You are on page 1of 30

Chapter 3

Ginzburg-Landau theory

Many important properties of superconductors can be explained by the theory, devel-


oped by Ginzburg an Landau in 1950, before the creation of the microscopic theory.
The Ginzburg-Landau (GL) theory often provides much simpler and more tractable
description than the BCS theory. Originally the GL theory was developed as a phe-
nomenological model. In 1958 Lev Gor’kov demonstrated that it can be derived from
the microscopic theory within certain restrictions on the parameters. The most im-
portant restrictions are |1| ≪ kB Tc , i.e. temperature should be close to the transition
temperature, and λL ≫ ξ0 , i.e. electrodynamics is considered local, like in the London
model. Qualitatively, however, predictions of the GL theory are applicable in a much
wider parameter range.
The GL theory has some limitations. It deals only with the gap parameter 1 and
does not describe behavior of quasiparticles. It is also less successful in description of
non-stationary (time-dependent) states.
The starting point of the GL theory is the Landau theory of the symmetry-breaking
phase transitions, which we will briefly introduce.

3.1 Landau theory of phase transitions


Landau considered phase transitions where some symmetry is spontaneously broken:
The Hamiltonian possesses a certain symmetry on both sides of the transition, while
the ground state obeys this symmetry on one side of the transition (symmetric phase)
and looses this symmetry on another side (broken-symmetry phase). He introduced an
order parameter ψ which measures departure of the system state from the symmetric
one. That is, ψ = 0 in the symmetric phase and non-zero in the broken-symmetry
phase.

53
54 CHAPTER 3. GINZBURG-LANDAU THEORY

Re Ψ
∆ χ Im Ψ

Figure 3.1: Below the transition temperature, the free energy Eq. (3.1) has a minimum
at a nonzero order parameter magnitude. The minimum energy is degenerate with
respect to the order parameter phase χ.

For example, for the transition from the paramagnetic to ferromagnetic state the
magnetic moment M can be used as an order parameter. Here the symmetry with re-
spect to space rotations is broken. For the superconducting transition the gap 1 (or a
quantity, proportional to it) is a good order parameter. Since 1 measures expectation
value of the product of two creation operators, Eq. (2.24), the superconducting state
with 1 6= 0 violates particle number conservation. The symmetry behind this con-
servation law, usually called global U(1) symmetry, is broken in the superconducting
state.
Close to the transition, where the order parameter is small, Landau suggested to
expand the free energy of the system in terms of the order parameter, keeping only
contributions which are compatible with the symmetry of the system in the symmetric
phase. For the uniform superconductor in the absence of magnetic field the proper
expansion is
β
Fsn = Fs − Fn = α|1|2 + |1|4 . (3.1)
2
Here α and β are expansion parameters, generally dependent on temperature, pressure
and other relevant conditions. To find the equilibrium value of 1 we have to minimize
the free energy. In order to have the minimum at 1 = 0 when T > Tc one should have
α(T ) > 0 at T > Tc . For the minimum at 1 6= 0 one should have α(T ) < 0, β > 0 at
T < Tc . Close to the transition we can thus write

α(T ) = α ′ (T − Tc ), β(T ) ≈ β(Tc ) > 0 . (3.2)


3.2. GINZBURG-LANDAU EQUATIONS 55

Minimizing Fsn at T < Tc we obtain

α α ′ (Tc − T ) α2
|1|2 = 12GL = − = , Fsn = − . (3.3)
β β 2β

Thus at T → Tc we have |1| ∝ (1 − T /Tc )1/2 as we found from the BCS theory in
Problem 2.2. Comparing (3.3) with the result of this problem we find

α 8π 2 2
= k Tc (T − Tc ) . (3.4)
β 7ζ(3) B
Comparing Fsn from (3.3) with equation (1.2) we get
 
α2 H2 p T
= c , Hc = 2α ′ Tc π/β 1 − . (3.5)
2β 8π Tc

Using expression for Hc from the BCS theory derived in Problem 2.3 we find

α2 8π 2
= N(0)kB2 (T − Tc )2 . (3.6)
β 7ζ(3)
From equations (3.4) and (3.6) we obtain expressions for α and β from the microscopic
theory
T − Tc 7ζ(3)N(0)
α = N(0) , β= . (3.7)
Tc 8π 2 kB2 Tc2
Note that the condition of the minimum of the free energy fixes only the amplitude
(3.3) of the order parameter 1 = |1|eiχ . With respect to the phase χ the energy is
degenerate. This is illustrated in Fig. 3.1. The fact that the system chooses a particular
phase among a set of equivalent values is the manifestation of the broken symmetry.
(In the case of ferromagnets this is the choice of a particular direction of M among
equivalent ones.) Since the phase factor eiχ is a 1 × 1 unitary matrix, this symmetry is
called U(1) symmetry.

3.2 Ginzburg-Landau equations


To describe non-uniform superconductors and superconductors in the external field,
Ginzburg and Landau added to the Landau free energy (3.1) terms depending on the
gradient of the order parameter and on the magnetic field:
  2
2 β 4
2e h2
FGL = α|1| + |1| + γ −i h̄∇ − A 1 +
, (3.8)
2 c 8π
Z
FGL [1, A] = FGL dV . (3.9)
56 CHAPTER 3. GINZBURG-LANDAU THEORY

In principle, the GL model can be considered as a generalization of the London model


(1.6): Energy of the superconducting state in the absence of currents and fields is taken
from the Landau theory of phase transitions, while the gradient term resembles the
kinetic energy of superconducting electrons in view of results of Sec. 2.11 and in par-
ticular Eq. (2.86) for vs . Another way to look at the gradient term in (3.8) is to consider
it as a first term of expansion of the free energy over the gradients of the order param-
eter, written in a gauge-invariant form. Of cause, the fact that 2e enters the expression
(3.8) cannot be derived from general considerations, but reflects the microscopic origin
of superconductivity.
As in section 1.4 the free energy FGL is a functional of the fields 1 and A. To
minimize it we take the variation and equals it to zero. Since 1 is complex, 1 and
1∗ are linearly independent. One can check, though, that variations with respect to 1
and 1∗ give equations which are complex conjugate of each other. Thus we will take
variation only with respect to 1∗ . We have

δ1∗ (|1|2 ) = δ1∗ (11∗ ) = 1δ1∗ ,


δ1∗ (|1|4 ) = δ1∗ (12 1∗2 ) = 212 1∗ δ1∗ = 2|1|21δ1∗ ,

and
  2     
2e 2e 2e ∗
δ1∗ −i h̄∇ − A 1 = −i h̄∇ − A 1
i h̄∇ − A |{z}
δ1
c c c
| {z } a
X
   
2e 2e
= i h̄X ∇a + − A X a = i h̄a(−∇X) + i h̄ div(aX) + a − A X
c c
 
2e
= a −i h̄∇ − A X + i h̄ div(aX)
c
 2    
∗ 2e ∗ 2e
= δ1 −i h̄∇ − A 1 + i h̄ div δ1 −i h̄∇ − A 1 , (3.10)
c c

where we used vector identity X ∇a = −a div X + div(a X) and we note that notations
div X and ∇X are equivalent. Combining all expressions we have
Z (  2 )
∗ 2e 2
δ1∗ FGL = dV δ1 γ −i h̄∇ − A 1 + α1 + β|1| 1 (3.11)
c
Z  
∗ 2e
+ i h̄γ dS δ1 n · −i h̄∇ − A 1 . (3.12)
c

Here we converted volume integral from divergence to the surface integral, n is the
normal to the surface.
3.2. GINZBURG-LANDAU EQUATIONS 57

Requiring δ1∗ FGL = 0 for arbitrary δ1∗ we get from the volume term (3.11) the
Ginzburg-Landau equation
 2
2e
γ −i h̄∇ − A 1 + α1 + β|1|21 = 0 (3.13)
c
and from the surface term (3.12) the boundary condition
 
2e
n · −i h̄∇ − A 1 = 0 . (3.14)
c
Note that this boundary condition we obtained from FGL which does not include the
energy of interaction between the superconductor and its surroundings. Thus condition
(3.14) is applicable, say, on the boundary between the superconductor and vacuum.
More general consideration leads to the condition
 
2e i
n · −i h̄∇ − A 1 = 1 , (3.15)
c bs
where the parameter bs depends on the material which is in contact with the supercon-
ductor.
Now let us take variation with respect to A. We have
b a
z }| { z}|{
2 2
δA (h ) = δA (curl A) = 2 curl A curl δA = 2 (div[δA × curl A] + δA curl curl A) .

