0% found this document useful (0 votes)
105 views79 pages

Final PHYSICS PROJECT

The document is a Physics project submitted by V Keerthana from St. Francis School on the topic of Superconductivity, guided by Ms. Shruti Chatrad. It includes a certificate of completion, acknowledgement, an index, and detailed sections on superconductivity principles, the Meissner-Ochsenfeld effect, thermodynamic properties, and the BCS theory. The project aims to explore the fascinating properties of superconductivity and its applications, supported by various figures and references.

Uploaded by

vkeerthana2007
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
105 views79 pages

Final PHYSICS PROJECT

The document is a Physics project submitted by V Keerthana from St. Francis School on the topic of Superconductivity, guided by Ms. Shruti Chatrad. It includes a certificate of completion, acknowledgement, an index, and detailed sections on superconductivity principles, the Meissner-Ochsenfeld effect, thermodynamic properties, and the BCS theory. The project aims to explore the fascinating properties of superconductivity and its applications, supported by various figures and references.

Uploaded by

vkeerthana2007
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 79

St.

FRANCIS SCHOOL
Koramangala, Bengaluru -560 034
(Affiliated to CISCE Board, New Delhi)

Board Examination Project

Name: …………………………..

Class: …………………………..

U.I.D. Number: …………………………..

Subject: …………………………..
1
2
St. FRANCIS SCHOOL
Koramangala, Bengaluru -560 034
(Affiliated to CISCE Board, New Delhi)

Physics Project
Topic:

Submitted to:
Ms. Shruti Chatrad
Submitted by:
V Keerthana
Grade XII

3
4
Certificate of Completion

This is to certify that V Keerthana ,a student of Grade XII (PCMB), has


successfully completed the Physics Project titled:
Topic: Superconductivity
under the guidance of Ms. Shruti Chatrad, Faculty of Physics, during the
Academic Session 2024-2025.

Internal Examiner

Name: ………………………………………

Signature: ………………………………………

Date: ………………………………………

External Examiner

Name: ………………………………………

Signature: ………………………………………
\

Date: ………………………………………

5
6
Acknowledgement

The project titled ‘Superconductivity’ is done as part of the internal


assessment for the Physics at ISC level. As the curriculum demands I have
made it as original as possible avoiding plagiarism as far as possible. I have
acknowledged in the bibliography section all the resources that I used to
complete the project

I place on record my sincere thanks to Rev. Bro. Antony Vayalil, the


principal, for his constant guidance and encouragement. I am deeply
indebted to Ms. Shruti Chatrad , my Physics teacher, for her valuable
assistance and guidance throughout the project.

Thanks to all my friends who helped me in different ways to complete the


project.

V Keerthana

Grade XII, PMCB

7
Sl. No Index Page
Number
1. Introduction 10
2. Meissner-Ochsenfeld Effect And London Equation 14
3. Thermodynamic Properties Of Semiconductors 22
4. Basic Concepts And Results Of The BCS Theory 42
5. Quantisation Of Magnetic Flux 66
6 Josephson Effects 74
7. Conclusion 77

8
8. Bibliography 79

9
Fig. 1: The discovery of superconductivity by Kamerlingh Onnes

Fig. 2: The low-temperature resistivity of copper, tin and YBa2Cu3O7

Fig. 3: Induction of a persistent current in a superconducting ring.

10
1.INTRODUCTION
I want to give an introduction into the physical principles of superconductivity
and its fascinating properties. More detailed accounts can be found in the
excellent text books by W. Buckel [1] and by D.R. Tilley and J. Tilley [2].
Superconductivity was discovered [3] in 1911 by the Dutch physicist
H. Kamerlingh Onnes, only three years after he had succeeded in liquefying
helium. During his investigations on the conductivity of metals at low
temperature he found that the resistance of a mercury sample dropped to an
unmeasurably small value just at the boiling temperature of liquid helium. The
original measurement is shown in Fig. 1. Kamerlingh Onnes called this totally
unexpected phenomenon ‘superconductivity’ and this name has been retained
since. The temperature at which the transition took place was called the critical
temperature Tc. Superconductivity is observed in a large variety of materials but,
remarkably, not in some of the best normal conductors like copper, silver and
gold, except at very high pressures. This is illustrated in Fig. 2 where the
resistivity of copper, tin and the ‘high-temperature’ superconductor YBa 2Cu3O7
is sketched as a function of temperature. Table 1 lists some important
superconductors together with their critical temperatures at vanishing magnetic
field.

Table 1: The critical temperature of some common materials at


vanishing magnetic field.
materia Ti Al Hg Sn Pb N NbT Nb3S
l b i n
Tc [K] 0. 1.1 4.1 3.7 7. 9. 9.2 18
4 4 5 2 9 2

A conventional resistance measurement is far too insensitive to establish


infinite conductivity, a much better method consists in inducing a current in a
ring and determining the decay rate of the produced magnetic field. A schematic
experimental setup is shown in Fig. 3. A bar magnet is inserted in the still
normal-conducting ring and removed after cooldown below Tc. The induced
current should decay exponentially
I(t) = I(0)exp(−t/τ)

11
12
with the time constant given by the ratio of inductivity and resistance, τ = L/R,
which for a normal metal ring is in the order of 100 µs. In superconducting
rings, however, time constants of up to 105 years have been observed [4] so the
resistance must be at least 15 orders of magnitude below that of copper and is
indeed indistinguishable from zero. An important practical application of this
method is the operation of solenoid coils for magnetic resonance imaging in the
short-circuit mode which exhibit an extremely slow decay of the field of
typically 3 · 10−9 per hour [5].

There is an intimate relation between superconductivity and magnetic


fields. W. Meissner and R. Ochsenfeld [6] discovered in 1933 that a
superconducting element like lead completely expelled a weak magnetic field
from its interior when cooled below Tc while in strong fields superconductivity
broke down and the material went to the normal state. The spontaneous
exclusion of magnetic fields upon crossing Tc could not be explained in terms of
the Maxwell equations and indeed turned out to be a non- classical phenomenon.
Two years later, H. and F. London [7] proposed an equation which offered a
phenomenological explanation of the Meissner-Ochsenfeld effect but the
justification of the London equation remained obscure until the advent of the
Bardeen, Cooper and Schrieffer theory [8] of superconductivity in 1957. The
BCS theory revolutionized our understanding of this fascinating phenomenon. It
is based on the assumption that the supercurrent is not carried by single electrons
but rather by pairs of electrons of opposite momenta and spins, the so-called
Cooper pairs. All pairs occupy a single quantum state, the BCS ground state,
whose energy is separated from the single-electron states by an energy gap
which in turn can be related to the critical temperature. The BCS theory has
turned out to be of enormous predictive power and many of its predictions and
implications like the temperature dependence of the energy gap and its relation
to the critical temperature, the quantisation of magnetic flux and the existence of
quantum interference phenomena have been confirmed by experiment and, in
many cases, even found practical application.

