PHYS40451 LectureNotes2023
PHYS40451 LectureNotes2023
Andrei Golov
September 2023
Novel Phenomena
p2 3
= kB T.
2m 2
At low temperature when their de Broglie wavelength,
h
λdB = ∼ h(3mkB T )−1/2 ,
p
exceeds the average distance between them, n−1/3 , particles’ individual trajec-
tories lose sense and we have to describe the gas in terms of a many-particle
1
wave function. This characteristic temperature of “quantum degeneracy” is thus
h2 n2/3
T∗ ∼ . (1.1)
3mkB
Below T ∗ the quantum statistics (whether bosons or fermions) is important.
Bosons make “Bose-Einstein condensate”; fermions make “Fermi sea”.
The particle’s total spin, sh̄, defines the statistics. Bosons (photon, phonon,
4
He, 1 H, 23 Na, 87 Rb) have s integer. Fermions (quark, electron, neutron, proton,
3
He, 86 Rb) have s half-integer.
a) Quantum effects can keep some of them liquid down to low temperatures.
Two types of quantum liquids, stable at T = 0, are available:
(i) liquid 3 He and 4 He. Without applying substantial external pressure, helium
does not solidify down to T = 0 – because its zero-point kinetic energy is large
(due to the very weak interatomic attraction and low atomic mass m) and is
comparable with the energy of interatomic attraction;
(ii) electrons in metals (conduction electrons can be thought of as a liquid filling
the metal; here T ∗ ∼ 105 K is huge compared to even room temperature T ∼
300 K).
a) Reduce the particle density to ∼ 10−5 of that of air to slow down their re-
combination into molecules (which requires three-particle collisions to take away
energy from recombination); this should be done in a sealed chamber.
b) Prevent atoms from reaching the walls of the chamber (by a magnetic trap
– this only works with particular spin states of atoms).
c) Cool down below T ∗ in two steps:
(i) laser precooling (red-shifted laser tricks atoms into absorbing less and then
emitting more energy);
(ii) evaporative cooling by gradual reduction of the depth of the trapping poten-
tial well: the fastest particles keep escaping leaving slower ones behind (however,
2
this crucially relies on the interaction between particles for their thermalization;
i.e. will not work with a truly ideal gas).
They are labeled by the allowed wavenumbers k that match the boundary con-
2 2
ditions. The energy levels are ϵ(k) = h̄2m
k
, where k = |k|. The state of the
lowest energy ϵ = 0 corresponds to k = 0, and its wavefunction is
For a single particle, the arbitrary choice of the phase θ1 has no observable
consequences.
For N bosons (density n = N/V ), below the critical temperature for ‘Bose-
Einstein condensation’ (BEC)
2πh̄2 n 2/3
Tc = , (1.4)
mkB 2.612
a macroscopic number of particles N0 condenses into the same quantum state
with the lowest energy so that
3/2
N0 (T ) T
=1− , (1.5)
N Tc
This is a phase transition into a more ordered state (in momentum space).
Macroscopic wavefunction
In the Bose condensate at rest (i.e. k = 0), all N0 particles are in the same
quantum state described by wavefunctions (1.3) in which the phases θ1 become
−1/2
close to each other, with a very small uncertainty ∆θ1 ∼ N0 ≪ 1 when their
3
number N0 is large. Hence, we can use their mean phase θ =< θ1 > as another
macroscopic variable alongside with their density n0 = N0 /V . Thus properties
of all particles in the condensate are described by just two parameters, n0 and
θ, through the ‘macroscopic wavefunction’
1/2
N0 1/2
ψ= exp(iθ) = n0 exp(iθ), (1.7)
V
Exercises
4
2 Superfluid helium and macroscopic wavefunc-
tion
At low temperatures and pressures, helium liquefies but never solidifies – thanks
to the low mass of its atoms and very weak attraction between them (and hence a
relatively substantial kinetic energy of zero-point motion). We thus have a liquid
at arbitrarily low temperatures! At Tλ = 2.17 K (at P = 0) liquid 4 He undergoes
a sharp phase transition into a state with a new internal order. Historically, this
state was called “He II” (and the normal state at T > Tλ is “He I”) but we will
refer to it as “superfluid 4 He” or simply “superfluid helium”. In this state,
steady flow is possible without any losses, heat transport is extremely fast,
oscillating flow behaves as if there were some viscous (“normal”) component in
the liquid but of a reduced density ρn (T ) < ρ (where ρ is the density of all
liquid); at T = 0 this “normal component” vanishes: ρn (0) = 0. The remaining
component of density ρs (T ) = ρ − ρn (T ) behaves as an inviscid ordered liquid
(called the “superfluid component”).
Macroscopic coherence
∇ × vs = 0. (2.9)
5
Quantization of circulation
Quantized vortices
To allow rotation, the magnitude of the order parameter |ψ(r)| becomes inho-
mogeneous, creating continuous lines along which ψ = 0 (hence, making the
phase θ undefined on those lines). This makes the superfluid formally multiply-
connected, so non-zero circulation around such a line becomes possible. The
superfluid velocity around the vortex line is vs (r) = κ/2πr.