Here we used vector identity div[a × b] = b curl a − a curl b. We also have


  2       
2e 2e 2e
δA −i h̄∇ − A 1 = δA −i h̄∇ − A 1 i h̄∇ − A 1∗
c c c
     
2e 2e
+ −i h̄∇ − A 1 δA i h̄∇ − A 1∗
c c
     
2e 2e 2e
= − δA 1∗ −i h̄∇ − A 1 + 1 i h̄∇ − A 1∗ .
c c c
Thus we obtain
Z       
j 2e 2e 2e
δA FGL = dV δA − γ 1∗ −i h̄∇ − A 1 + 1 i h̄∇ − A 1∗
c c c c
Z
1
+ dS n · [δA × curl A] . (3.16)

Here we used Maxwell equation j = (c/4π) curl h = (c/4π) curl curl A. Setting this
variation to zero we obtain the expression for the supercurrent
     
2e 2e
j = 2eγ 1∗ −i h̄∇ − A 1 + 1 i h̄∇ − A 1∗ (3.17)
c c
 
2e
= 4eγ |1|2 h̄∇χ − A , (3.18)
c
58 CHAPTER 3. GINZBURG-LANDAU THEORY

where 1 = |1|eiχ . By comparing this expression with results of section 2.11 and
Problem 2.7 we can find the coefficient γ in the clean limit, see Problem 3.1.
In general case the coefficient γ depends on purity of the sample. The purity is
characterized by the parameter xs = τs kB Tc /h̄ ∼ τs /τp , where τs is the electronic
mean free time due to the scattering by impurities and τp is the characteristic time
of the pairing interaction, see Sec. 1.9. Superconductors are called clean when this
parameter is large, and they are dirty in the opposite case. One has

N(0)πD
γ = y(xs ) (3.19)
8h̄kB Tc

where D = vF2 τs /3 is the diffusion coefficient, and



8 X 1
y(xs ) = (3.20)
π2 (2n + 1)2 [(2n + 1)2πxs + 1]
n=1

This function is y = 1 for the dirty limit xs ≪ 1, and it is

7ζ(3)
y=
2π 3 xs
for the clean case xs ≫ 1. Here ζ(3) ≈ 1.202 is the Riemann zeta function. Therefore
(
πN(0)D/8kB Tc h̄ , dirty
γ = 2 2 2
(3.21)
7ζ(3)N(0)vF /48π (kB Tc ) , clean

so that γdirty/γclean ∼ xs ≪ 1.
In the end we note that the GL functional (3.9) can be written in a different useful
form. Using the same transformations as in (3.10) but keeping 1∗ instead of δ1∗ we
get
Z "  2 #
2 β 4 ∗ 2e h2
FGL = dV α|1| + |1| + γ 1 −i h̄∇ − A 1 +
2 c 8π
Z  
2e
+ i h̄γ dS 1∗ n · −i h̄∇ − A 1
c
Z " (  2 )#
β h 2 2e
= dV − |1|4 + + 1∗ γ −i h̄∇ − A 1 + α1 + β|1|2 1
2 8π c
" # "    2 #
|1| 4
Z Z
β 4 h2 Hc2 h
= dV − |1| + = dV − + (3.22)
2 8π 8π 1GL Hc

where the surface integral is zero due to the boundary condition (3.14) and the part in
curly braces is zero due to the GL equation (3.13).
3.3. COHERENCE LENGTH AND PENETRATION DEPTH 59

3.3 Coherence length and penetration depth


Let us introduce the dimensionless order parameter ψ̃ so that 1 = ψ̃1GL and the
equilibrium order-parameter magnitude
r s
α |α|
1GL = − = . (3.23)
β β
Then the equation (3.13) becomes
 2
2ie
−ξ 2 ∇ − A ψ̃ − ψ̃ + |ψ̃|2 ψ̃ = 0 . (3.24)
h̄c
Here the length
T −1/2
r  
p γ
ξ(T ) = h̄ γ /|α| = h̄ 1 − (3.25)
α ′ Tc Tc
defines the characteristic length scale of variation of the order parameter. It is called
the Ginzburg-Landau coherence length. Using expressions (3.7) and (3.21) and the
definition of the zero-temperature coherence length ξ0 from (1.23) we find in the clean
limit
T −1/2
   
7ζ(3) 1/2
ξ(T ) = ξ0 1 − (3.26)
12 Tc
and in the dirty limit
T −1/2
 
π p
ξ(T ) = √ ξ0 ℓ 1 − . (3.27)
2 3 Tc
Here ℓ = vF τs is the mean free path between scatterings. Since xs = ℓ/2πξ0 , the
clean limit corresponds to ℓ ≫ ξ0 and the dirty limit to ℓ ≪ ξ0 . Thus in dirty materials
ξ(T ) is generally much smaller that in clean materials.
Comparing expressions (3.18) and (1.16) for the supercurrent we see that they co-
incide if we define the Ginzburg-Landau superconducting density as

ns = 8γ m|1|2 . (3.28)

We already saw that equation (1.16) [and thus equation (3.18) with definition (3.28)]
explains the Meissner effect and gives the penetration depth value
!1/2 s
T −1/2
 
mc2 c c β
λL = = √ = 1 − . (3.29)
4πns e2 4|e| 2πγ |1| 4|e| 2πα ′ γ Tc Tc

We see that within the GL model both λL (T ) and ξ(T ) have the same temperature
dependence. Their ratio is called the Ginzburg-Landau parameter
r
λL (T ) c β
κ= = . (3.30)
ξ(T ) 4|e|h̄γ 2π
60 CHAPTER 3. GINZBURG-LANDAU THEORY

We can also express the critical field (3.5) through ξ and λL :

h̄c 80
Hc = √ = √ . (3.31)
2 2|e|λL ξ 2 2πλL ξ

The GL model is applicable when |1(T )| ≪ |1(T = 0)| ∼ kB Tc . In particular, if


the amplitude of the order parameter is at the equilibrium value 1GL the requirement
becomes Tc − T ≪ Tc . Another requirement is the local electrodynamics, so that the
penetration depth λL (T ) ≫ ξ0 . As we see from (3.29), λL (T ) → ∞ when T → Tc .
So this condition is also satisfied sufficiently close to Tc .
Since ξ(T ) limits the maximum gradient of the order parameter also for its phase,
(∇χ)max ∼ 1/ξ(T ), we can define the characteristic maximum current density, the
Ginzburg-Landau critical current, from Eq. (3.18) as