13
Fig. 4: A lead cylinder in a magnetic field. Two possible ways to reach the superconducting
final state with H > 0 are sketched.

14
Fig. 5: The phase diagram in a (T,H) plane.

2. MEISSNER-OCHSENFELD EFFECT AND LONDON


EQUATION
We consider a cylinder with perfect conductivity and raise a magnetic field
from zero to a finite value H. A surface current is induced whose magnetic field,
according to Lenz’s rule, is opposed to the applied field and cancels it in the
interior. Since the resistance is zero the current will continue to flow with
constant strength as long as the external field is kept constant and consequently
the bulk of the cylinder will stay field-free. This is exactly what happens if we
expose a lead cylinder in the superconducting state (T < Tc) to an increasing
field, see the path (a) → (c) in Fig. 4. So below Tc lead acts as a perfect
diamagnetic material. There is, however, another path leading to the point (c).
We start with a lead cylinder in the normal state (T > Tc) and expose it to a field
which is increased from zero to H. Eddy currents are induced in this case as well
but they decay rapidly and after a few hundred microseconds the field lines will
fully penetrate the material (state (b) in Fig. 4). Now the cylinder is cooled
down. At the very instant the temperature drops below Tc , a surface current is
spontaneously created and the magnetic field is expelled from the interior of the
cylinder. This surprising observation is called the Meissner-Ochsenfeld effect
after its discoverers; it cannot be explained by the law of induction because the
magnetic field is kept constant.

In a (T,H) plane, the superconducting phase is separated from the normal


phase by the curve Hc(T) as sketched in Fig. 5. Also indicated are the two ways on
which one can reach the point (c). It is instructive
to compare this with the response of a ‘normal’ metal of perfect conductivity.
The field increase along the path (a) → (c) would yield the same result as for the
superconductor, however the cooldown along the path (b) → (c) would have no
effect at all. So superconductivity means definitely more than just vanishing
resistance.

I have already used the terms ‘superconducting phase’ and ‘normal phase’
to characterize the two states of lead. These are indeed phases in the

15
thermodynamical sense, comparable to the different phases of H 2O which is in
the solid, liquid or gaseous state depending on the values of the parameters

temperature and pressure. Here the relevant parameters are temperature and
magnetic field (for some materials also pressure). If the point (T,H) lies below
the curve Hc(T) the material is superconducting and expels the magnetic field,
irrespective of by which path the point was reached. If (T,H) is above the curve
the material is normal-conducting.

16
The first successful explanation of the Meissner-Ochsenfeld effect was
achieved in 1935 by Heinz and Fritz London. They assumed that the
supercurrent is carried by a fraction of the conduction electrons in the metal.
The ‘super-electrons’ experience no friction, so their equation of motion in an
electric field is

.
This leads to an accelerated motion. The supercurrent density is

where ns is the density of the super-electrons. This immediately yields the


equation

. (1)
Now one uses the Maxwell equation

and takes the curl (rotation) of (1) to obtain

.
Since the time derivative vanishes the quantity in the brackets must be a
constant. Up to this point the derivation is fully compatible with classical
electromagnetism, applied to the frictionless acceleration of electrons. An
example might be the motion of electrons in the vacuum of a television tube or
in a circular accelerator. The essential new assumption H. and F. London made
is that the bracket is not an arbitrary constant but is identical to zero. Then one
obtains the important London equation

. (2)

17
It should be noted that this assumption cannot be justified within classical
physics, even worse, in general it is wrong. For instance the current density in a
normal metal will vanish when no electric field is applied, and whether a static
magnetic field penetrates the metal is of no importance. In a superconductor of
type I, on the other hand, the situation is such that Eq. (2) applies. Combining
the fourth Maxwell equation (for time-independent fields)

18
and the London equation and making use of the relation

×~ ( ×~ B~) = −
2B~

(this is valid since ·~B~ = 0) we get the following equation for the magnetic
field in a superconductor

. (3)
It is important to note that this equation is not valid in a normal conductor. In
order to grasp the significance of Eq. (3) we consider a simple geometry, namely
the boundary between a superconducting half space and vacuum, see Fig. 6a.
Then, for a magnetic field parallel to the surface, Eq. (3) becomes

with the solution


By(x) = B0 exp(−x/λL) .

Here we have introduced a very important superconductor parameter, the


London penetration depth

19
Fig. 6: (a) Exponential attenuation of a magnetic field in a superconducting half plane. (b)
Shielding current in a superconducting cylinder induced by a field parallel to the axis. (c) A
current-carrying wire made from a type I superconductor.

Fig. 7: Temperature dependence of the London penetration depth.

20
. (4)
So the magnetic field does not stop abruptly at the superconductor surface but
penetrates into the material with exponential attenuation. For typical material
parameters the penetration depth is quite small, 20 – 50 nm. In the bulk of a
thick superconductor there can be no magnetic field, which is just the
MeissnerOchsenfeld effect. Here it is appropriate to remark that in the BCS
theory not single electrons but pairs of electrons are the carriers of the
supercurrent. Their mass is mc = 2me, their charge −2e, their density nc = ns/2.
Obviously the penetration depth remains unchanged when going from single
electrons to Cooper pairs.
We have now convinced ourselves that the superconductor can tolerate a
magnetic field only in a thin surface layer (this is the case for type I
superconductors). An immediate consequence is that current flow is restricted to
the same thin layer. Currents in the interior are forbidden as they would generate
magnetic fields in the bulk. The magnetic field and the current which are caused
by an external field parallel to the axis of a lead cylinder are plotted in Fig. 6b.
Another interesting situation occurs if we pass a current through a lead wire
(Fig. 6c). It flows only in a very thin surface sheet of about 20 nm thickness, so
the overall current in the wire is small. This is a first indication that type I
superconductors are not suitable for winding superconducting magnet coils.

The penetration depth has a temperature dependence which can be


calculated in the BCS theory. When approaching the critical temperature, the
density of the supercurrent carriers goes to zero, so λL must become infinite:
λL → ∞ for T → Tc .

This is shown in Fig. 7. An infinite penetration depth means no attenuation of a


magnetic field which is just what one observes in a normal conductor.

21
Fig. 8: (a) Free energy of aluminium in the normal and superconducting state as a function of
T (after N.E. Phillips)..

