid phase


Above the melting line in figure 2, interactions (and
pointlike disorder) are less important, and we must consider a liquid of weakly confined lines. The mean square
distance that a line wanders in a distance z along the
field direction can be estimated by using equation 1 in
box 1 (the cage potential is ignored by using a setting of
k = 0). In this approximation, the line wandering is given
by a diffusion-like expression
<|r(z)-r(0)|2> = - j - | z | .
The line "diffuses" as a function of the timelike variable

Box 2, Vortex Statistical Mechanics and the Boson Mapping


he statistical mechanics of line vortices offers many opportunities to exploit rich analogies between spatial configurations of interacting lines in equilibrium and the non-relativistic quantum mechanics of interacting bosons in two
dimensions.17 One conclusion of the boson analogy is that
the reduction in quantum zero-point energy that accompanies
the melting of quantum Wigner crystals is represented by the
increased "braiding" entropy of the wiggling lines in the vortex
liquid. Thus, the same physics that drives the melting of the
quantum Wigner crystal appears classically in the macroscopic
vortex system.
Space does not permit a detailed exposition of the boson
mapping. However, a brief qualitative outline illustrates the
flavor of the argument. It can be shown that the classical
partition function Z of a system of N flexible vortices with
pairwise interaction V(r,j) and subject to an arbitrary z-independent pinning potential U(r) in a sample of thickness L is
related to a quantum mechanical matrix element of an TV-particle system propagating in imaginary time:

2 = Jdr 1 ...dr A Jdr / 1 ...dr / N

where H w = - 5 X ;

XA>)

Z ( l i / l )

Here, g is the vortex tilt modulus discussed in box 1. The


symmetrical integration over the vortex entry and exit points
{r;} and {r';} ensures that only boson states contribute to 2.
The states in the quantum mechanical matrix element in the
above equation for 2 define the vortex positions at the two
surfaces normal to the field direction, and the operator in the
matrix element is the exponential of a "Hamiltonian" that is
identical in form to that of correlated electrons in metals. The
boson mapping thus allows decades of work on correlated
electron physics to be transferred to research on vortex lines,
subject to the exchange of fermi for boson statistics.
The boson analogy can be exploited experimentally by
measuring the positions of vortices on opposite surfaces of a
superconducting slab, thus defining the states of the quantum
mechanical matrix element discussed above. Vortex positions
are traditionally probed with decoration experiments, in which
small magnetic particles are allowed to settle on the sample
surface, marking the relatively high-field regions in the vortex
cores. Double-sided vortex decoration experiments,18 leading
to the triangulated configurations of entry and exit points
shown in the inset figure, allow information about vortex

2, with "diffusion constant" D = kBT/gy where g is the vortex


tilt modulus. Once the line wanders a distance equal to
the intervortex spacing a0, the system is said to be "entangled," in the sense that its behavior can now be strongly
influenced by the barriers to cutting and reconnecting
vortex lines. The entanglement length /, is the distance
along the field required for the vortex line to wander
approximately one intervortex spacing a0:
tzBl
2

nBli3

where a0 ~ (f)0/B and <f)0 is the flux quantum.


Three regimes of entanglement can be defined by
comparing lz with characteristic lengths in the sample.
The disentangled regime occurs for low fields such that
the crystal thickness L is smaller than the entanglement
length: B <Bxl=gcj>QILkBT. Here, the lines are disentan-

DOUBLE-SIDED
TRIANGULATION of

vortex entry (a) and


exit (b) positions from
a decoration of
BSCCO at low
magnetic fields.18
Shaded triangles touch
"defect" sites, which
are five- or
seven-coordinated.
Dislocations (clusters
of red triangles) appear
as (5-7)-pairs in this
representation. The
green triangles
represent highly
disordered regions,
while the violet cluster
is a "twisted bond."
The two images are
remarkably similar,
despite the sample
thickness of
approximately 20 /xm.
The small differences
contain information
about vortex
correlations deep
inside the sample.

positions deep in the superconductor to be inferred from their


positions on the entrance and exit surfaces. The fluctuating
vortex trajectories act like a spatial filter transmitting information across the sample. The decay rate of spatial correlations
in the direction of the field as a function of sample thickness
is governed by the "phonon-roton" spectrum of the equivalent
superfluid boson system, and it displays the same minimum at
finite wavevector in vortex systems18 as is familiar from liquid
helium-4. Much useful quantitative and qualitative information can be obtained from the double-sided decorations, including the long-wavelength elastic constants, the decay of
crystalline and hexatic order across the crystal and the crossover
from logarithmic to exponential vortex pair interactions.