4|e|h̄γ 12GL h̄c2


jcGL = 4|e|γ 12GLh̄(∇χ)max = = ∝ (1 − T /Tc )3/2 . (3.32)
ξ 8π|e|λ2L ξ

Sometimes it is convenient to normalize the order parameter in such a way that



|ψ|2 = ns /2, ψ = 2 γm1. (3.33)

This is the same normalization which we used for the wave function ψ in section 1.6.
The GL free energy becomes
  2 2
b 4
2 1 2e + h ,

FGL = a|ψ| + |ψ| + −i h̄∇ − A ψ (3.34)
2 4m c 8π
Z Z " # Z
b 4 h2 Hc2 h i
FGL = FGL dV = − |ψ| + dV = −|ψ̃|4 + (h/Hc )2 dV ,
2 8π 8π
(3.35)
where
α β
a= , b= . (3.36)
4mγ 16m2 γ 2
The coherence length, the penetration depth and the equilibrium magnitude of the order
parameter become
!1/2 r
h̄ mc2 |a|
ξ(T ) = √ , λL = 2
, ψGL = , (3.37)
4m|a| 8πe2 ψGL b

and the current  


e 2e
j= |ψ|2 h̄∇χ − A . (3.38)
m c
For the dimensionless order parameter ψ̃ = ψ/ψGL equation (3.24) obviously holds.
3.4. CRITICAL FIELD OF A SUPERCONDUCTING SLAB 61

z y
H

d/2
j
−d/2 x

Figure 3.2: Superconducting slab in the parallel magnetic field.

3.4 Critical field of a superconducting slab


As an application of the GL theory let us perform calculation of the critical field for
an infinite slab with thickness d < ξ(T ) in the magnetic field H applied parallel to
the slab. This was one of the first prediction of the GL theory which was checked
experimentally and which demonstrated the superiority of the GL approach compared
to the London model. Other simple applications of the GL theory one can find in
Problems 3.3 and 3.4.
The setting is shown in Fig. 3.2: The slab occupies space between z = −d/2 and
z = d/2, the field is in the y direction and the current and the vector potential are in
the x direction. First we remind the prediction of the London model which was found
in problem 1.6:
d −1
 
2 2λL
Hc,slab = Hc2 1 − tanh . (3.39)
d 2λL
Here Hc is the thermodynamic critical field. In the London model the transition at
Hc,slab is always of the first order, since ns is not allowed to change. In the GL model
we may allow ns = 2|ψ|2 to deviate from the equilibrium value. Since we consider
slabs thinner than the coherence length we can assume that the order parameter is
constant in the slab: ψ = ψ̃ψGL with ψ̃ = const but not necessarily unity. For
ψ = const we have from the equation (3.38)

2e2 c
− |ψ|2 A = j = curl curl A
mc 4π
or
∂ 2A 8πe2 ψGL
2
2 ψ̃ 2
= ψ̃ A = A.
∂z2 mc2 λ2L
The general solution of this equation is

A = A1 ezψ̃/λL + A2 e−zψ̃/λL .

The constants A1 and A2 should be determined from the boundary condition for the
62 CHAPTER 3. GINZBURG-LANDAU THEORY

magnetic field h = ∂A/∂z = H at z = ± d/2. We find

cosh(ψ̃z/λL ) H λL sinh(ψ̃z/λL )
h=H , A= . (3.40)
cosh(ψ̃d/2λL ) ψ̃ cosh(ψ̃d/2λL )
In the fixed external field we should consider the Gibbs free energy G = F − HB/4π
with B = hhi. We have per unit area of the slab

Zd/2  
hH H2
Gsn = Gs (H ) − Gn (H ) = dz FGL − +d
4π 8π
−d/2

Zd/2 "  2 #
b 1 2e h2 − 2hH H2
= 2
dz a|ψ| + |ψ|4 + A 2
|ψ| + +d
2 4m c 8π 8π
−d/2

Zd/2 " #
1 A2 ψ̃ 2 H2
= dz Hc2 (−2ψ̃ 2 4
+ ψ̃ ) + + h2
− 2hH + d
8π λ2L 8π
−d/2
 
Hc2 2 4 H2 2λL
=d (−2ψ̃ + ψ̃ ) + − tanh(ψ̃d/2λL ) + d
8π 8π ψ̃

To find ψ̃ we minimize Gsn by setting ∂Gsn /∂ ψ̃ = 0 which gives the equation


 2 " #−1
H 2 2 2 sinh(ψ̃d/λL )
= 4ψ̃ (1 − ψ̃ ) cosh (ψ̃d/2λL ) −1 , (3.41)
Hc ψ̃d/λL

which is valid for any H including H = Hc,slab. The critical field corresponds to
Gsn = 0 which gives the equation
 2 " #−1
Hc,slab 2 2 tanh(ψ̃d/2λL )
= ψ̃ (2 − ψ̃ ) 1 − . (3.42)
Hc ψ̃d/2λL

Equations (3.41) and (3.42) give the solution of the problem in the implicit form.
Let us consider limiting cases. If d ≫ λL then ψ̃ ≈ 1 and from Eq. (3.42) we get
the same result as in the London model (3.39). Using condition d ≫ λL it can be
simplified as
Hc,slab = Hc (1 + λL /d) . (3.43)

Now let us consider thin films, d ≪ λL . Expanding hyperbolic functions in


Eq. (3.41) we get

2 4ψ̃ 2 (1 − ψ̃ 2 ) λ2L
Hc,slab = Hc2 = 24Hc2 (1 − ψ̃ 2 ) , (3.44)
(ψ̃d/λL )2 /6 d2
3.4. TYPE I AND TYPE II SUPERCONDUCTORS 63

3
d

Figure 3.3: Measurements of the critical field


of the tin films of different thicknesses d versus
2 temperature-controlled London penetration depth
λL (symbols) compared to the GL model for
Hc,slab /Hc

√ Hc,slab (solid line). Broken line shows the hys-


2 6λL /d
teresis in the transition observed for thicker films,
1 indicating that the transition is of the first order.
The hysteresis disappears for the thin films, which
hysteresis
is consistent with the transition of the second or-
der (N.V. Zavaritski, 1952).
0 0.2 0.4 0.6
λL(T )/d

and from Eq. (3.42) we get

2 ψ̃ 2 (2 − ψ̃ 2 ) λ2L
Hc,slab = Hc2 = 12Hc2 (2 − ψ̃ 2 ) . (3.45)
(ψ̃d/2λL )2 /3 d2

Solving these two equations we obtain

ψ̃ = 0 ,

i.e. the transition is of the second order and


√ λL
Hc,slab = 2 6 Hc ≫ Hc . (3.46)
d
With decreasing film thickness the transition changes from the first order to the second

order at d = 5λL , see Problem 3.6.
These predictions of the GL theory were confirmed by the experiment, see Fig. 3.3.

3.5 Energy of the normal–superconducting boundary.


Type I and type II superconductors
If one places a large slab in the perpendicular magnetic field, then the magnetic field
starts to penetrate into the slab already at H < Hc , see Fig. 3.4, since expelling the
field in this geometry would require screening currents to exceed the critical magni-
tude. Thus normal and superconducting regions in the same material may coexist. It
is clear, that for the stable situation the field in normal regions should be equal to Hc .
64 CHAPTER 3. GINZBURG-LANDAU THEORY

S
S N N S N S
B=0

B = Hc

B = H < Hc

Figure 3.4: Intermediate state of type I superconductors. In the external field H < Hc
the sample is divided to normal and superconducting domains so that in the normal
phase magnetic induction B = Hc while in the superconducting domains magnetic
field is absent.