3. THERMODYNAMIC PROPERTIES OF SUPERCONDUCTORS


3.1 The superconducting phase
A material like lead goes from the normal into the superconducting state when
it is cooled below Tc and when the magnetic field is less than Hc(T). It has been

22
mentioned already that this is a phase transition comparable to the transition
from water to ice at 0◦ C and normal pressure. Phase transitions take place when
the new state is energetically favoured. The relevant thermodynamic energy is
here the Gibbs free energy (see Appendix A)

G = U − T S − µ0M~ · H~ (5)
where U is the internal energy, S the entropy and M the magnetisation of the
superconductor (the magnetic moment per unit volume). A measurent of the free
energy of aluminium is shown in Fig. 8a. Below Tc the superconducting state has
a lower free energy than the normal state and thus the transition normal →
superconducting is associated with a gain in energy. The entropy of the
superconducting state is lower because there is a higher degree of order in this
state. From the point of view of the BCS theory this is quite understandable
since the conduction electrons are paired and collect themselves in a single
quantum state. Numerically the entropy difference is small, though, about 1
milli-Joule per mole and Kelvin, from which one can deduce that only a small
fraction of the valence electrons of aluminium is condensed into Cooper pairs. It
should be noted, that also normal conduction is carried by just a small fraction of
the valence electrons, see sect. 4.1.

3.2 Energy balance in a magnetic field


We have argued that a lead cylinder becomes superconductive for T < Tc
because the free energy is reduced that way:
Gsup < Gnorm for T <
Tc .
What happens if we apply a magnetic field? A normal-conducting metal
cylinder is penetrated by the field so its free energy does not change: Gnorm(H) =
Gnorm(0). In contrast to this, a superconducting cylinder is strongly affected by
the field. It sets up shielding currents which generate a magnetic moment m~
antiparallel to

23
the applied field. The magnetic moment has a positive potential energy in the
magnetic field

Epot = −µ0 m~ · H~ = +µ0 |m~ ||H~ | . (6)

24
In the following it is useful to introduce the magnetisation M as the magnetic
moment per unit volume. The magnetisation of a superconductor inside a
current-carrying coil resembles that of an iron core. The ‘magnetising’ field H is
generated by the coil current only and is unaffected by the presence of a
magnetic

material while the magnetic flux density 1 B is given by the superposition of H


and the superconductor magnetisation M:

B~ = µ0(H~ + M~ ) . (7)

In the following I will call both H and B magnetic fields. For a type I
superconductor we have

M~ (H~ ) = −H~ and B~ = 0 (8)

as long as H < Hc. The potential energy per unit volume is obtained by
integration

. (9)
This corresponds to the increase in the Gibbs free energy that is caused by the
magnetic field, see Fig. 8b.
. (10)
Here and in the following G denotes the Gibbs free energy per unit volume.
The critical field is achieved when the free energy in the superconducting state
just equals the free energy in the normal state

1 There is often a confusion whether the H or the B field should be used. Unfortunately, much of the superconductivity
literature is based on the obsolete CGS system of units where the distinction between B and H is not very clear and the two
fields have the same dimension although their units were given different names: Gauss and Oerstedt.

25
. (11)
Since the energy density stored in a magnetic field is (µ0/2)H2, an alternative
interpretation of Eq. (11) is the following: in order to go from the normal to the
superconducting state the material has to push out the magnetic energy, and the
largest amount it can push out is the difference between the two free energies at

26
vanishing field. For H > Hc the normal phase has a lower energy, so
superconductivity breaks down.

27
Fig. 9: Magnetisation of type I and type II superconductors as a function of the magnetic
field.

28
3.3 Type II superconductors
For practical application in magnets it would be rather unfortunate if only type I
superconductors existed which permit no magnetic field and no current in the
bulk material. Alloys and the element niobium are so-called type II
superconductors. Their magnetisation curves exhibit a more complicated
dependence on magnetic field (Fig. 9). Type II conductors are characterized by
two critical fields, Hc1 and Hc2 , which are both temperature dependent. For fields
0 < H < Hc1 the substance is in the Meissner phase with complete exclusion of
the field from the interior. In the range Hc1 < H < Hc2 the substance enters the
mixed phase, often also called Shubnikov phase: part of the magnetic flux
penetrates the bulk of the sample. Above Hc2 , finally, the material is normal-
conducting. The area under the curve M = M(H) is the same as for a type I
conductor as it corresponds to the free-energy difference between the normal
and the superconducting state and is given by (µ0/2)Hc2.

29
Fig. 10: The measured magnetisation curves [14] of lead-indium alloys of various
composition, plotted against B = µ0H.

Fig. 11: Flux tubes in a type II superconductor.

30
It is instructive to compare measured data on pure lead (type I) and lead-
indium alloys (type II) of various composition. Figure 10 shows that the upper
critical field rises with increasing indium content; for Pb-In(20.4%) it is about
eight times larger than the critical field of pure lead. Under the assumption that
the free-energy difference is the same for the various lead-indium alloys, the
areas under the three curves A, B, C in Fig. 10 should be identical as the diagram
clearly confirms.

A remarkable feature, which will be addressed in more detail later, is the


observation that the magnetic flux does not penetrate the type II conductor with
uniform density. Rather it is concentrated in flux tubes as sketched in Fig. 11.
Each tube is surrounded with a super-vortex current. The material in between the
tubes is field- and current-free.

31
Fig. 12: Top: Magnetic moment of a type I and a type II sc cylinder in a field Hc1 < H < Hc.
Bottom: The Gibbs free energies of both cylinders as a function of field.

Fig. 13: (a) The phase diagram of a type II superconductor. (b) The upper critical field Bc2 =
µ0Hc2 of several high-field alloys as a function of temperature.

32
The fact that alloys stay superconductive up to much higher fields is easy
to understand: magnetic flux is allowed to penetrate the sample and therefore
less magnetic field energy has to be driven out. Figure 12 shows that a type II
superconducting cylinder in the mixed phase has a smaller magnetic moment
than a type I cylinder. This implies that the curve Gsup(H) reaches the level
Gsup(H) = Gnorm at a field Hc2 > Hc .
In a (T,H) plane the three phases of a type II superconductor are separated
by the curves Hc1(T) and Hc2(T) which meet at T = Tc, see Fig. 13a. The upper
critical field can assume very large values which make these substances
extremely interesting for magnet coils (Fig. 13b).

33
Fig. 14: Attenuation of field (a) in a thick slab and (b) in thin sheet. (c) Subdivision of a thick
slab into alternating layers of normal and superconducting slices.