2
κ
The vortex energy per unit length is ρs 4π ln(R/a0 ), where R is the radius of
container (or half the distance to the nearest vortex in an array). Rotation of
the superfluid component at angular velocity Ω in a simply-connected volume
takes forms of an array of quantized vortex lines, each carrying one quantum of
circulation, κ = h/m, and the areal density of lines is 2Ω/κ (‘Feynman’s rule’).
Exercises
1. Look up the lecture materials and review main properties of superfluid he-
lium.
2. Repeat derivation of the quantization of circulation.
3. Calculate the value of the quantum of circulation κ.
4. Sketch the succession of flow patterns (velocity and phase maps, as seen from
above) while one vertical vortex traverses a flow across the channel (phase slip,
see lecture materials).
5. Estimate the distance between vortex lines in a bucket of superfluid helium
rotating at Ω = 1 rad s−1 .
6
3 Excitations. Two-fluid description
Ground state and excitations. Landau criterion.
Particle-like excitations of all liquid have the momentum p and energy ϵ(p).
Landau criterion for superfluidity: non-dissipative motion at a speed v past an
object (wall) is possible if v < vL :
ϵ
vL = min . (3.11)
p
The ground state of the non-interacting Bose gas is when all particles are in
the lowest energy state. Hence, excitations are associated with free particles
in higher energy states. In contrast, weakly repulsive particles have a wave-
like collective mode (sound) as low-energy excitations, with dispersion ϵ = cp.
Hence, a non-interacting Bose-Einstein condensate (ϵ = p2 /2m) is not superfluid
(vL = 0)! But a weakly repulsive Bose gas (ϵ = cp for p → 0) is superfluid
(vL = c > 0). So is liquid helium below Tλ .
∂U
in which the chemical potential, µ ≡ ∂N S,V
, is set to zero (µ = 0) – because
the number of excitations is not conserved (their number becomes a free param-
eter with which the system minimizes its free energy when in thermodynamic
equilibrium).
7
The temperature T reflects the density of excitations,
Z ∞
d3 p
n(T ) = f (p) . (3.13)
0 (2πh̄)3
and of rotons,
nrot (T ) ∝ T 1/2 exp(−∆/kB T ). (3.15)
Phonons dominate at T < 0.6 K and rotons – at T > 0.7 K. The phonon con-
tribution to the normal component, specific heat and entropy, ρn (T ) ∝ T 4 ,
Cph (T ) ∝ T 3 , Sph (T ) ∝ T 3 , scale as power laws of T ; while the leading term of
the roton contributions is exponential ∝ exp(−∆/kB T ).
Exercises
8
4 Two-fluid hydrodynamics
Classical Fluids
∂ρ
+ ∇ · (ρv) = 0. (4.16)
∂t
The conservation of momentum (Newton’s 2nd law): Navier-Stokes equation,
∂v
ρ + v · ∇v = −∇P + η∇2 v. (4.17)
∂t
For sufficiently small velocities, the LHS can be linearized (only term linear in
v left):
∂v
ρ = −∇P + η∇2 v. (4.18)
∂t
The relaxation time τ (mean free path l = v̄τ ) is set by collisions with
other particles of fluid (Poiseuille regime when this is dominant), volume de-
fects (Drude regime when this is dominant – often for electrons in metals) and
9
ultimately (for either low density of particles or pure metals) by boundaries such
as pipe walls or surface of wire (Knudsen regime).
For the Poiseuille regime in a cylindrical pipe of radius R and pressure drop
∆P over distance ∆x (∇P = ∆P ∆x ), the total volume throughput per unit time
in the steady-state is
dV πR4
= ∇P (4.20)
dt 8η
(i. e. requires pressure gradient). For a gas with the mean velocity of particles
v̄, the viscosity is η = 31 ρlv̄. Similarly, the thermal conductivity is κ = 13 CV lv̄.
∂ρ
+ ∇ · (ρs vs + ρn vn ) = 0. (4.21)
∂t
The conservation of entropy S (per unit volume of all fluid) for reversible pro-
cesses (only the normal component carries entropy, at velocity vn ):
∂S
+ ∇ · (Svn ) = 0. (4.22)
∂t
The dynamics of the superfluid component (the force on atoms of the frictionless
superfluid component is due to the gradient of their potential energy µ),
∂vs
m + v · ∇vs = −∇µ = −v∇P + vS∇T,
∂t
where the Gibbs-Duhem relation, dµ = vdP −vSdT , was used (µ being chemical
potential per atom, v = m/ρ atomic volume, m mass of a 4 He atom). For small
velocities vs it can be linearized and re-written as
∂vs 1 S
= − ∇P + ∇T. (4.23)
∂t ρ ρ
The conservation of the total momentum (here ηn is the viscosity of the normal
component):
∂vs ∂vn
ρs + ρn = −∇P + ηn ∇2 vn . (4.24)
∂t ∂t
∇ × vs = 0. (4.25)
10
Fountain effect
With superfluid helium in a pipe, one can administer a pressure gradient along
it by maintaining a difference in temperatures of pipe’s ends – and hence push
the fluid through the pipe without any pump. This is called fountain effect.