When doing calculation of the stable configuration it turns out that additionally one
has to require that normal-superconducting (NS) boundary have some extra energy,
associated with it. This energy can be determined experimentally by comparing the
experimentally observed NS patterns to the calculated ones. Where does this energy
originate?
Let us consider this problem within the GL theory. We will look at a single bound-
ary in the field H = Hc and will assume the system to be homogeneous in y and z
directions, and to be fully superconducting at x → −∞ and normal at x → +∞. The
goal is thus to solve the GL equations (3.24) and (3.38) with the boundary conditions

ψ̃ → 1, h → 0, at x → −∞,
(3.47)
ψ̃ → 0, h → Hc , at x → +∞.

In the general case this can be done only numerically. But the qualitative picture of
the solution is clear: ψ̃ changes from 1 to 0 in the layer of thickness ξ and h changes
from 0 to Hc in the layer of thickness λL , Fig. 3.5. What energy is associated with this
profile of the order parameter?
Remember that we have to consider the Gibbs free energy. For the pure super-
conducting or the pure normal state at H = Hc this energy is equal to −Hc2 /8π, see
Sec. 1.3. Let us consider this value as a zero of energy. Using (3.35) we get

Z∞ " # Z∞ "  2 #
hH Hc2 H2 h − Hc Hc2
G= dx FGL − + = c 4
dx −|ψ̃| + = δ,
4π 8π 8π Hc 8π
−∞ −∞
(3.48)
3.5. TYPE I AND TYPE II SUPERCONDUCTORS 65

h(x) ∼ λL 9(x) h(x) 9(x)

∼ λL
(a) (b)

0 x 0 x
∼ξ ∼ξ

Figure 3.5: Change of the order parameter amplitude |ψ(x)| and of the local magnetic
field h(x) at the boundary between coexisting normal and superconducting domains.
√ √
(a) Type I superconductor with 2λL < ξ . (b) Type II superconductor with 2λL >
ξ.

where δ is a characteristic quantity of the length dimension. Using conditions (3.47)


we see that the expression under integral in (3.48) is zero everywhere except the NS
boundary.
In the case ξ ≫ λL , i.e. κ ≪ 1, Fig. 3.5a, we have a region of size ∼ ξ − λL where
|ψ̃| and h are both small which makes a positive contribution to the integral (3.48)
and thus δ ∼ ξ − λL . Exact calculation in this limit (Problem 3.8) gives δ ≈ 1.89ξ .
Superconductors with the positive NS energy are called type I superconductors. In this
case the picture of reasonably thick normal and superconducting slices in Fig. 3.4 has
sense.
In the opposite case ξ ≪ λL , i.e. κ ≫ 1, Fig. 3.5b, we have a region of size ∼ λL −
ξ where |ψ̃| ∼ 1 and h ∼ Hc which makes a negative contribution to the integral (3.48)
and thus δ ∼ −(λL − ξ ) = ξ − λL . Exact calculation in this limit gives δ ≈ −1.104λL.
Superconductors with negative NS energy are called type II superconductors. In this
case the picture of reasonably thick normal and superconducting slices in Fig. 3.4 can
not be stable: number of NS boundaries will tend to increase and the size of normal
regions to decrease. The limit to this process is set by the flux quantization: A normal
region with the trapped magnetic field which is surrounded by the superconductor can
at minimum carry one quantum of circulation 80 and thus it cannot disappear. Such
objects are called quantized vortices and we will consider them further in this chapter.
Special consideration shows that the transition from the positive to the negative
energy of the NS boundary occurs at
1
κ= √ .
2
√ √
Thus superconductors with κ < 1/ 2 are of type I, and those with κ > 1/ 2 are of
type II.
What material parameters determine value of κ? In the clean limit we have from
66 CHAPTER 3. GINZBURG-LANDAU THEORY

equations (3.30), (3.7), (3.21) and (2.6)


r r s
c β 3c π k B Tc 3π 2 h̄c kB Tc e2 m
κ= = = √
4|e|h̄γ 2π 7ζ(3)N(0) vF2
|e|h̄ 14ζ(3) e2 EF π h̄pF
s
3π 2 h̄c kB Tc e2 /a0 3π 2 h̄c kB Tc k B Tc
= √ ≈ √ ∼ 103 . (3.49)
14ζ(3) e2 EF EF 14ζ(3) e2 EF EF

Here we used

mvF2 p2 2π h̄ pF2 a0
EF = = F, a0 = , pF = = (2mEF )
2 2m pF pF 2π h̄

and the fact that the ratio of the potential energy of conducting electrons, e2 /a0 , to their
kinetic energy EF is of the order of unity for good metals, although it may become
larger for systems with strong correlations between the electrons. The last factor in
Eq. (3.49) is usually small: kB Tc /EF is below 10−3 for usual superconductors, but it
is of the order of 10−1 − 10−2 for high temperature superconductors with Tc ∼ 100K
and EF ∼ 1000K.
We see that for usual clean superconductors the Ginzburg-Landau parameter is nor-
mally small, though, in some cases it may be of the order of 1. On the contrary, for
high temperature superconductors, which have a tendency to be strongly correlated
systems with a not very low ratio of kB Tc /EF , the parameter κ is usually very large.
The Ginzburg-Landau parameter increases for dirty superconductors with xs ≪ 1:

κdirty ∼ κclean /xs . (3.50)

Therefore, dirty alloys normally have a large κ.

3.6 Abrikosov vortices. Critical field Hc1


Magnetic field penetrates into type II superconductors in the form of small regions
which carry one quantum of magnetic flux 80 each. As can be seen from Eq. (1.17),
this corresponds to the increase of the order-parameter phase χ by 2π on a loop around
this region. This phase winding is a topological invariant: It cannot be continuously
reduced to zero. Homotopy group theory connects this topological invariant to exis-
tence of linear defects of the order parameter: continuous lines where ψ = 0 and
around which the phase of ψ winds by 2π. In a superconductor such linear defects
carry magnetic flux and are called Abrikosov vortices, Fig. 3.6. They were theoreti-
cally predicted by A. Abrikosov in 1957. As a topological object, a vortex cannot end
in superconductor bulk: It can only end at the boundary or form a closed loop.
3.6. ABRIKOSOV VORTICES. CRITICAL FIELD HC1 67

Figure 3.6: Abrikosov vortices in type II superconductor are topologically-protected


linear defects of the order parameter: Order parameter is zero at the vortex axis and the
phase of the order parameter winds by 2π on a loop around the vortex. Each vortex
carries a single quantum of magnetic flux.