Fig. 15: The exponential drop of the magnetic field and the rise of the Cooper-pair density at a
boundary between a normal and a superconductor.
3.4 When is a superconductor of type I or type II?

34
3.41 Thin sheets of type I superconductors
Let us first stick to type I conductors and compare the magnetic properties of a
very thin sheet (thickness d < λL) to those of a thick slab. The thick slab has a
vanishing B field in the bulk (Fig. 14a) while in the thin sheet (Fig. 14b) the B
field does not drop to zero at the centre. Consequently less energy needs to be
expelled which implies that the critical field of a very thin sheet is much larger
than the Bc of a thick slab. From this point of view it might appear energetically
favourable for a thick slab to subdivide itself into an alternating sequence of thin
normal and superconducting slices as indicated in Fig. 14c. The magnetic energy
is indeed lowered that way but there is another energy to be taken into
consideration, namely the energy required to create the normal-superconductor
interfaces. A subdivision is only sensible if the interface energy is less than the
magnetic energy.

3.42 Coherence length


At a normal-superconductor boundary the density of the supercurrent carriers
(the Cooper pairs) does not jump abruptly from zero to its value in the bulk but
rises smoothly over a finite length ξ, the coherence length, see Fig. 15.

The relative size of the London penetration depth and the coherence length
decides whether a material is a type I or a type II superconductor. To study this
in a semi-quantitative way, we first define the thermodynamic critical field by
the energy relation
. (12)

For type I this coincides with the known Hc , see Eq. (11), while for type II
conductors Hc lies between
Hc1 and Hc2. The difference between the two free energies, Gnorm − Gsup(0), can
be intepreted as the Cooper-pair condensation energy.

35
For a conductor of unit area, exposed to a field H = Hc parallel to the
surface, the energy balance is as follows:

36
(a) The magnetic field penetrates a depth λL of the sample which corresponds
to an energy gain since magnetic energy must not be driven out of this layer:
. (13)
(b) On the other hand, the fact that the Cooper-pair density does not assume its
full value right at the surface but rises smoothly over a length ξ implies a loss of
condensation energy
. (14)

37
Obviously there is a net gain if λL > ξ. So a subdivision of the superconductor
into an alternating sequence of thin normal and superconducting slices is
energetically favourable if the London penetration depth exceeds the coherence
length.
A more refined treatment is provided by the Ginzburg-Landau theory [9].
Here one introduces the Ginzburg-Landau parameter
κ = λ L/ ξ . (15)
The criterion for type I or II superconductivity is found to be
type I: type
II: .
In reality a type II superconductor is not subdivided into thin slices but the
field penetrates the sample in flux tubes which arrange themselves in a triangular

38
pattern. The core of a flux tube is normal. The following table lists the
penetration depths and coherence lengths of some important superconducting
elements. Niobium is a type II conductor but close to the border to type I, while
indium, lead and tin are clearly type I conductors.
The coherence length ξ is proportional to the mean free path ` of the
conduction electrons in the metal. This quantity can be large for a very pure
crystal but is strongly reduced by lattice defects and impurity atoms. In alloys
the mean free path is generally much shorter than in pure metals so alloys are
always type II conductors. In the Ginzburg-Landau theory the upper critical field
is given by

(16)
where Φ0 is the flux quantum (see sect. 5.2). For niobium-titanium with an
upper critical field Bc2 = 10 T at 4.2 K this formula yields ξ = 6nm. The
coherence length is larger than the typical width of a grain boundary in NbTi
which means that the supercurrent can move freely from grain to grain. In high-
Tc superconductors the coherence length is often shorter than the grain boundary
width, and then current flow from one grain to the next is strongly impeded.
There exists no simple expression for the lower critical field. In the limit
one gets
. (17)

39
Fig. 16: (a) Specific heat C(T)/T of normal and superconducting gallium as a function of T2.
(b) Experimental verification of Eq. (20).

Fig. 17: Measured heat conductivity in niobium samples with RRR = 270 and RRR = 500 as a
function of temperature [16].

40
3.5 Heat capacity and heat conductivity
The specific heat capacity per unit volume at low temperatures is given by the
expression

CV (T) = γ T + AT3 . (18)

The linear term in T comes from the conduction electrons, the cubic term from
lattice vibrations. The coefficients can be calculated within the free-electron-gas
model and the Debye theory of lattice specific heat (see any standard textbook
on solid state physics):

. (19)

Here kB = 1.38 · 10−23 J/K is the Boltzmann constant, EF the Fermi energy, n the
density of the free electrons, N the density of the lattice atoms and ΘD the Debye
temperature of the material. If one plots the ratio C(T)/T as a function of T 2 a
straight line is obtained as can be seen in Fig. 16a for normal conducting gallium
[15]. In the superconducting state the electronic specific heat is different because
the electrons bound in Cooper pairs no longer contribute to energy transport. In
the BCS theory one expects an exponential rise of the electronic heat capacity
with temperature
Ce,s(T) = 8.5γ Tc exp(−1.44Tc/T) (20)

The experimental data (Fig. 16a, b) are in good agreement with this prediction.
There is a resemblance to the exponential temperature dependence of the
electrical conductivity in intrinsic semiconductors and these data can be taken as
an indication that an energy gap exists also in superconductors.
The heat conductivity of niobium is of particular interest for
superconducting radio frequency cavities. Here the theoretical predictions are
rather imprecise and measurements are indispensible. The low temperature
values depend strongly on the residual resistivity ratio RRR = R(300K)/R(10K)
of the normal-conducting niobium. Figure 17 shows experimental data [16].

41
Fig. 18: The allowed states for conduction electrons in the pxpy plane and the Fermi sphere.
The occupied states are drawn as full circles, the empty states as open circles.