∇P = S∇T. (4.26)
To maintain this pressure difference, the vertical levels of free surfaces of helium
in two reservoirs have to be different by ∆H given by:
One can see that the superfluid and normal components are accelerated in
the same direction when acted upon by a pressure gradient ∇P , but are accel-
erated in the opposite directions when acted upon by a temperature gradient
∇T . This results in two different types of oscillating modes (sounds).
First sound is a wave of pressure and density, while T = const and S = const,
vn = vs . Oscillations of ρ(x, t) and P (x, t) propagate with the velocity c1 , given
by the adiabatic compressibility like in any other fluid :
2 ∂P
c1 = . (4.28)
∂ρ S
11
propagate with the velocity c2 ,
ρs S 2 T
c22 = , (4.29)
ρn ρCP
Exercises
1. Look up the lecture materials and review main differences between the equa-
tions describing the hydrodynamics of classical fluids and of superfluid helium.
2. Sketch instantaneous velocities of the superfluid and normal components and
the associated variations of:
(i) pressure and density in a plane wave of first sound;
(ii) temperature and entropy in a plane wave of second sound.
3. Sketch the temperature dependence of the velocity of second sound c2 (T ).
4 (optional). Derive the formulas for the velocities of first and second sound in
superfluid helium.
12
Mid-term checklist
(superfluid helium means ‘liquid 4 He below Tλ = 2.17 K’):
13
5 Electrodynamics of Superconductors
Revision: Electromagnetic Fields (valid for all materials)
Fundamental fields E and B can be determined by measuring the force F
they exert on a charge, q, moving at a speed v:
F = q(E + v × B)
E = −∇ϕ
B=∇×A
For the same fields, E and B, ϕ and A are not uniquely defined. We often fix
the uncertainty by choosing ϕ(∞) = 0 and ∇ · A = 0
D = ϵ0 E + P,
H = B/µ0 − M,
−12
where ϵ0 = 8.85 · 10 F/m (or A2 s2 /Nm2 ) and µ0 = 4π · 10−7 H/m (or N/A2 ).
∇·B=0
∂B
∇×E+ =0
∂t
∇ · D = qn
∂D
∇×H=j+ .
∂t
14
Properties of Normal Metals (reference)
Macroscopic variables, all depending on position and time (r, t): number
density of charged particles n, their charge q (usually, q = −e), electrostatic
potential ϕ, drift velocity v, current density j = qnv.
where the relaxation time τ = l/vF is introduced to account for their finite
mean free path before exchanging acquired momentum with the crystal lattice
and other electrons (vF = pF /m is the Fermi velocity).
In the Drude regime, particles mainly collide with defects randomly dis-
tributed in the volume of a metal wire. A non-zero potential gradient along the
wire, ∇ϕ = −E, is required to maintain the steady flow of charge of current
density j = qnv – hence the Ohm’s law (from Eq. 5.31):
j = σE, (5.32)
ne q 2 τ ne q 2 l
σ= = , (5.33)
m mvF
where ne is the density of conducting electrons.
15
ne σ −1 (77 K) σ −1 (273 K) m/me ΘD TF vF
10 cm−3
23
µΩ cm µΩ cm K 105 K 108 cm/s
Cu 0.85 0.2 1.56 1.3 310 0.82 1.57
Pb 1.32 4.7 19.0 1.9 88 1.10 1.83
Properties of superconductors
16
Two-fluid model and London equations
nn e2 τ
jn = σE = E. (5.39)
m
- The superconducting electrons have zero resistivity and entropy. They obey
the London equations (F. and H. London, 1935):
ns e2 µ0
∇2 B = B. (5.42)
m
Take a plane surface of a superconductor, x-axis perpendicular to the surface,
parallel external field Bz = B. The decaying solution is:
where
m
λ2L = . (5.44)
µ0 ns e2
17
Macroscopic wave function for superconducting electrons
[With hindsight, we can think of charge carriers called Cooper pairs of mass
m∗ = 2m, charge q = −2e and density n∗ (T ) = ns /2]:
qh̄ q2
js = ∗
|ψ|2 ∇θ − ∗ |ψ|2 A. (5.47)
m m
This is the London equation in the most general form: apply ∇× to both sides
and you will recover the second Londons equation (because ∇ × ∇θ = 0).
Exercises
18
6 Cooper Instability
Electron-electron interaction
For the Cooper effect (formation of pairs) we need some effective attractive
interaction between electrons. This should overcome the Coulomb repulsion,
which is fortunately screened by other electrons on pretty short distances.