Let nv be density of vortex lines in a plane perpendicular to the direction of the


magnetic field (and to vortex axes). Since each vortex carries flux 80 , the magnetic
induction is simply
B = 80 nv (3.51)
and the magnetization of the sample
80 nv − H 80
M= = nv + M0 ,
4π 4π
where M0 = −H /4π is the magnetization (1.14) with the complete Meissner effect.
From the magnetization curve of type II superconductor in Fig. 1.4 we see that in
the fields slightly above Hc1 difference M − M0 is small and it goes to zero when
H → Hc1 . Thus in this limit vortex density nv is small, vortices are far apart and we
can consider each of them individually.
Let us study within the GL theory a single straight vortex. We place it at r = 0
of the cylindrical coordinate system (r, φ, z) and assume the problem to be axially-
symmetric and uniform in the z direction. Thus we can write

ψ = ψGL ψ̃, ψ̃ = f (r) eiφ , (3.52)


h = h(r) ẑ . (3.53)

We can choose
A = A(r) φ̂ (3.54)
so that
1 d
h(r) = rA(r) (3.55)
r dr
68 CHAPTER 3. GINZBURG-LANDAU THEORY

and
c c dh(r)
j= curl h = − φ̂ . (3.56)
4π 4π dr
Inserting (3.52) and (3.54) into the GL equation (3.24) we get the equation
" #
2 ′′ f′ 4e2 2
ξ f + − 2 Q f + f −f3 = 0, (3.57)
r h̄ c2

where the prime denotes derivative over r and we introduced function Q(r) so that

h̄c
A(r) = Q(r) + . (3.58)
2er
Note that
1 d
h(r) = rQ(r) . (3.59)
r dr
Now inserting equations (3.52) and (3.54) into the current expression (3.38) and ex-
pressing the current via Q from (3.56) and (3.59) we get the equation

Q′ Q f 2Q
Q′′ + − 2 − 2 = 0. (3.60)
r r λL

We obtained the system of two equations (3.57) and (3.60) for two functions f (r) and
Q(r) with boundary conditions

f → 0 and A finite, r → 0,
(3.61)
f → 1 and h → 0, r → ∞.

In general case the system can be solved only numerically. An example of the solution
is plotted in Fig. 3.7. It is clear that when λL ≫ ξ , i.e. κ ≫ 1 there are three distinct
regions in the vortex structure:
(i) r ≫ λL . Here f = 1 and h is decaying exponentially with increasing distance
from the vortex axis.
(ii) ξ ≪ r ≪ λL . Here 1 − f ≪ 1 and h is changing.
(iii) r ≪ ξ . Here f ≪ 1 and increasing with r while h ≈ const.
The region r . ξ where the order parameter substantially deviates from the equilibrium
value is called the vortex core.
In the each of the regions (i)–(iii) we can solve the equations (3.57) and (3.60)
analytically. We start with equation (3.60) in the regions (i) and (ii). Here we can set
f = 1 and the Eq. (3.60) becomes the modified Bessel equation. We need the solution
which decays when r → ∞. It is

Q(r) = Q0 K1 (r/λL ) . (3.62)


3.6. ABRIKOSOV VORTICES. CRITICAL FIELD HC1 69

0 1 2 3 r/λL
2

9(r)/9GL
1

h(r)/Hc1
0
0 5 10 15 r/ξ

Figure 3.7: Dependence of the order-parameter amplitude and of the magnetic field on
the distance from the axis of the Abrikosov vortex found from numerical solution of
the GL equations for κ = 5.

Here K1 is the modified Bessel function of the second kind with the asymptotic behav-
ior1

K1 (x) = 1/x, x ≪ 1,
r
2 −x
K1 (x) = e , x ≫ 1,
πx
and the constant Q0 is determined from the boundary condition at r → 0
λL h̄c
A = Q0 + finite,
r 2er
which gives
h̄c
Q0 = − . (3.63)
2eλL
Magnetic field we find from equation (3.59)
h̄c 80
h(r) = 2
K0 (r/λL ) = − K0 (r/λL ) . (3.64)
2eλL 2πλ2L

Note that with our choice of the direction of phase winding in Eq. (3.52) the magnetic
field turns out to be directed against the z axis. The asymptotic behavior of K0 is

K0 (x) = − ln x, x ≪ 1,
r
2 −x
K0 (x) = e , x ≫ 1,
πx
1 Digital Library of Mathematical Functions http://dlmf.nist.gov is highly recommended.
70 CHAPTER 3. GINZBURG-LANDAU THEORY

In the region (iii) the magnetic field is approximately constant and thus is equal to the
value given by Eq. (3.64) at r = ξ .
Now we consider equation (3.57) in the regions (ii) and (iii). Here
λL h̄c
Q(r) = Q0 =−
r 2er
and equation (3.57) becomes
 
2 ′′ f′ f
ξ f + − 2 + f − f 3 = 0. (3.65)
r r
In the region (iii) f ≪ 1, we can omit f 3 in this equation. Then it becomes the Bessel
equation with the solution

f = CJ1 (r/ξ ) = f0 r/ξ . (3.66)

The constant f0 ∼ 1 can be found only numerically. In the region (ii) we put g =
1 − f ≪ 1 into Eq. (3.65) and obtain in the first order
ξ2
g= . (3.67)
2r 2
We can summarize all findings in the following table
r ≪ξ f = f0 r/ξ h = −h0 ln(λL /ξ )

ξ ≪ r ≪ λL f = 1 − ξ 2 /(2r 2 ) h = −h0 ln(λL /r) (3.68)


p
λL ≪ r f =1 h = −h0 2λL /πre−r/λL
where
80
h0 = .
2πλ2L
Now we can calculate the energy of the vortex compared to the state when there is
no vortex (f = 1, h = 0). We have from Eq. (3.35)
Z∞ h
H2 i
Fv = vortex
FGL no vortex
− FGL = c 1 − f 4 + (h/Hc )2 2πrdr . (3.69)

0
The integral can be evaluated separately in the ranges (0, ξ ), (ξ, λL ) and (λL , ∞) using
expressions (3.68). It is easy to see that among the six terms [(1 − f 4 ) and (h/Hc )2
contributions in each of the three ranges] the largest by the factor of ln κ comes from
the 1 − f 4 term in the (ξ, λL ) range. The reason is that it results in a logarithmically
divergent integral, while all other terms are regular. We thus have
ZλL ZλL
H2 4 Hc2 H2 λL 820
Fv ≈ c (1 − f ) 2πr dr = 4g r dr = c ξ 2 ln = ln κ .
8π 4 2 ξ 16π 2 λ2L
ξ ξ
(3.70)
3.7. INTERACTION OF AN ABRIKOSOV VORTEX WITH ELECTRIC CURRENT71

The first critical field Hc1 corresponds to penetration of the first vortices to the
superconductor. It is thus determined by the condition that the Gibbs free energy of the
state without vortices is equal to the energy of the state with the far separated vortices.
We have

0 = Gvortex(Hc1 ) − Gno vortex(Hc1 ) =


1 1
nv Fv − (B vortex − B no vortex )Hc1 = nv Fv − nv 80 Hc1 . (3.71)
4π 4π
From this equation we obtain

80 ln κ
Hc1 = 2
ln κ = Hc √ . (3.72)
4πλL 2κ

Comparing to (3.68) we find that in the vortex core h(0) = 2Hc1 . Note also that the
energy difference in (3.71) does not depend on the density of vortices in the approx-
imation of individual vortices. That means that at Hc1 vortices start to penetrate in
the sample until their mutual interaction breaks the condition (3.71). This corresponds
to the vertical derivative dM/dH at H = Hc1 on the magnetization curve of type II
superconductor, Fig. 1.4.