42
4. BASIC CONCEPTS AND RESULTS OF THE BCS THEORY
4.1 The ‘free electron gas’ in a normal metal
4.11 The Fermi sphere
In a metal like copper the positively charged ions form a regular crystal lattice.
The valence electrons (one per Cu atom) are not bound to specific ions but can
move through the crystal. In the simplest quantum theoretical model the
Coulomb attraction of the positive ions is represented by a potential well.
The Schro¨dinger equation with boundary conditions, and then the electrons are
placed on these levels paying attention to the Pauli exclusion principle: no more
than two electrons of opposite spin are allowed on each level. The electrons are
treated as independent and non-interacting particles, their mutual Coulomb
repulsion is taken into account only globally by a suitable choice of the depth of
the potential well. It is remarkable that such a simple-minded picture of a ‘free
electron gas’ in a metal can indeed reproduce the main features of electrical and
thermal conduction in metals. However, an essential prerequisite is to apply the
Fermi-Dirac statistics, based on the Pauli principle, and to avoid the classical
Boltzmann statistics which one uses for normal gases. The electron gas has indeed
rather peculiar properties. The average kinetic energy of the metal electrons is by
no means given by the classical expression

which amounts to about 0.025 eV at room temperature. Instead, the energy


levels are filled with two electrons each up to the Fermi energy EF . Since the
electron density n is very high in metals, EF assumes large values, typically 5 eV.
The average kinetic energy of an electron is 3/5EF ≈ 3 eV and thus much larger
than the average energy of a usual gas molecule. The electrons constitute a
system called a ‘highly degenerate’ Fermi gas. The Fermi energy is given by the
formula

. (21)

The quantity ~ = h/2π = 1.05 · 10−34 Js = 6.58 · 10−15 eVs is Planck’s


constant, the most important constant in quantum theory. In order to remind the

43
reader I will shortly sketch the derivation of these results. Consider a three-
dimensional region in the metal of length L = N a, where a is the distance of
neighbouring ions in the lattice and The Schro¨dinger equation
with potential V = 0
and with periodic boundary conditions ψ(x + L,y,z) = ψ(x,y,z) etc. is solved by

44
ψ(x,y,z) = L−3/2 exp(i(k1x + k2y + k3z)) (2
2)
where the components of the wave vector ~k are given by
(2
with nj = 0,±1,±2,... 3)

The electron momentum is p~ = ~~k, the energy is E = ~2~k2/(2me). It is useful


to plot the allowed quantum states of the electrons as dots in momentum space.
In Fig. 18 this is drawn for two dimensions.

45
In the ground state of the metal the energy levels are filled with two
electrons each starting from the lowest level. The highest energy level reached is
called the Fermi EF . At temperature T → 0 all states below EF are occupied, all
states above EF are empty. The highest momentum is called the ‘Fermi
momentum’ , the highest velocity is the Fermi velocity vF = pF/me
which is in the

46
order of 106 m/s. In the momentum state representation, the occupied states are
located inside the ‘Fermi
sphere’ of radius pF, the empty states are outside.

What are the consequences of the Pauli principle for electrical conduction?
Let us apply an electric field pointing into the negative x direction. In the time
δt a free electron would gain a momentum
δpx = eE0 δt . (24)

However, most of the metal electrons are unable to accept this momentum
because they do not find free states in their vicinity, only those on the right rim
of the Fermi sphere have free states accessible to them and can accept the
additional momentum. We see that the Pauli principle has a strong impact on
electrical conduction. Heat conduction is affected in the same way because the
most important carriers of thermal energy are again the electrons. An anomaly is
also observed in the heat capacity of the electron gas. It differs considerably
from that of an atomic normal gas since only the electrons in a shell of thickness
kBT near the surface of the Fermi sphere can contribute. Hence the electronic
specific heat per unit volume is roughly a fraction kBT/EF of the classical value

.
This explains the linear temperature dependence of the electronic specific heat,
see eq. (19).

The origin of Ohmic resistance


Before trying to understand the vanishing resistance of a superconductor we
have to explain first why a normal metal has a resistance. This may appear
trivial if one imagines the motion of electrons in a crystal that is densely filled
with ions. Intuitively one would expect that the electrons can travel for very
short distances

47
Fig. 19: Temperature dependence of the resistivity of OFHC (oxygen-free high conductivity)
copper and of 99.999% pure annealed copper. Plotted as a dashed line is the calculated
resistivity of copper without any impurities and lattice defects (after M.N. Wilson [17]).

only before hitting an ion and thereby loosing the momentum gained in the
electric field. Collisions are indeed responsible for a frictional force and one can
derive Ohm’s law that way. What is surprising is the fact that these collisions are
so rare. In an ideal crystal lattice there are no collisions whatsoever. This is
impossible to understand in the particle picture, one has to treat the electrons as

48
matter waves and solve the Schro¨dinger equation for a periodic potential. The
resistance is nevertheless due to collisions but the collision centres are not the
ions in the regular crystal lattice but only the imperfections of this lattice:
impurities, lattice defects and the deviations of the metal ions from their nominal
position due to thermal oscillations. The third effect dominates at room
temperature and gives rise to a resistivity that is roughly proportional to T while
impurities and lattice defects are responsible for the residual resistivity at low
temperature (T < 20 K). A typical curve ρ(T) is plotted in Fig. 19. In very pure
copper crystals the low-temperature resistivity can become extremely small. The
mean free path of the conduction electrons may be a million times larger than
the distance between neighbouring ions which illustrates very well that the ions
in their regular lattice positions do not act as scattering centres.

49
Fig. 20: A pair of electrons of opposite momenta added to the full Fermi sphere.

4.2 Cooper pairs


We consider a metal at T → 0. All states inside the Fermi sphere are filled with

electrons while all states outside are empty. In 1956 Cooper studied [18] what

would happen if two electrons were added to the

50
filled Fermi sphere with equal but opposite momenta whose magnitude

was slightly larger than the Fermi momentum pF (see Fig. 20). Assuming that a

weak attractive force existed he was able to show that the electrons form a

bound system with an energy less than twice the Fermi energy , Epair < 2EF . The

mathematics of Cooper pair formation will be outlined in Appendix B.

What could be the reason for such an attractive force? First of all one has to
realize that the Coulomb repulsion between the two electrons has a very short
range as it is shielded by the positive ions and the other electrons in the metal.
So the attractive force must not be strong if the electrons are several lattice
constants apart. Already in 1950, Fro¨hlich and, independently, Bardeen had
suggested that a dynamical lattice polarization may create a weak attractive
potential. Before going into details let us look at a familiar example of attraction
caused by the deformation of a medium: a metal ball is placed on an elastic
membrane and deforms the membrane such that a potential well is created. A
second ball will feel this potential well and will be attracted by it. So effectively,
the deformation of the elastic membrane causes an attractive force between the
two balls which would otherwise not notice each other. This visualisation of a
Cooper-pair is well known in the superconductivity community (see e.g. [1]) but
it has the disadvantage that it is a static picture.
You will find this quite cumbersome, there is a lot of ‘resistance’. Now you
discover a track made by another skier, a ‘Loipe’, and you will immediately
realize that it is much more comfortable to ski along this track than in any other
direction. The Loipe picture can be adopted for our electrons. The first electron
flies through the lattice and attracts the positive ions.

. (25)

51
Fig. 21: Dynamical deformation of the crystal lattice caused by the passage of a fast electron.
(After Ibach, L¨uth [19]).

Fig. 22: Visualization of Cooper pairs and single electrons in the crystal lattice of a
superconductor. (After Essmann and Tr¨auble [12]).