gled and the properties are those of a fluid of short


vortex segments t h a t easily slide past each other. Note
t h a t Bxl > 0 in the limit of infinite sample thickness.
When B > v l , the vortices entangle, and moving one
vortex independently of the others requires flux cutting.
Entanglement continues to increase until B exceeds Bx2 =
g4>0/skBT, which is the field such that lz = s, the spacing
between the double CuO 2 planes. Here, the flux lines are
"superentangled," or decoupled, in the sense that vortices
wander more than an intervortex spacing while passing
from one set of CuO 2 planes to the next. In this regime
the flux liquid can be viewed as planes of weakly interacting pointlike pancake vortices. The disentangled, entangled and decoupled regimes are illustrated in figure 2.
Entanglement of vortex lines in the flux liquid can
be probed through an elegant pseudo-transformer experiment. 13 As illustrated in figure 4a, current enters and
APRIL 1997

PHYSICS TODAY

43

FIGURE 5. DYNAMIC PHASE DIAGRAM at constant magnetic

Solid

Liquid

field ol steady-state motion of vortices in a random array of


pins in two dimensions, derived from numerical molecular
dynamics simulations.13 The vertical line marks the melting
temperature of the perfect vortex lattice and is the upper limit
for elastic motion in the disordered solid. The dashed line
indicates a crossover from strong pinning to plastic motion.
J

leaves by way of the top layer of the sample, and the


voltage is monitored across two symmetrically placed pairs
of taps on the top and bottom surfaces. Because of the
anisotropic geometry, the current density remains higher
near the top than the bottom, producing a stronger Lorentz
force driving the vortices near the top. The voltage difference across a pair of taps on the top or bottom directly
measures the number of vortices per unit time passing
between them. A voltage difference between the top and
bottom taps implies that the vortices on the top and bottom
move at different velocities, which requires flux cutting
and reconnecting. Such cutting and reconnection should
be particularly easy in the decoupled limit of independent
sheets of pancake vortices. If, however, the vortex lines
are disentangled, or the barriers to flux cutting (and
subsequent reconnection) are sufficiently high, then the
moving lines will remain intact in the field direction and
no voltage difference between the top and bottom taps
will be observed.
Figure 4b shows the voltages across the top and
bottom taps of YBCO crystals with and without the planar
pinning defects called twin boundaries. 13 The sharp drop
in both voltages in the untwinned samples at Tm marks
the first-order freezing transition. The two voltages are
different throughout the flux liquid phase, indicating entanglement and low barriers to flux cutting.
The behavior of samples with twin boundaries is even
more interesting.
In these crystals the planar twin
boundaries run parallel to the field direction and at 45
to the current direction. Such defects pin vortices very
strongly at low temperatures and occur in sufficient densities to destroy the vortex lattice. Both the top and
bottom voltages now go smoothly to zero at a lower
temperature 71,, suggesting a continuous transition to a
glassy state. Remarkably, however, the voltages at the
top and bottom of the sample become equal at a temperature T th , which is well above the point where the resistance
vanishes. This is also the temperature at which dissipationless supercurrents begin to flow along the c-axis. The
simplest interpretation is that the correlated disorder
embodied in the twin boundaries has "combed the hair"
of the flux lines sufficiently to produce a disentangled flux
liquid in the temperature interval T} <T < Tth.
Is this disentangled vortex liquid a distinct thermodynamic phase of the infinite system, 2 or is it a finite-size
effect caused by the entanglement length exceeding the
sample thickness? An answer to this question is suggested
by the behavior of the temperatures T, and T th as a
function of sample thickness. 13 T, is approximately independent of sample thickness, indicating a true phase
transition, presumably to a glassy state where the vortices
are strongly pinned by the twin boundaries. 7\ h , however,
is thickness dependent, as would be expected if it marked
the equality of the entanglement length and the sample
thickness. Tth appears to approach Ty with increasing
sample thickness, suggesting that the disentangled liquid
does not occur in an infinite sample. The diverging
entanglement length as T > T^ is consistent with a theory
of correlated pinning, which predicts a continuous transition to a u Bose glass" phase at low temperatures in the
44

APRIL 1997

PHYSICS TODAY

Elastic
motion

m
j

0
MH

Plastic motion

TEMPERATURE
presence of twin boundaries. 2 Wandering vortex lines
have many other important consequences that can be
explored with an analogy between vortex trajectories and
the world lines of quantum mechanical bosons, as sketched
in box 2 on page 43.