3.7 Interaction of an Abrikosov vortex with electric cur-


rent
Appearance of vortices changes the electrical and magnetic properties of a supercon-
ductor quite substantially. In particular, if vortices are free to move, then the resistivity
becomes non-zero. We can see this by considering interaction of a vortex with applied
electric current jex within the London model, Eq. (1.6). Let vortex be in the position
r0 . The current around the vortex is j = j(r − r0 ). The current makes the contribution
to the energy
m ex 2 m h2 ex 2 ex
i
(j + j ) = j + (j ) + 2j j .
2ns e2 2ns e2
The first two terms here are the energy of the vortex and of the external current. The
last term is the interaction energy
m ex
Fint = j j. (3.73)
ns e2
For the force per unit length of the vortex we have
Z Z
m
fL = −∇r0 Fint dS = − ∇r jex j dS ,
ns e2 0
72 CHAPTER 3. GINZBURG-LANDAU THEORY

where the integration is in the plane perpendicular to the vortex axis. Applied current
jex is independent of r0 while ∇r0 j = −∇r j. We replace ∇r0 with −∇r , considering jex
to be constant, and use vector identity ∇(Ca) = (C∇)a + C × curl a valid for constant
vector C:
Z Z Z 
m ex m ex ex
fL = ∇(j j) dS = [j × curl j] dS + (j ∇)j dS =
ns e2 ns e2
Z Z  Z
m ex ex m
[j × curl j] dS − j div j dS = [jex × curl j] dS .
ns e2 ns e2
Here we integrated by parts the second term and use div jex = 0. Using result of
Problem 3.9 we find
c80
curl j = δ(r − r0 )b ,
4πλ2L
where b is the unit vector along the direction of the vortex. Finally we have
80 ex
fL = [j × b] . (3.74)
c
This is the Lorentz force acting on the vortex. Multiplying it on the vortex density nv
we find the force FL acting per unit volume. Since magnetic induction B = nv 80 b we
have
FL = c−1 [jex × B] . (3.75)
Under action of this force vortices may start to move with velocity vL . Let us assume
that they experience viscous damping force fv = −ηvL . In steady motion fL + fv = 0
and thus
vL = (80 /ηc)[jex × b] . (3.76)
In the frame moving with vortices there is magnetic field B. Thus in the laboratory
frame electrical field appears

E = c−1 [B × vL ] . (3.77)

Substituting (3.76) and taking into account b = B/B and B ⊥ jex we obtain

E = (80 B/ηc2 )jex . (3.78)

But appearance of the electric field along the current means that the resistivity is no
longer zero:
ρ = E/j ex = 80 B/ηc2 . (3.79)
Thus superconducting properties are lost.
In order to preserve the zero resistivity vortices should be immobile. This can be
achieved by pinning: In an inhomogeneous superconductor vortex energy may depend
3.8. UPPER CRITICAL FIELD HC2 73

on the vortex position. A vortex thus can settle to a local minimum of energy and then
application of the current above some critical value would be required before the vortex
will start to move.
Note that the current jex may also be produced by another vortex. This results in
the interaction between vortices. Using result of Problem 3.9 we find that two vortices
with the same direction of flux repel each other with the force fL ∝ δ −1 inversely
proportional to the distance between them δ, if δ < λL . If δ ≫ λL , then the interaction
is much smaller. In a sense this resembles the interaction between ions in a metal:
They repel each other at small distances due to Coulomb interaction while at larger
distances the Coulomb repulsion is screened. With this analogy we expect that at lower
temperatures vortices tend to form a crystalline-like lattice (or, if disorder is strong, a
glass-like state). At higher temperatures, if thermal fluctuations are strong enough and
vortices are sufficiently mobile, the lattice can melt and vortex liquid is formed. The
vortex matter including dynamics of vortices in superconductors is by itself a large area
of research, which we won’t discuss in this course.

3.8 Upper critical field Hc2


When in a type II superconductor magnetic field is increased above Hc1 , distance be-
tween Abrikosov vortices decreases. At sufficiently high fields vortices may come
closer to each other then the coherence length ξ , and their cores would start to overlap.
Since in the core the amplitude of the order parameter is suppressed, the average am-
plitude of the order parameter will start to decrease in this situation, until eventually it
will become zero everywhere at some field which is called Hc2 . Thus one can expect
that for the type II superconductors transition to the normal state in the field is of the
second order.
One can arrive to the same conclusion also from a different point of view. We have
seen in Sec. 3.4 that for a thin superconducting slab in the parallel field, the critical
field may significantly exceed Hc . So why the sample should go completely normal
at the field Hc , while it can split to thin normal and superconducting slabs? In type
I superconductors this scenario is prevented by positive energy of the NS interface,
but this limitation is not applicable to type II superconductors. Thus we conclude that
in the type II superconductor transition to the normal state occurs only when even
infinitesimally small superconducting region cannot exist.
We can find the corresponding critical field Hc2 using GL theory. Since in our case
|ψ| ≪ ψGL , i.e. |ψ̃| ≪ 1, we can linearize the GL equation (3.24):
 2
2ie
−ξ 2 ∇ − A ψ̃ − ψ̃ = 0 . (3.80)
h̄c
74 CHAPTER 3. GINZBURG-LANDAU THEORY

Additionally, such a small superconducting region cannot affect the magnetic field.
Thus the vector potential is fully determined by the external field H = H ẑ and we
can select A = H x ŷ. Then the equation (3.80) separates for variables x, y and z and
for the last two it becomes the Schrodinger equation for a free particle, which has the
plane-wave solutions. Thus we look for the overall solution in the form

ψ̃ = f (x)eiky y eikz z . (3.81)

Inserting this into Eq. (3.80) we get


" #
2 ′′ 2 2 4ie
2 4e2 2 2
−ξ f − ξ (ikz ) + (iky ) − H x(iky) − 2 H x f = f
h̄c h̄ c2
or  
2eH 2
−ξ 2 f ′′ + ξ 2 ky − x f = (1 − kz2 ξ 2 )f .
h̄c
Introducing
h̄c
x0 = ky
2eH
we arrive to the equation
 2
2eH ξ
−ξ 2 f ′′ + (x − x0 )2 f = (1 − kz2 ξ 2 )f . (3.82)
h̄c
We are looking for the solution of this equation which goes to zero when x → ±∞.
But Eq. (3.82) coincides with the equation of the harmonic oscillator
!
h̄2 d 2 1 2 2
− + mω x ψ = Eψ
2m dx 2 2

with
4|e|H ξ 2
ω= ,
h̄2 c
for which the solution satisfying such boundary conditions exists only when
 
1
E = h̄ω n +
2
with integer n. Applying this to Eq. (3.82) we find
 
4|e|H ξ 2 1
1 − kz2 ξ 2 = h̄ n +
h̄2 c 2
or
h̄c 1/ξ 2 − kz2
H = .
4|e| n + 1/2
3.8. UPPER CRITICAL FIELD HC2 75

Normal
Vortex

H
Hc2
Hc

H c1
Meissner

T Tc

Figure 3.8: Phase diagram of a type II superconductor.

The maximum possible value of the field obviously corresponds to kz = 0 and n = 0.