52
Obviously, the lattice deformation attracts the second electron because
there is an accumulation of positive charge. The attraction is strongest when the
second electron moves right along the track of the first one and when it is a
distance d behind it, see Fig. 21. This explains why a Cooper pair is a very
extended object, the two electrons may be several 100 to 1000 lattice constants
apart. For a simple cubic lattice, the lattice constant is the distance between
adjacent atoms.
In the example of the cross-country skiers or the electrons in the crystal
lattice, intuition suggests that the second partner should preferably have the same
momentum, although opposite momenta are not so bad either.
Quantum theory makes a unique choice: only electrons of opposite
momenta form a bound system, a Cooper pair. I don’t know of any intuitive
argument why this is so. (The quantum theoretical reason is the Pauli principle
but there exists probably no intuitive argument why electrons obey the Pauli
exclusion principle and are thus extreme individualists while other particles like
the photons in a laser or the atoms in superfluid helium do just the opposite and
behave as extreme conformists. One may get used to quantum theory but certain
mysteries and strange feelings will remain.)
The binding energy of a Cooper pair turns out to be small, 10−4−10−3 eV, so
low temperatures are needed to preserve the binding in spite of the thermal
motion. According to Heisenberg’s Uncertainty Principle a weak binding is
equivalent to a large extension of the composite system, in this case the above-
mentioned d = 100 − 1000 nm. As a consequence, the Cooper pairs in a
superconductor overlap each other. In the space occupied by a Cooper pair there
are about a million other Cooper pairs. Figure 22 gives an illustration. The
situation is totally different from other composite systems like atomic nuclei or
atoms which are tightly bound objects and well-separated from another. The
strong overlap is an important prerequisite of the BCS theory because the
Cooper pairs must change their partners frequently in order to provide a
continuous binding.

53
Fig. 23: Various Cooper pairs in momentum space.

Fig. 24: (a) Energy gap between the BCS ground state and the single-electron states. (b)
Reduction of energy gap in case of current flow.

54
4.3 Elements of the BCS theory
After Cooper had proved that two electrons added to the filled Fermi sphere
are able to form a bound system with an energy Epair < 2EF , it was immediately
realized by Bardeen, Cooper and Schrieffer ,

also the electrons inside the Fermi sphere should be able to group themselves
into pairs and thereby reduce their energy. The attractive force is provided by
lattice vibrations whose quanta are the phonons. The highest possible phonon
energy is

~ωD = kBΘD ≈ 0.01 − 0.02 eV . (26)

Therefore only a small fraction of the electrons can be paired via phonon
exchange, namely those in a shell of thickness ±~ωD around the Fermi energy.
This is sketched in Fig. 23. The inner electrons cannot participate in the pairing
because the energy transfer by the lattice is too small. For vanishing electric
field a Cooper pair is a loosely bound system of two electrons whose momenta
are of equal magnitude but opposite direction. All Cooper pairs have therefore
the same momentum P~ = 0 and occupy exactly the same quantum state.
The reason why Cooper pairs are allowed and even prefer to enter the same
quantum state is that they behave as Bose particles with spin 0. This is no
contradiction to the fact that their constituents are spin 1/2 Fermi particles.
Figure 23 shows very clearly that the individual electrons forming the Cooper
pairs have different momentum vectors p, which however cancel
pairwise such that the pairs have all the same momentum zero. It should be
noted, though, that Cooper pairs differ considerably from other Bosons such as
helium nuclei or atoms: They are not ‘small’ but very extended objects, they
exist only in the BCS ground state and there is no excited state. An excitation is
equivalent to breaking them up into single electrons.
The BCS ground state is characterized by the macroscopic wave function
Ψ and a ground state energy that is separated from the energy levels of the
unpaired electrons by an energy gap. In order to break up a pair an energy of 2∆
is needed, see Fig. 24.

55
Fig. 25: Temperature dependence of the energy gap according to the BCS theory and
comparison with experimental data.

There is a certain similarity with the energy gap between the valence band
and the conduction band in a semiconductor but one important difference is that
the energy gap in a superconductor is not a constant but depends on temperature.
For T → Tc one gets ∆(T) → 0. The BCS theory makes a quantitative prediction

56
for the function ∆(T) which is plotted in Fig. 25 and agrees very well with
experimental data.
One of the fundamental formulae of the BCS theory is the relation between
the energy gap ∆(0) at T = 0, the Debye frequency ωD and the electron-lattice
interaction potential V0 :

. (27)

Here N(EF ) is the density of single-electron states of a given spin orientation at


E = EF (the other spin orientation is not counted because a Cooper pair consists
of two electrons with opposite spin). Although the interaction potential V0 is
assumed to be weak, one of the most striking observations is that the exponential
function cannot be expanded in a Taylor series around V0 = 0 because all
coefficients vanish identically. This implies that Eq. (27) is a truely non-
perturbative result. The fact that superconductivity cannot be derived from
normal conductivity by introducing a ‘small’ interaction potential and applying
perturbation theory (which is the usual method for treating problems of atomic,
nuclear and solid state physics that have no analytical solution) explains why it
took so many decades to find the correct theory. The critical temperature is
given by a similar expression

. (28)
Combining the two equations we arrive at a relation between the energy gap
and the critical temperature which does not contain the unknown interaction
potential

∆(0) = 1.76kBTc . (29)

The following table shows that this remarkable prediction is fulfilled rather
well.

element Sn I T Ta N H Pb
n l b g
∆(0)/ 1.7 1 1 1.7 1.7 2 2.1
kBTc 5 .8 .8 5 5 .3 5

57
Fig. 26: The
critical temperature of various tin isotopes.

In the BCS theory the underlying mechanism of superconductivity is the


attractive force between pairs of electrons that is provided by lattice vibrations.
It is of course highly desirable to find experimental support of this basic
hypothesis. According to Eq. (28) the critical temperature is proportional to the
Debye frequency which in turn is inversely proportional to the square root of the
atomic mass M:
Tc ωD 1/√M .

58
If one produces samples from different isotopes of a superconducting element
one can check this relation.

Figure 26 shows Tc measurements on tin isotopes. The predicted 1/√M law is


very well obeyed.

59
4.4 Supercurrent and critical current
The most important task of a theory of superconductivity is of course to
explain the vanishing resistance. We have seen in sect. 4.1 that the electrical
resistance in normal metals is caused by scattering processes so the question is
why Cooper pairs do not suffer from scattering while unpaired electrons do. To
start a current in the superconductor, let us apply an electric for a short
time δt. Both electrons of a Cooper pair receive an additional momentum δp~ =
−eE0 δt so after the action of the field all Cooper pairs have the same non-
vanishing momentum
.