Positively charged ions are attracted to the nearest electron and slowly (be-
cause they are heavy) accelerate towards it. Hence, a moving electron leaves
a wake of enhanced positive charge density. Other electrons are attracted to
this wake. We thus have a short-range electron-electron attraction mediated
by the deformation of the crystal lattice (i. e. emission and absorption of a
phonon with momentum h̄q and energy h̄ωq , during which one electron changes
its momentum from h̄k1 to h̄k1 − h̄q and another – from h̄k2 to h̄k2 + h̄q). In
quantum mechanics, such interaction is described by the matrix elements Vkk′
connecting all possible initial and final electron states:
g 2 h̄ωq
Vkk′ = , (6.50)
(ϵk′ − ϵk )2 − (h̄ωq )2
19
Cooper pairs
In the presence of attractive interaction between electrons the normal ground
state is unstable against the formation of bound states (Cooper 1956). Consider
the “Fermi sea” at T = 0: non-interacting electrons fill all levels (each specified
by a wave vector k and spin projection σ ≡ 1/2 or −1/2) with kinetic energy
ϵk = h̄2 k 2 /2m up to the Fermi energy EF = h̄2 kF2 /2m. Add two special elec-
trons (coordinates r1 , r2 ) that interact via a pair potential V (r1 − r2 ). Their
coordinate wave function is Ψ(r1 , r2 ). The Scrödinger equation:
h̄2
− 2m (∇21 + ∇22 ) + V (r1 − r2 ) Ψ(r1 , r2 ) = EΨ(r1 , r2 ), (6.52)
where E is the eigenvalue of their energy. Our goal is to calculate the lowest
possible E.
After switching to the coordinate of the centre of mass R = (r1 + r2 )/2 and
∂R ∂r
the relative coordinate r = (r1 − r2 ) it becomes (we use ∇1 = ∇R ∂r 1
+∇r ∂r 1
):
h̄2 2 h̄2
− ∇R − ∇2r + V (r) Ψ(R, r) = EΨ(R, r). (6.53)
4m m
After separating the centre-of-mass and relative coordinates we get:
Let us express Ψ(R) as a sum over plane-wave states which are still available
a(k)eik·r .
P
Ψ(r) = |k|>kF (6.57)
′
We multiply it by e−ik ·r and integrate over volume:
Z
X ′
(2ϵk′ − E) a(k′ )Ω + a(k) V (r)ei(k−k )·r d3 r = 0, (6.59)
k>kF
20
where we used the relation (Ω is the volume of the system)
Z
′
ei(k−k )·r d3 r = Ωδ(k − k′ ). (6.60)
The sum can be expressed as an integral over the energies in terms of the density
of states D(ϵ) = dN/dϵ . In the narrow range from EF to EF + h̄ωD , D(ϵk ) is
nearly constant: D(ϵk ) ≈ D(EF ):
Z EF +h̄ωD
dϵ 1 2(EF + h̄ωD ) − E
1 = V0 D(EF ) = V0 D(EF ) ln . (6.67)
EF 2ϵ − E 2 2EF − E
Without interaction the energy of the pair would be 2EF . It is convenient to
define the energy difference Eb ≡ E − 2EF . We thus have
2h̄ωD
Eb = − . (6.68)
exp(2/V0 D(EF )) − 1
In the weak-coupling limit V0 D(EF ) ≪ 1 we finally have
Eb ≈ −2h̄ωD exp − D(E2F )V0 . (6.69)
For repulsive interactions (V0 < 0) Eb > 0; the normal state (Fermi sea) is
the ground state. But for any attraction (V0 > 0), there is a bound state with
Eb < 0. More Cooper pairs will form (all in state K = 0) to reduce total energy.
At finite temperature, the bound state is stable at T < Tc ∼ |Eb |/kB . Eb
is much smaller than h̄ωD . We thus have three well-separated energy scales:
kB Tc ∼ |Eb | ≪ h̄ωD ≪ EF .
21
7 BCS theory
BCS theory at T = 0
(Bardeen, Cooper, Schriffer 1957) Let us treat all electrons on the same foot-
ing. The Hamiltonian is Ĥ = K̂ + V̂ . Variational method: we will vary the
parameters of the trial wave function Φ to find the state at which the energy is
a minimum.
The ‘BCS wave-function’: only pairs (k, −k) are allowed. For any k, a pair
of electrons with (k, ↑) and (−k, ↓) is either present (|11 >) or absent (|00 >):
Q
Φ= k ψk , (7.70)
For comparison, the normal ground state has sharp step-like functions:
uk = 0, vk = 1 k < kF
(7.73)
uk = 1, vk = 0 k > kF .
In the presence of attractive interaction one can decrease the total energy by
rounding these distributions – similar to the Cooper problem, Eq. (6.65).
In this approach we do not fix the total number of particles N (as only
the probabilities of occupation are specified). Even though there exists a final
probability to find the system with any N , for a macroscopic system the main
contribution arises from configurations with the number of particles nearly equal
to the expectation value < Φ|N̂ |Φ >. To find the minimum of < Φ|Ĥ|Φ >
while keeping < Φ|N̂ |Φ > constant, we will minimize the expectation value of
< Φ|Ĥ − µN̂ |Φ >, where the chemical potential is µ = EF , and the single
particle kinetic energies ξk ≡ ϵk − EF are measured relative to EF .