We thus obtain
h̄c 1 80 T
Hc2 = 2
= 2
∝1− . (3.83)
2|e| ξ 2πξ Tc

Comparing with Eq. (3.31) we find



Hc2 = Hc 2κ . (3.84)

For κ > 1/ 2 we have Hc2 > Hc . The phase diagram of type II superconductor is
shown in Fig. 3.8
In a type I superconductor Hc2 < Hc , but it also has a physical meaning. In this
case the superconducting transition at Hc is of the first order. As a result, normal state
can be supercooled and can exist as a metastable state at H < Hc . The reason behind
this is that owing to positive energy of NS interface, a superconducting seed of a finite
size should be formed to be stable. However, at the field Hc2 even infinitesimally small
superconducting region becomes energetically favourable. Thus the field Hc2 put the
absolute limit for supercooling of the normal phase.
In this section we have considered formation of the superconducting seed in an
infinite system. Close to the boundary (with vacuum or dielectric), however, stable
superconducting region can exist up to the field

Hc3 = 2.4κHc = 1.7Hc2 . (3.85)

Since Hc3 > Hc2 it is Hc3 which normally puts the limit on supercooling of the normal
phase.
76 CHAPTER 3. GINZBURG-LANDAU THEORY

3.9 Fluctuations. Applicability of the GL theory


The GL theory is a mean-field theory: In a homogeneous system ψ does not depend
on space coordinates or time, ψ = ψGL . Actually at non-zero temperature the system
will fluctuate around this mean value: ψ(r, t) = ψGL + δψ(r, t). Consider a particular
instantaneous realization of the fluctuation δψ(r). The fluctuation changes the energy
of the system by

δFGL [δψ] = FGL [ψGL + δψ] − FGL [ψGL ] . (3.86)

The probability to encounter such a fluctuation is


 
1 δFGL [δψ]
P [δψ] = exp − , (3.87)
Z kB T
where the normalization factor Z is called partition function
 
X δFGL [δψs ]
Z= exp − . (3.88)
s
kB T

Here the sum is taken over all possible realizations of fluctuations labeled by the in-
dex s.
The average value of any quantity a[ψ] can be calculated as
X
ha[ψ]i = a[ψGL + δψs ]P [δψs ] . (3.89)
s

In particular, the free energy turns out to be

F = −kB T ln Z . (3.90)

One may also consider deviation of ψ from the equilibrium value as creation of
some (bosonic) quasiparticles or excitations, which are called collective modes of the
order parameter. In a homogeneous system we can perform Fourier transform of the
order parameter (as before, we perform calculations for a unit volume V = 1)
X
δψ(r) = eikr δψk (3.91)
k

and present δFGL as X


δFGL [δψ(r)] = εk δψk∗ δψk .
k
Here εk is a contribution to the energy of the new excitations which we may call the
potential energy. To find the energy spectrum of the collective modes one would need
to consider time dependence and thus corresponding kinetic energy. We, however, will
only discuss the “rest mass” of the excitations, i.e. potential energy when k → 0.
3.9. FLUCTUATIONS. APPLICABILITY OF THE GL THEORY 77

3.9.1 Uncharged superfluid, T > Tc


For simplicity we first look at the case of uncharged superfluid with e = 0. This might
be a hypothetical 3 He superfluid with the s-wave pairing. (In actual superfluid 3 He
pairing is of the p-wave type.) We will also use the simplest possible approximation,
the so-called Gaussian approximation, where in Eq. (3.86) we leave only the second-
order terms with respect to fluctuations. (Note that the first-order terms disappear since
we consider fluctuations around the equilibrium value.) We start with the case T > Tc .
Here ψGL = 0 and ψ = δψ. We obtain
Z  
3 ∗ 1 ∗
δFGL [δψ(r)] = d r a(δψ(r)) δψ(r) + (−i h̄∇δψ(r)) (−i h̄∇δψ(r))
4m
Z  
X
3 −ikr ik′ r ∗ 1 ∗ ′
= d re e aδψk δψk′ + (−i h̄ ik δψk ) (−i h̄ ik δψk′ )
4m
kk′
!
X
∗ h̄2 k 2 ∗
= aδψk δψk + δψk δψk
4m
k
!
X h̄2 k 2 h i
= a+ (Re δψk )2 + (Im δψk )2 . (3.92)
4m
k
R ′
We have used relation d 3 r e−ikr eik r = δkk′ . Since δψ(r) is complex, Re δψk and
Im δψk are independent. Thus, we have found two collective modes with potential
energy
h̄2 k 2
εk = a + .
4m
Since a > 0 at T > Tc these modes have a gap. Or, in the language of field theory,
they are massive.
With the help of Eq. (3.92) one can calculate the partition function and then the
free energy and other thermodynamic quantities, like heat capacity. It turns out that the
heat capacity due to fluctuations diverges at T → Tc + 0 as

Cfluct ∝ (T − Tc )−1/2 .

3.9.2 Uncharged superfluid, T < Tc


We may expect that the situation changes in the broken-symmetry phase, since now in
the Mexican-hat potential, Fig. 3.1, the uniform change of the phase does not change
the energy of the system. We may select undisturbed order parameter to be real ψ =
ψGL and with fluctuations included it becomes

ψ = (ψGL + δψ)eiφ ,
78 CHAPTER 3. GINZBURG-LANDAU THEORY

where the fluctuations δψ and φ are real. We have


Z h b
FGL [ψ] = d 3 r a(ψGL + δψ)2 + (ψGL + δψ)4
2
1 ∗
+ − i h̄(∇δψ)eiφ + (−i h̄)i(ψGL + δψ)(∇φ)eiφ
4m
i
− i h̄(∇δψ)eiφ + (−i h̄)i(ψGL + δψ)(∇φ)eiφ (3.93)

and thus up to the second order in fluctuations


Z h b 2 1 n
δFGL = d 3 r a(δψ)2 + 6ψGL (δψ)2 + i h̄(∇δψ)∗ e−iφ (−i h̄)(∇δψ)eiφ
2 4m
oi
+ h̄ψGL (∇φ)∗ e−iφ h̄ψGL (∇φ)eiφ
Z "  #
3 2 h̄2 2 |a| 2
= d r 2|a|(δψ) + |∇δψ| + |∇φ| . (3.94)
4m b

2 = −a/b = |a|/b. Performing Fourier transform


Here we used ψGL
X X
δψ(r) = eikr δψk , φ(r) = eikr φk
k k

we obtain
( ! )
X h̄2 k 2 |a| h̄2 k 2 ∗
δFGL = 2|a| + δψk∗ δψk + φ φk . (3.95)
4m b 4m k
k

Since δψ(r) and φ(r) are real, the real and imaginary parts of δψk and φk are not
independent:
δψ−k = δψk∗ , φ−k = φk∗ . (3.96)

Thus equation (3.95) describes one massive amplitude mode with the energy

h̄2 k 2
εka = 2|a| + (3.97)
4m
and one massless phase mode with the energy

ph |a| h̄2 k 2
εk = . (3.98)
b 4m
The appearance of gapless (massless) Goldstone modes is a characteristic feature of
systems with broken continuous symmetries.
3.9. FLUCTUATIONS. APPLICABILITY OF THE GL THEORY 79

3.9.3 Ginzburg number


Since the GL theory neglects fluctuations, it is applicable only when the fluctuations
are small compared to the mean-field value of the order parameter. For the amplitude
we have

2 2
δψ ≪ ψGL (3.99)

and for the phase




φ2 ≪ 1 . (3.100)

Let us estimate the magnitude of the amplitude fluctuations. Using Eq. (3.89) we
write R a ∗

2 X′ d 2 δψk δψk∗ δψk e−(εk/ kB T )δψk δψk
δψ = R a ∗ . (3.101)
k
d 2 δψk e−(εk/ kB T )δψk δψk

The prime at the sum means that due to condition (3.96) we have to sum only over a
hemisphere of k values. Since the expressions under the integrals do not depend on the
argument of δψk we perform integration over the argument

Z Z∞
d 2 δψk → 2π|δψk | d|δψk |
0

and use integral expressions

Z∞ Z∞
−ax 2 1 2 1
xe dx = , x 3 e−ax dx =
2a 2a 2
0 0

to find
2π/ξ
Z

X′ 1/[2(εa /kB T )2 ] kB T X 1 kB T 4πk 2 dk 1
2 k
δψ = = =
1/[2εka /kB T ] 2 a
εk 2 (2π)3 |a| + h̄2 k 2 /4m
k k 0
2π/ξ
Z
kB T m k 2 dk kB T h i
= −2
= 1 − (2π)−1 arctan 2π . (3.102)
π 2 h̄2 k2+ξ 2π|a|ξ 3
0

Here we converted sum over k to the integral, took into account that the shortest wave-
length of fluctuations is ∼ ξ and used expression (3.37) for ξ .


For applicability of the GL theory one requites δψ 2 ≪ ψGL 2 , i.e.

1 k B Tc a2 Hc2
≪ =
2π ξ 3 b 8π
80 CHAPTER 3. GINZBURG-LANDAU THEORY

or  
k B Tc T 2
 −3/2 ≪ Hc (0)2 1 − .
T Tc
ξ03 1 − Tc
Thus
T
1− ≫ Gi , (3.103)
Tc
where the Ginzburg number
" #2 " #2
k B Tc k B Tc (kB Tc )4
Gi = ∼ ∼ . (3.104)
Hc (0)2 ξ03 N(0)120 (h̄vF /10 )3 EF4

Here we used Eqs. (2.59) and (1.20). In typical superconductors kB Tc ∼ 10−3 EF and
the Ginzburg number is very small Gi ∼ 10−12. Thus the applicability range of the GL
theory starts from an almost immediate vicinity of the transition.

3.10 The Anderson-Higgs mechanism


We return to discussion of the GL model with electric charge. Fluctuations in the
charged system are evidently coupled to the electromagnetic field. We will consider the
external field to be absent, so that non-zero value of the vector potential A and magnetic
field h = curl A are due to fluctuations. We continue to use Gaussian approximation.
If T > Tc and ψ = δψ one can see from the functional (3.34) that in the second order
there is no terms that couple the field and the order parameter, so their fluctuations in
the Gaussian approximation are independent.
At T < Tc the situation is more interesting. Similar to Eqs. (3.93) and (3.94) we
have
Z h b
FGL [ψ] = d 3 r a(ψGL + δψ)2 + (ψGL + δψ)4
2
1 2e 2
+ − i h̄(∇δψ)e + (−i h̄)i(ψGL + δψ)(∇φ)eiφ − A(ψGL + δψ)eiφ

4m c
1 i
+ | curl A|2

and
" ( ) #
2e 2
Z
3 2 h̄2 2 |a| 1 2
δFGL = d r 2|a|(δψ) + |∇δψ| + ∇φ − A + | curl A| .
4m b h̄c 8π
(3.105)
Owing to the gauge invariance we can add to the vector potential gradient of any func-
tion. We choose the transformation
 
h̄c
A → A′ + ∇ φ .
2e
3.10. THE ANDERSON-HIGGS MECHANISM 81

Note that curl A = curl A′ . We obtain


Z " #
3 2 h̄2 2 |a| e2 ′ 2 1 ′ 2
δFGL = d r 2|a|(δψ) + |∇δψ| + |A | + | curl A | .
4m b mc2 8π

The phase fluctuations have been absorbed to the vector potential. We drop prime from
A and perform Fourier transform of fluctuations. Since [curl A]k = −i(k × Ak ) we get

| curl A|2 → (−ik×Ak )∗ (−ik×Ak ) = (k×A∗k )·(k×Ak ) = k 2 |Ak |2 −(kA∗k )(kAk ) .

Here we used vector identity (a × b) · (c × d) = (ac)(bd) − (bc)(ad). Thus we get


( ! ! )
X h̄2 k 2 2 |a| e2 k2 2 1 ∗
δFGL = 2|a| + |δψk | + + |Ak | − (kAk )(kAk ) .
4m b mc2 8π 8π
k

We split the vector potential in parts parallel to k̂ = k/k and perpendicular to it


k k
Ak = Ak + A⊥
k, Ak = k̂(k̂Ak )
k k
so that |Ak |2 = |Ak |2 + |A⊥ 2 ∗ 2 2
k | and (kAk )(kAk ) = k |Ak | . Finally we obtain
( ! ! )
2ψ 2 2ψ 2
X h̄2 k 2 e k e k 2
δFGL = 2|a| + |δψk |2 + GL
|Ak |2 + GL
+ |A⊥
k|
2
.
4m mc2 mc2 8π
k
(3.106)
We see that above the transition, when ψGL = 0, the electromagnetic field has only
transverse degrees of freedom, which are massless (photons). In the superconducting
state the longitudinal component appears and all components become massive. Com-
pare this to uncharged system where the phase mode is Goldstone, i.e. massless. This
mechanism in which the Goldstone mode acquires a mass is known as the Anderson-
Higgs mechanism. In this context the massive amplitude mode is called the Higgs
boson.
Similar mechanism is thought to be working in the high-energy physics, where the
symmetry breaking at the electroweak transition leads to appearance of the Higgs field
and the gauge W and Z bosons become massive.
82 CHAPTER 3. GINZBURG-LANDAU THEORY

Problems
Problem 3.1. Find expression for the coefficient γ in the Ginzburg-Landau free energy
(3.8) from the microscopic theory in the clean limit.

Problem 3.2. Find the Ginzburg-Landau superconducting density in the dirty limit and
compare it with ns in the clean limit.

Problem 3.3. Calculate the critical current for a thin film with thickness d ≪ ξ(T ).

Problem 3.4. A thin superconducting film with a thickness d ≪ λL and d ≪ ξ(T ) is


deposited on a dielectric filament (cylinder) of a radius R ≫ d. The filament is placed
into a longitudinal magnetic field H at a temperature T > Tc and then cooled down
below Tc . Find the dependence of Tc on H . (The Little and Parks effect.)

Problem 3.5. A thin superconducting slab with thickness d ≪ ξ(T ) is placed in the
magnetic field H oriented parallel to the slab. Find the dependence of the magnitude
of the order parameter in the slab on H for cases d ≫ λL and d ≪ λL .

Problem 3.6. A thin superconducting slab with thickness d ≪ ξ(T ) is driven form
the superconducting to normal state by the magnetic field oriented parallel to the slab.
Find the critical thickness dc so that at d > dc the transition is of the first order and at
d < dc it is of the second order.

Problem 3.7. The film with thickness d ≫ ξ has the same upper critical field as a bulk
superconductor. Find the upper critical field for a film with a thickness d ≪ ξ placed
in a magnetic field tilted by an angle 2 from the normal of the film.

Problem 3.8. Calculate the energy of the normal-superconductor boundary in an ex-


treme type I superconductor in the limit λL → 0.

Problem 3.9. Calculate electric current around a straight Abrikosov vortex.

Problem 3.10. Abrikosov vortex is pinned on a void with diameter d so that d & ξ
and d ≪ λL . Estimate the critical current to release the vortex from the pinning site.
In the calculations ignore the unpinned part of the vortex.

Problem 3.11. Calculate the mean square of phase fluctuations hφ 2 i in uncharged


superfluid. Show that the conditions hφ 2 i ≪ 1 results in the same Ginzburg criterion
(3.103) as for the amplitude fluctuations.

Problem 3.12. Calculate fluctuation contribution Cfluct to the heat capacity. Show that
the condition Cfluct ≪ 1C, where the 1C is the heat capacity jump at the transition,
results in the Ginzburg criterion (3.103).

You might also like