60
Associated with this coherent motion of the Cooper pairs is a supercurrent
density

. (30)
Here nc is the Cooper-pair density. It can be shown (see e.g. Ibach, Lu¨th [19])
that the Cooper-pair wave function with a current flowing is simply obtained by
multiplying the wave function at rest with the phase factor exp(iK~ · R~) where
R~ = (~r1 + ~r2)/2 is the coordinate of the centre of gravity of the two electrons.
Moreover the electron-lattice interaction potential is not modified by the current
flow. So all equations of the BCS theory remain applicable and there will remain
an energy gap provided the kinetic- energy gain δEpair of the Cooper pair is less
than 2∆, see Fig. 24b. It is this remaining energy gap which prevents scattering.
As we have seen there are two types of scattering centres: impurities and thermal
lattice vibrations. Cooper pairs can only scatter when they gain sufficient energy
to cross the energy gap and are then broken up into single electrons. An impurity
is a fixed heavy target and scattering cannot increase the energy of the electrons
of the pair, therefore impurity scattering is prohibited for the Cooper pairs.
Scattering on thermal lattice vibrations is negligible as long as the average
thermal energy is smaller than the energy gap (that means as long as the
temperature is less than the critical temperature for the given current density). So
we arrive at the conclusion that there is resistance-free current transport
provided there is still an energy gap present (2∆ − δEpair > 0) and the temperature
is sufficiently low (T < Tc(Js)).
The supercurrent density is limited by the condition that the energy gain
δEpair must be less than the energy gap. This leads to the concept of the critical

61
current density Jc. The energy of the Cooper pair is, after application of the
electric field,

with δEpair ≈ pFP/me. From the condition δEpair ≤ 2∆ we get

Js ≤ Jc ≈ 2enc ∆/pF . (31)

Coupled to a maximum value of the current density is the existence of a critical


magnetic field. The current flowing in a long wire of type I superconductor is
confined to a surface layer of thickness λL, see Fig. 6c. The maximum
permissible current density Jc is related to the critical field:

Hc(T) = λL Jc(T) ≈ λL 2enc ∆(T)/pF . (32)

62
The temperature dependence of the critical field is caused by the temperature
dependence of the gap energy.

63
Fig. 27: Persistent ring currents in a benzene molecule and in a superconducting ring which
have been induced by a rising field

The above considerations on resistance-free current flow may appear a bit


formal so I would like to give a more familiar example where an energy gap
prevents ‘resistance’ in a generalized sense. We compare crystals of diamond
and silicon. Diamond is transparent to visible light, silicon is not. So silicon
represents a ‘resistance’ to light. Why is this so? Both substances have exactly
the same crystal structure, namely the ‘diamond lattice’ that is composed of two
face-centred cubic lattices which are displaced by one quarter along the spatial
diagonal. The difference is that diamond is built up from carbon atoms and is an

64
electrical insulator while a silicon crystal is a semiconductor. In the band theory
of solids there is an energy gap Eg between the valence band and the conduction
band. The gap energy is around 7 eV for diamond and 1 eV for silicon. Visible
light has a quantum energy of about 2.5 eV. A photon impinging on a silicon
crystal can lift an electron from the valence band to the conduction band and is
thereby absorbed. The same photon impinging on diamond is unable to supply
the required energy of 7 eV, so this photon simply passes the crystal without
absorption: diamond has no ‘resistance’ for light. (Quantum conditions of this
kind have already been known in the Stone Age. If hunters wanted to catch an
antelope that could jump 2 m high, they would dig a hole 4 m deep and then the
animal could never get out because being able to jump 2 m in two successive
attempts is useless for overcoming the 4 m. The essential feature of a quantum
process, namely that the energy gap has to be bridged in a single event, is
already apparent in this trivial example).
Finally, I want to give an example for frictionless current flow. The
hexagonal benzene molecule C6H6 is formed by covalent binding and contains 24
electrons which are localised in σ bonds in the plane of the molecule and 6
electrons in π bonds below and above this plane. The π electrons can move
freely around the ring. By a time-varying magnetic field a ring current is induced
(benzene is a diamagnetic molecule) which will run forever unless the magnetic
field is changed. This resembles closely the operation of a superconducting ring
in the persistent mode (see Fig. 27).

65
Fig. 28: Magnetic field and vector potential of a solenoid.

5. QUANTISATION OF MAGNETIC FLUX


Several important superconductor properties, in particular the magnetic flux
quantisation, can only be explained by studying the magnetic vector potential
and its impact on the so-called ‘canonical momentum’ of the charge carriers.
Since this may not be a familiar concept I will spend some time to discuss the
basic ideas and the supporting experiments which are beautiful examples of
quantum interference phenomena.

66
5.1 The vector potential in electrodynamics
In classical electrodynamics it is often a matter of convenience to express the
magnetic field as the curl (rotation) of a vector potential
B~ = ×~A~ .
The magnetic flux through an area F can be computed from the line integral of
A
~ along the rim of F by using Stoke’s theorem:

. (33)
We apply this to the solenoidal coil sketched in Fig. 28.

The magnetic field has a constant value B = B0 inside the solenoid and
vanishes outside if we the length of the coil is much larger than its radius R. The
vector potential has only an azimuthal component and can be computed using
Eq. (33):
for r <
R

for r > R .

Evaluating B~ = ×~A~ in cylindrical coordinates gives the expected result

for r < R
z
0 for r > R .

67
What do we learn from this example?
(a) The vector potential is parallel to the current but perpendicular to the
magnetic field.
(b) There are regions in space where the vector potential is non-zero while the
magnetic field vanishes. Here it is the region r > R. A circular contour of
radius r > R includes magnetic flux, namely B0πR2 for all r > R, so A~ must be
non-zero, although B~ = 0.

68
The vector potential is not uniquely defined. A new potential A~0 = A~ + ~
χ with an arbitrary scalar function χ(x,y,z) leaves the magnetic field B~ invariant
because the curl of a gradient vanishes identically. For this reason it is often said
that the vector potential is just a useful mathematical quantity without physical
significance of its own. In quantum theory this point of view is entirely wrong,
the vector potential is of much deeper physical relevance than the magnetic field.