The potential energy is (summation over all processes Vkk′ each scattering the
state (k occupied, k′ empty) with the amplitude u∗k′ vk into the state (k empty,
k′ occupied) with the amplitude vk′ u∗k ):
X
< Φ|V̂ |Φ >= Vkk′ uk vk∗ ′ u∗k′ vk . (7.75)
k,k′
22
All vk and uk have the same phase, so we assume they are real. After combining
the previous two equations, we get the expectation value of the total energy
2ξk vk2 +
P P
< Φ|Ĥ − µN̂ |Φ >= k k,k′ Vkk′ uk vk′ uk′ vk . (7.76)
It is convenient to define
P
∆k ≡ − k′ Vkk′ uk′ vk′ , (7.79)
vk = sin θk
uk = cos θk , (7.81)
or
ξk sin 2θk − ∆k cos 2θk = 0. (7.83)
If we define
Ek ≡ (ξk2 + ∆2k )1/2 , (7.84)
23
BCS used a simplified form of Vkk′ :
−V0 |ξk | < h̄ωD ,
|ξk′ | < h̄ωD
Vkk′ = (7.87)
0 otherwise.
with summation over the narrow interaction shell (7.87) |ξk | < h̄ωD , |ξk′ | < h̄ωD .
We can reduce summation to integration (D(ξ) = dN/dξ):
X Z h̄ωD
= D(0) dξ (7.89)
k −h̄ωD
Then
Z h̄ωD
dξ h̄ωD
1 = 2D(0)V0 = D(0)V0 sinh−1 , (7.90)
0 2(ξ 2 + ∆2 )1/2 ∆
and hence,
h̄ωD
∆= . (7.91)
sinh(1/D(0)V0 )
For many superconductors, D(0)V0 < 0.3 (except for Pb and Hg where: D(0)V0 =
0.39 and 0.35, respectively). For the weak-coupling limit, D(0)V0 ≪ 1,
1
∆ = 2h̄ωD exp(− D(0)V 0
). (7.92)
We can now find the final expressions for uk and vk . Using Eq. (7.84) and
Eq. (7.85) we get
u2k − vk2 = ξk (ξk2 + ∆2 )−1/2 . (7.93)
Together with the normalization condition Eq. (7.72) this gives:
1
u2k = 1 + ξk (ξk2 + ∆2 )−1/2 ,
2
1
2
vk = 1 − ξk (ξk2 + ∆2 )−1/2 . (7.94)
2
Condensation Energy
Let’s evaluate the total energy of the BCS state, using Eq. (7.76) and Eq. (7.79):
2
− ξk2 /Ek − ∆2k /2Ek ).
P P
E0 = k (2ξk vk − ∆k uk vk ) = k (ξk (7.95)
24
We expect E0 to be less than the energy of the normal state EN :
P
EN = k<kF 2ξk . (7.96)
2
E0 − EN = −2D(0)(h̄ωD )2 exp(− D(0)V 0
). (7.100)
This is the condensation energy at T = 0 which for weak coupling limit can be
rewritten using Eq. (7.92):
E0 − EN = − 21 D(0)∆2 . (7.101)
−1
fk = (1 + exp(Ek /kB T )) , (7.103)
25
The gap equation Eq. (7.86) becomes
X ∆k′
∆k = − Vkk′ (1 − 2fk′ ) . (7.106)
′
2Ek′
k
Replacing sum with an integral and using the simplified BCS form of Vkk′ gives:
Z h̄ωD
dξ
1 = D(0)V0 (1 − 2f ((ξ 2 + ∆2 )1/2 )). (7.107)
0 (ξ + ∆2 )1/2
2
Unconventional superconductors
In many metals electrons make spherically-symmetric spin-less Cooper pairs
(both total orbital momentum and total spin are zero), the collective motion of
which is described by a common complex order parameter – very much like in
superfluids made of bosonic atoms and molecules. It turned out that in some
fermionic systems (like electrons in high-Tc and ‘heavy-fermion’ superconductors
but also atoms in liquid 3 He), Cooper pairs are anisotropic and can have internal
degrees of freedom (orientation of the orbital and spin angular momenta). This
greatly enriches the physics of their superconducting state, allowing variations,
in space, of the values of parameters describing their multi-component wave-
function; for instance, several types of quantized vortices and other topological
defects of the order parameter are possible. The microscopic BCS theory can
often be modified to account for the anisotropy of Cooper pairs and non-phonon
mechanism of pairing in these materials.
26
8 Thermodynamics of superconductors in the
Meissner state
When a system is held at constant applied magnetic field, Ba , (with use of, say,
an external magnet and its current supply) – the equilibrium state corresponds
to the minimum of the Gibbs free energy (here g is its density per unit volume):
g ≡ U − T S + P − Ba · M, (8.112)
where U and S are the densities of energy and entropy. After using the first law
of thermodynamics, for constant T and P the differential is
dg = −M · dB. (8.113)
We thus have (valid for any phase)
Z Ba
g(Ba ) = g(0) − M · dB. (8.114)
0
27
- While in the normal phase (almost non-magnetic), Mn ≈ 0, hence:
Tc , K Bc (0), mT
Al 1.19 10
Hg 4.16 39.5
Nb 9.2 195
Pb 7.2 80
Sn 3.72 30.4
Zn 0.86 5.2
∂g
The entropy is S = − ∂T , hence:
B
Bc dBc
Sn (T, 0) − Ss (T, 0) = − . (8.120)
µ0 dT
For T < Tc , the RHS is always positive (because dBc /dT < 0), hence: Sn ≥ Ss ,
i. e. the superconducting state is more highly ordered.
28
Exercises
29
9 Ginzburg–Landau Theory
After a phase transition into a more highly ordered state, it usually differs from
the old one by a new symmetry; this new symmetry can often be quantified
using an observable quantity (called order parameter), which is non-zero only
in the more highly ordered phase. Continuous phase transitions (such as the
superconducting transition in zero magnetic field) have the benefit of the fact
that the order parameter is continuous through the transitions (i.e. it grows
from zero as one crosses into the ordered phase) – hence, one might guess the
analytical dependence of the order parameter from some general principles. A
typical example of the order parameter in a ferromagnet is its magnetization M
– a vector pointing in a certain spontaneously-chosen direction (while magneti-
zation was zero in the paramagnetic phase above the critical temperature). In
the superconducting state, for the order parameter it is convenient to take the
complex-valued macroscopic wavefunction ψ(r) (which is non-zero only below
Tc ); the square of its magnitude gives the density of Cooper pairs: |ψ(r)|2 = n∗
while the phase θ is chosen spontaneously.
where Fn0 is the free energy of the normal state in zero field, and the difference
in the density of free energy,
1 2 2 1 4 (∇ × A)2
fs − fn0 = |(−ih̄∇ − qA)ψ| + α(T )|ψ| + β(T )|ψ| + , (9.123)
2m∗ 2 2µ0
can be expanded in powers of the order parameter ψ(r) just below Tc (see
below). In equilibrium the superconductor distributes ψ(r) and B(r) (i. e. n∗ (r)
and js (r) = µ10 ∇ × B) to minimize the total free energy Fs .
30
Justification of the above expression for fs − fn0 :
Now let’s find the wave function |ψ| = const with which fs − fn0 = min:
β 4
fs − fn0 = α(T )|ψ|2 + |ψ| , (9.127)
2
dfs
0= = α(T ) + β|ψ|2 , (9.128)
d|ψ|2
α a
|ψ|2 = − = (Tc − T ), (9.129)
β β
a2
(fs − fn0 )min = − (Tc − T )2 . (9.130)
2β
B2
This is the condensation energy, − 2µc0 , hence Bc ∝ (Tc − T ).
31
Ginzburg-Landau equations
We have to find the conditions for a minimum of Fs (ψ(r), A(r)). Bu taking the
variational derivative with respect to ψ we arrive at:
1
(−ih̄∇ − qA)2 ψ + αψ + β|ψ|2 ψ = 0. (9.133)
2m∗
And taking the variational derivative with respect to A, we get:
ih̄q ∗ ∗ q2
js = − (ψ ∇ψ − ψ∇ψ ) − A|ψ|2 . (9.134)
2m∗ m∗
These two are called the Ginzburg-Landau equations. The first one looks like the
Shrödinger equation with a non-linear, superfluid density dependent, potential
energy. And the second one can be rewritten as the London equation,
q
js = |ψ|2 (h̄∇θ − qA). (9.135)
m∗
h̄2 d2 ψ
− + αψ = 0. (9.137)
2m∗ dx
For the given boundary condition ψ(0) = 0, the general solution of this ODE,
for x > 0, is ψ = A sin(x/ξGL ). Plugging it into (9.137) yields the characteristic
length scale of the first Ginzburg-Landau equation:
2 h̄2 h̄2
ξGL =− = . (9.138)
2m∗ α 2m∗ a(Tc − T )
32
In the Londons theory, there was no allowance for any variations of ψ inside a
superconductor); only the magnetic field could vary over distances greater than
the London penetration depth λL , which also arises from the second Ginzburg-
Landau equation (because it is basically equivalent to the second Londons equa-
tion).
Exercises
33
10 Type II superconductors. Vortex state.
For type-I superconductors (if λ < ξ, hence σ > 0) – to minimize the total
energy, the N-S surface area should not be too large. At certain values of Ba
some macroscopic parts of the sample go normal and allow the magnetic flux
through: this is the intermediate state (not really a homogeneous phase, but
just coexisting macroscopic regions of the superconducting and normal phases),
where the particular shape of the N-S boundary depends on the sample’s shape.
For type-II superconductors (when λ > ξ), hence σ < 0 – the N-S surface
becomes unstable with respect to increasing its area (by increasing the number
of smaller normal islands penetrated by magnetic field) until flux quantization
will limit this growth. The magnetic flux now is arranged in a dense regular
lattice of quantized flux lines, each carrying a single flux quantum, Φ0 = h/2e.
This is the vortex state – a new homogeneous phase in which an average magnetic
field co-exists with superconductivity.
34
Intermediate state in type-I superconductors (optional)
Type-II superconductors
In the vortex state, the flux penetrates as quantized flux lines – each with the
smallest possible flux Φ0 . The field is B0 at the axis of a line but is screened
within radius ∼ λ by superconducting current (hence, another name for this
line defect: Abrikosov vortex); hence πλ2 B0 ∼ Φ0 . On the axis, the superfluid
order parameter is suppressed (|ψ| = 0) but recovers to the bulk value at the
35
distance from the axis ∼ ξ (“core size”); the phase of the order parameter θ
winds up 2π round the axis.
The energy per unit length of a flux line is (solution of G-L equations):
2
1 Φ0
Γ= (ln κ + ϵ), (10.145)
4πµ0 λ
(the small parameter ϵ represents the contribution from the core region; it is
negligible for κ ≫ 1). Γ is positive, that means the lines are under tension.
Two parallel flux lines repel each other by a short-range force of radius ∼ λ.
This mutual repulsion stabilizes a 2-d hexagonal lattice of flux lines.
ln κ Φ0 2
Because flux lines have positive energy Γ ≈ 4πµ 0 λ (for κ ≫ 1), the vortex
state can only become energetically favorable over the Meissner state if the
applied magnetic field Ba exceeds a finite value Bc1 :
√
Bc1 = Bc ln κ/ 2κ ∼ Bc κ−1 , for κ ≫ 1, (10.146)
2|α|m∗ h̄ √
Bc2 = = 2 = 2Bc κ. (10.147)
qh̄ qξ
√
In the limit κ = 1/ 2, the difference between type-I and type-II superconduc-
tors disappears, and Bc1 = Bc2 = Bc . And for κ → ∞, Bc2 → ∞, i. e. extreme
type-II superconductors with high κ ≫ 1, the vortex state survives up to fields
B ∼ Bc2 ≫ Bc .
Flux pinning
FL = j × B. (10.148)
36
Without pinning, the flux lines would move across the current; this would gen-
erate emf (voltage in the directions of current) and hence energy dissipation.
In a particular material, there exists a certain maximum pinning force per unit
volume F p (max). Hence, there is a critical current density,
jc = F p (max)/B, (10.149)
Certain flux lines are normally pinned to the regions of weakened supercon-
ductivity, i. e. impurities or crystalline defects (as the vortex core is normal, the
loss of condensation energy is smaller if the core passes through such a region).
The other lines are restrained because of the repulsion from other lines, i. e. the
rigidity of the flux lattice – this effect is called collective pinning: the critical
pinning force F p (max) becomes dependent on field B.
Flux creep is a slow relaxation of pinned arrays of lines from one metastable
pinned configuration to another. Creep is always associated with finite voltage
and dissipation.
Applications
Because of high jc and high Bc2 , superconducting magnets made of type-II wires
can produce very stable B ∼ 20 T at zero power. To maximize pinning, one can
use a combination of cold working and heat treatment (for NbTi: material is
divided into cells by dislocation walls, small particles of Ti-rich phase precipitate
on these walls or pinning to grain boundaries (for Nb3 Sn). To make the wire
stable against sudden local temperature increase caused by creep (“flux jump”),
they are made of thin filaments (∼ 30 µm) enclosed into a good heat conductor
(usually copper).
37
Why would one want to have high magnetic fields?
1. NMR imaging of various elements in tissue, lungs and bloodstream.
2. Physics research:
• condensed matter research (e. g. of graphene);
• particle physics: to guide charged particles in accelerators by strong
magnetic field;
• to contain plasma in fusion reactors (tokamaks);
• cooling to ∼ 10−6 K by demagnetizing a paramagnet from field ∼ 8 T.
3. Magnetic levitation.
4. Magnetic separation: e.g. Cu-Ni ore.
5. Energy storage.
6. Compact and powerful motors, generators, etc.
ln κ Φ0 2
Because flux lines have positive energy Γ ≈ 4πµ 0 λ (for κ ≫ 1), the vortex
state can only become energetically favorable over the Meissner state if the
applied magnetic field Ba exceeds a finite value Bc1 . At this field, Ba = Bc1 ,
the Gibbs free energies of these two competing states become equal:
38
√
2Φ0
Using Bc = 4πλξ (which can be obtained from the expressions for condensation
α2 Bc2
energy 2β = 2µ0 , and for ξ and λ in terms of α and β), we get
Bc
Bc1 = √ ln κ, (10.155)
2κ
As the applied field Ba approaches the upper critical value Bc2 , the supercon-
ducting order parameter vanishes, |Ψ| → 0. We thus can use the first Ginzburg-
Landau equation without its non-linear term:
1
(−ih̄∇ − qA)2 ψ = |α|ψ. (10.156)
2m∗
Our task is to find the highest possible Ba that still allows a non-trivial (i. e.
with |Ψ| =
̸ 0) solution of this equation. Fortunately, this equation is equivalent
to the Schroedinger equation for a stationary state Ψ of a particle of charge q
in magnetic field B = ∇ × A and with the eigenvalue of energy Ec = |α|. The
solutions of the latter, for a uniform field along z-axis, B = (0, 0, B), corresponds
qB
to quantized cyclotron rotation in xy-plane at frequency ωc = m ∗ )while moving
h̄2 kz2
1 h̄qB
Ec = n + + , (10.157)
2 m∗ 2m∗
39
Exercises
1. Sketch |ψ(r)| and B(r) as function of the distance r from the axis of a flux
line when the value of an applied field is:
(i) just above Bc1 , (ii) just below Bc2 .
2. For the Pb-In alloy and YBa2 Cu3 O7 , using the parameters given in the table
in the notes (and inventing a way to estimate the value of Bc ), make estimates
on Bc1 and Bc2 at T = 0.
3*. Complete the derivations of the formulae for Bc1 and Bc2 , given in the notes.
40
11 Josephson junctions
The long-range order of the phase of the order parameter means that fixing its
value at one point determines its value at any other point of the same piece of
superconductor. Consider two isolated pieces of superconductor with phases θ1
and θ2 . If Cooper pairs have a small (but finite) probability of tunneling between
two bulk pieces, the superconducting current I between the pieces will depend
on the phase difference ∆θ ≡ θ2 − θ1 – and can hence be used to determine
the value of ∆θ. The function I(∆θ) should be periodic with a period 2π. The
simplest one is I ∝ sin ∆θ which is generally valid for very weak links (in the
solutions to today’s example sheet, see a derivation of this formula using the
G-L equations). Such links are called Josephson junctions and are typically
made of a thin (∼ 20 Å) insulating film between two superconductors.
Josephson equations
41
After plugging this result in the first equation we obtain
q
Is = Ic sin ∆θ0 − V t , (11.165)
h̄
which tells us that an application of a constant voltage to the Josephson junction
generates an AC current with frequency
|qV | |V |
νJ = = . (11.166)
h Φ0
This current doesn’t dissipate energy because the time average of Is V is zero!
Since 1990 the volt has been maintained using the Josephson effect, where
the Josephson constant is fixed by the 18th General Conference on Weights and
Measures as: νJ /V = 2e/h = 0.4835979 GHz/µV. This is typically used with
an array of several thousand or tens of thousands of junctions at microwave
frequences between 10 and 80 GHz.
Without magnetic field, θ was the same in all bulk parts of the superconductor
where Is = 0. With field, only γ is the same everywhere, while θ(r) becomes a
function of position r and subject to the choice of gauge (A).
42
Interference effects in magnetic field. DC SQUID
DC SQUID: Consider two Josephson junctions in parallel (both having the same
critical current Ic ): one from point (a) to (b) conducts current I1 , the other from
point (d) to (c) conducts current I2 . The total current
1 1
I = I1 +I2 = Ic (sin ∆γ1 +sin ∆γ2 ) = 2Ic cos (∆γ1 − ∆γ2 ) sin (∆γ1 + ∆γ2 ) .
2 2
(11.170)
We will now show that the value of ∆γ1 −∆γ2 is determined by the magnetic
flux through the loop made by the wires connecting the junctions.
q b
Z
∆γ1 = θb − θa − A · dl (11.171)
h̄ a
and
q c
Z
∆γ2 = θc − θd − A · dl. (11.172)
h̄ d
q a q c
Z Z
∆γ1 − ∆γ2 = (θd − θa ) + (θb − θc ) + A · dl + A · dl. (11.173)
h̄ b h̄ d
Note that inside bulk superconducting wires, js = 0, hence (from the second
G-L equation, which is basically a definition of the value of js ),
q
js = ∗ |ψ|2 (h̄∇θ − qA) = 0, (11.174)
m
we get
q
∇θ = A. (11.175)
h̄
Hence,
Z d
q d
Z
θd − θa = ∇θ · dl = A · dl, (11.176)
a h̄ a
Z b
q b
Z
θb − θc = ∇θ · dl = A · dl. (11.177)
c h̄ c
q 2π
Finally, as h̄ = Φ0 ,
1 Φ
cos (∆γ1 − ∆γ2 ) = cos π , (11.179)
2 Φ0
43
and
πΦ 1
I = 2Ic cos sin (∆γ1 + ∆γ2 ) . (11.180)
Φ0 2
The first term, 2Ic cos( πΦ
Φ0 ), gives the total critical current for a given flux Φ.
The second term, sin( 12 (∆γ1 +∆γ2 )), automatically adjusts the parameters ∆γ1
and ∆γ1 to sustain the given total supercurrent I when I ≤ Ic .
The flux from a pick-up coil (which is near or around the sample) is often
transferred to the SQUID loop by the “flux transformer” – a closed loop of a su-
perconducting wire. The total flux in this loop is quantized (hence, conserved).
If the flux through one (pick-up) part of such loop has changed, a supercurrent
is induced to compensate for it. Hence, the flux through another part of the
transformer (which is next to the SQUID) reflects the change in the signal –
and can thus be monitored by the SQUID.
– in geology – mapping Earth’s magnetic field to detect iron, nickel ore, etc.
44
Exercises
45
End-term (second-half of the course) checklist
(for the first-half checklist go to page 13)
46
12 Numerical answers to exercises
3.3. Even in very narrow channels, the critical velocity for phase slip vc is
smaller than vL . Hence, dissipation due to phase slips always prevails.
47