Dynamic response of vortices


The richness of fundamental behavior in the dynamic
response of vortices matches that of their equilibrium
phase diagram and is responsible for the recent and rapid
emergence of vortex dynamics as an independent subject
of investigation. Experimenters and theorists are now
developing tools that are sensitive to the driven motion
of vortices, as well as to their equilibrium configurations.
It is easy to set up a steady state of driven motion in a
finite sample because vortices are naturally created and
destroyed as necessary at the sample boundaries. Vortices
display an impressive range of dynamical behavior, including avalanches, stick-slip dynamics, thresholds for
motion, nonlinear and hysteretic dynamic response and
plastic/elastic solid motion. Vortex dynamics promises to
be a major growth area for its impact on superconducting
properties and for the insight it provides into driven
dynamical systems in general.
An important early step in dynamical vortex studies
is the characterization of the various steady states of
motion in a dynamic phase diagram,14 illustrated in figure
5. In the presence of pinning and at low temperature,
there is a rest phase at low driving force and a moving
phase above a critical de-pinning force. The dashed line
in figure 5 indicates the crossover from rest to motion.
The nature of the motion above the critical force shows
wide variation. In systems with random pinning, the
motion is nearly always plastic at driving forces just above
the critical force; that is, the vortices move at different
velocities in different parts of the system.15 This creates
rifts along certain planes, where the velocity changes
discontinuously. Numerical simulations show that the
plastic motion can take the form of rivers of moving
vortices sliding between stationary "river banks," with the
rivers sometimes only as wide as a single intervortex
distance.15 At higher driving force, a remarkable transition to elastic motion occurs, and the vortices all move
with the same average velocity14 (There may be local
elastic distortions that heal as the motion proceeds.) The
same elastic forces between vortices that create the equilibrium lattice stabilize this dynamic state. Intuitively,
such a transition seems reasonable, because at sufficiently

ROWER AMPLIFIERS

high driving force the pinning forces disrupting uniform


motion become insignificant. Numerical simulations suggest that this kind of dynamic phase transition is sharp
in two dimensions,14 and neutron diffraction experiments
confirm that the correlation length for translational order
in moving three-dimensional lattices increases significantly at high velocity.16
As the phases in the dynamic phase diagram are
identified and understood, attention is turning naturally
to the transitions between them. Experimentally, examples of sharp (and hysteretic), as well as smooth, dynamic
transitions have already been found. It is tempting to
search for a scheme to analyze dynamic transitions by
order parameters and by the divergence or discontinuity
of appropriate correlation lengths, adapting the concepts
that have been successful in treating equilibrium phase
transitions. Some progress has been achieved with the
identification of symmetry elements that define the plastic
and elastic dynamic states. Nevertheless, the analysis of
dynamic vortex phase transitions with concepts adapted
from equilibrium analogs is largely unexplored. The direction and the success of such analysis schemes constitute
one of the fascinating questions in vortex physics whose
answers await us in the future.
George Crabtree's work is supported in part by the US Department
of Energy. David Nelson's work is supported by the National
Science Foundation.

References
1. J. G. Bednorz, K. A. Muller, Z. Phys. 64, 189 (1986).
2. G. Blatter, M. V. Feigel'man, V. B. Geshkenbein, A. I. Larkin,
V. M. Vinokur, Rev. Mod. Phys. 66, 1125 (1994).
3. V. B. Geshkenbein, M. V. Feigel'man, A. I. Larkin, Physica C
167, 177 (1990). E. Frey, D. R. Nelson, D. S. Fisher, Phys. Rev.
B 49, 9723(1994).
4. E. Br6zin, D. R. Nelson, A. Thiaville, Phys. Rev. B 31, 7124
(1985).
5. H. Safar et al, Phys. Rev. Lett. 69, 3370 (1992).
6. W. K Kwok et ai, Phys. Rev. Lett. 72, 1092 (1994).
7. H. Safar et al, Phys. Rev. Lett. 70, 3800 (1993).
8. J. A. Fendrich^a/., Phys. Rev. Lett. 74, 1210 (1995).
9. E. Zeldov, D. Majer, M. Konczykowski, V. B. Geshkenbein,
V. M. Vinokur, Nature 375, 373 (1995).
10. R. Liang, D. A. Bonn, W. N. Hardy, Phys. Rev. Lett. 76, 835
(1996). U. Welp, J. A. Fendrich, W. K. Kwok, G. W. Crabtree,
B. W. Veal, Phys. Rev. Lett. 76, 4809 (1996).
11. H. Pastoriza, M. F. Gofftnan, A. Arribere, F. de la Cruz, Phys.
Rev. Lett. 72, 2951 (1994). T. Hanaguri et al., Physica C 256,
111(1996).
12. A. Schillings a/., Nature 382, 791 (1996).
13. D. Lopez et al, Phys. Rev. B 53, 8895 (1996). D. Lopez, E. F.
Righi, G. Nieva, F. de la Cruz, Phys. Rev. Lett. 76, 4034 (1996).
14. A. E. Koshelev, V M. Vinokur, Phys. Rev. Lett. 73, 3580 (1994).
T. Giamarchi, P. Le Doussal, Phys. Rev. Lett. 76, 3408 (1996).
15. F. Nori, Science 271, 1373 (1996). M. J. Higgins, S. Bhattacharya, Physica C 257, 232 (1996).
16. P. Thorel, R. Kahn, Y. Simon, D. Cribier, J. Physique 34, 447
(1973). U. Yaron et al, Nature 376, 753 (1995).
17. D. R. Nelson, H. S. Seung, Phys. Rev. B 39, 9153 (1989).
18. S. Yoon, Z. Yao, H. Dai, C. M. Lieber, Science 270, 270 (1995).

APPLICATION
Uhrasound/EMI
RF Power Source
Plasma / Heating
NMR / MRI

FREQUENCY [Hz]
20k-400 M
50 k ~ 800 M
13.65/27.12/40.86 M
5 M - 800 M

POWER |\VJ

10-3 k
100 - 10 k
100-5 k
50-3 k

POST AMPLIFIERS T142- Series


BAND
POWER
GAIN
FLATNESS
H.D.
INPUT

3022AA
100k~400M
1 W
30 dB
0.5dB
-20dB
+ 5 dB MAX

4029A
100k~350M
10W
40 dB
ldB
-20dB
-f- 5 dB MAX

4749A
1M-250M
50 W
47 dB
1.5dB
- 2 0 dB
+ 5dBMAX

5059A
5M-220M
100W
50 dB
I.5dB
-20dB
+ 5dBMAX

CW / Pulse Operation
Stable for Any Load
Wide Band
Low Distortion
Compact Design

s* <tfeedcustomized^instruments?

Cattus!

Worldwide Sales - Sci Tran Products / 1734 Emery Dr., Allison Park, PA
15101 USA/Phn: +1(412)367-7063/ Fax: +1(412)367-8194/e-mail:
sci@pgh.net/ Headquarter - Thamway Co., Ltd. / 3-9-2, Imaizumi, Fuji-shi,
Shizuoka-ken 417 JAPAN/ Phn: +81(545)53-8965/ Fax: +81(545)53-8978
Circle number 20 on Reader Service Card

MAGNETIC FIELD
MEASUREMENT
3 Axis Fluxgate Magnetometers
Measure magnetic fields from
2 x l O n T t o 10 4 T
Highly stable - offset drift
less than 1 0 9 T / C
Small probe size - 3 orthogonal
axes in probe size of r f xl"x2.625"
Simultaneous analog outputs
for all axes
Offset Capability:
The APS520A also has
precision 3 axis offset
capability with a full
scale range of + l x l 0 4 T

APS520A

APS520

Call u s to discuss your applications!


Applied Physics Systems
897 Independence Avenue #1C
Mountain View, CA 94043

415-965-0500
Fax 415-965-0404

Circle number 21 on Reader Service Card


APRIL 1997

PHYSICS TODAY

45

Vous aimerez peut-être aussi