69
Fig. 29: Schematic arrangement for observing the phase shift due to a vector
potential.

5.2 The vector potential in quantum theory


In quantum theory the vector potential is a quantity of fundamental
importance:
A
(1) ~ is the wave function of the photons,
(2) in an electromagnetic field the wavelength of a charged particle is modified
by the vector potential. For the application in superconductivity we are
interested in the second aspect. The de Broglie relation states that the

70
wavelength we have to attribute to a particle is Planck’s constant divided by
the particle momentum

. (34)
For a free particle one has to insert p = mv. It turns out that in the presence of
an electromagnetic field this is no longer correct, instead one has to replace the
mechanical momentum m~v by the so-called ‘canonical momentum’
p~ = m~v + qA~ (35)

where q is the charge of the particle (q = −e for an electron). The wavelength is


then

.
If one moves by a distance ∆x, the phase ϕ of the electron wave function
changes in free space by the amount

.
In an electromagnetic field there is an additional phase change

.
This is called the Aharonov-Bohm effect after the theoreticians who predicted
the phenomenon [20]. The phase shift should be observable in a double-slit
experiment as sketched in Fig. 29. An electron beam is split into two coherent
sub-beams and a tiny solenoid coil is placed between these beams. The sub-
beam 1 travels antiparallel to A~, beam 2 parallel to A~. So the two sub-beams
gain a phase difference

. (36)

71
Fig. 30: Sketch of the M¨ollenstedt-Bayh experiment and observed
interference pattern.

Fig. 31: Observation of Aharonov-Bohm effect using electron holography (after Tonomura
[22]). The permanent toroidal magnet, encapsuled in superconducting niobium, and the
observed interference fringes are shown.

72
Here δϕ0 is the phase difference for current 0 in the coil. The Aharonov-Bohm
effect was verified in a beautiful experiment by Mo¨llenstedt and Bayh in Tu¨bingen
[21]. The experimental setup and the result of the measurements are shown in Fig.
30. An electron beam is split by a metalized quartz fibre on negative potential which
acts like an optical bi-prism.
Very sharp interference fringes are observed. Between the subbeams is a 14 µm–
diameter coil wound from 4 µm thick tungsten wire. The current in this coil is
first zero, then increased linearly with time and after that kept constant. The film
recording the interference pattern is moved in the vertical direction. Thereby the
moving fringes are depicted as inclined lines. The observed shifts are in
quantitative agreement with the prediction of Eq. (36).
An interesting special case is the phase shift δϕ = π that interchanges bright and
dark fringes. According to Eq. (36) this requires a magnetic flux
Φmag = π = e 2e
which turns out to be identical to the elementary flux quantum in
superconductors, see sect. 5.3. In the Mo¨llenstedt experiment however,
continuous phase shifts much smaller than π are visible, so the magnetic flux
through the normal-conducting tungsten coil is not quantised (there is also no
theoretical reason for flux quantisation in normal conductors).
Although the magnetic field is very small outside the solenoid, and the
observed phase shifts are in quantitative agreement with the expectation based
on the vector potential, there have nevertheless been sceptics who tried to
attribute the observed effects to some stray magnetic field. To exclude any such
explanation a new version of the experiment has recently been carried out by
Tonomura et al. [22] making use of electron holography (Fig. 31).
A metalized quartz fibre (the bi-prism) brings the two-part beam to an overlap
on the plate. The magnetic field is provided by a permanently magnetised ring of
with a few µm diameter. The magnet is enclosed in niobium and cooled by liquid
helium so the magnetic field is totally confined. The vector potential, however, is
not shielded by the
superconductor. The field lines of B~ and A~ are also drawn in the figure. The
holographic image shows again a very clear interference pattern and a shift of
the dark line in the opening of the ring which is caused by the vector potential.

73
Fig. 32: Schematic arrangement for studying the properties of a Josephson junction.

6.JOSEPHSON EFFECTS
In 1962 B.D. Josephson made a theoretical analysis of the tunneling of Cooper
pairs through a thin insulating layer from one superconductor to another and
predicted two fascinating phenomena which

74
were fully confirmed by experiment. A schematic experimental arrangement is
shown in Fig. 32.

DC Josephson effect. If the voltage V0 across the junction is zero there is a dc


Cooper-pair current which can assume any value in the range
−I0 < I < I0

where I0 is a maximum current that depends on the Cooper-pair densities and


the area of the junction.
AC Josephson effect. Increasing the voltage of the power supply eventually
leads to a non-vanishing voltage across the junction and then a new phenomenon
arises: besides a dc current which however is now carried by single electrons
there is an alternating Cooper-pair current
I(t) = I0 sin(2πfJt + ϕ0)
whose frequency, the so-called Josephson frequency, is given by the
expression
.
For a voltage V0 = 1µV one obtains a frequency of 483.6 MHz. The quantity ϕ0
is an arbitrary phase.
Equation (55) is the basis of extremely precise voltage measurement.

75
\

7.CONCLUSION
Superconductivity is a set of physical properties observed in superconductors:
materials where electrical resistance vanishes and magnetic fields are expelled from
the material. Unlike an ordinary metallic conductor, whose resistance decreases
gradually as its temperature is lowered, even down to near absolute zero, a
superconductor has a characteristic critical temperature below which the resistance
drops abruptly to zero. An electric current through a loop of superconducting wire
can persist indefinitely with no power source. The superconductivity phenomenon
was discovered in 1911 by Dutch physicist Heike Kamerlingh Onnes. Like
ferromagnetism and atomic spectral lines, superconductivity is a phenomenon

76
which can only be explained by quantum mechanics. It is characterized by the
Meissner effect, the complete cancelation of the magnetic field in the interior of the
superconductor during its transitions into the superconducting state. The occurrence
of the Meissner effect indicates that superconductivity cannot be understood simply
as the idealization of perfect conductivity in classical physics.

77
8.BIBLIOGRAPHY
[1] W. Buckel, Supraleitung, 4. Auflage, VCH Verlagsgesellschaft, Weinheim
1990
[2] D.R. Tilley & J. Tilley, Superfluidity and Superconductivity, Third Edition,
Institute of Physics Publishing Ltd, Bristol 1990
[3] H.K. Onnes, Akad. van Wetenschappen (Amsterdam) 14, 113, 818 (1911)
[4] J. File and R.G. Mills, Phys. Rev. Lett. 10, 93 (1963)
[5] M.N. Wilson, private communication

78
[6] W. Meissner and R Ochsenfeld, Naturwiss. 21, 787 (1933)
[7] F. London and H. London, Z. Phys. 96, 359 (1935)
[8] J. Bardeen, L.N. Cooper and J.R. Schrieffer, Phys. Rev. 108, 1175 (1957)
[9] V.L. Ginzburg and L.D. Landau, JETP USSR 20, 1064 (1950)
[10] L.P. Gorkov, Sov. Phys. JETP 9, 1364 (1960), JETP 10, 998 (1960)

79

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy