You are on page 1of 5

London equations

The London equations, developed by


brothers Fritz and Heinz London in
1935,[1] are constitutive relations for a
superconductor relating its
superconducting current to
electromagnetic fields in and around it.
Whereas Ohm's law is the simplest
constitutive relation for an ordinary
conductor, the London equations are
the simplest meaningful description of
superconducting phenomena, and form
the genesis of almost any modern
introductory text on the subject.[2][3][4]
A major triumph of the equations is
their ability to explain the Meissner
effect,[5] wherein a material
exponentially expels all internal
magnetic fields as it crosses the
superconducting threshold.

Description
As a material drops below its superconducting critical temperature,
There are two London equations when magnetic fields within the material are expelled via the Meissner
expressed in terms of measurable effect. The London equations give a quantitative explanation of this
effect.
fields:

Here is the (superconducting) current density, E and B are respectively the electric and magnetic fields
within the superconductor, is the charge of an electron or proton, is electron mass, and is a
phenomenological constant loosely associated with a number density of superconducting carriers. [6]

The two equations can be combined into a single "London Equation" [6][7] in terms of a specific vector
potential which has been gauge fixed to the "London gauge", giving: [8]

In the London gauge, the vector potential obeys the following requirements, ensuring that it can be
interpreted as a current density:[9]
in the superconductor bulk,
where is the normal vector at the surface of the superconductor.

The first requirement, also known as Coulomb gauge condition, leads to the constant superconducting
electron density as expected from the continuity equation. The second requirement is consistent
with the fact that supercurrent flows near the surface. The third requirement ensures no accumulation of
superconducting electrons on the surface. These requirements do away with all gauge freedom and
uniquely determine the vector potential. One can also write the London equation in terms of an arbitrary
gauge[10] by simply defining , where is a scalar function and is the change in
gauge which shifts the arbitrary gauge to the London gauge. The vector potential expression holds for
magnetic fields that vary slowly in space.[4]

London penetration depth


If the second of London's equations is manipulated by applying Ampere's law,[11]

then it can be turned into the Helmholtz equation for magnetic field:

where the inverse of the laplacian eigenvalue:

is the characteristic length scale, , over which external magnetic fields are exponentially suppressed: it is
called the London penetration depth: typical values are from 50 to 500 nm.

For example, consider a superconductor within free space where the magnetic field outside the
superconductor is a constant value pointed parallel to the superconducting boundary plane in the z
direction. If x leads perpendicular to the boundary then the solution inside the superconductor may be
shown to be

From here the physical meaning of the London penetration depth can perhaps most easily be discerned.

Rationale

Original arguments

While it is important to note that the above equations cannot be formally derived,[12] the Londons did
follow a certain intuitive logic in the formulation of their theory. Substances across a stunningly wide range
of composition behave roughly according to Ohm's law, which states that current is proportional to electric
field. However, such a linear relationship is impossible in a superconductor for, almost by definition, the
electrons in a superconductor flow with no resistance whatsoever. To this end, the London brothers
imagined electrons as if they were free electrons under the influence of a uniform external electric field.
According to the Lorentz force law

these electrons should encounter a uniform force, and thus they should in fact accelerate uniformly. Assume
that the electrons in the superconductor are now driven by an electric field, then according to the definition
of current density we should have

This is the first London equation. To obtain the second equation, take the curl of the first London equation
and apply Faraday's law,

to obtain

As it currently stands, this equation permits both constant and exponentially decaying solutions. The
Londons recognized from the Meissner effect that constant nonzero solutions were nonphysical, and thus
postulated that not only was the time derivative of the above expression equal to zero, but also that the
expression in the parentheses must be identically zero:

This results in the second London equation and (up to a gauge tranformation which is
fixed by choosing "London gauge") since the magnetic field is defined through

Additionally, according to Ampere's law , one may derive that:

On the other hand, since , we have , which leads to the spatial


distribution of magnetic field obeys :

with penetration depth . In one dimension, such Helmholtz equation has the solution form
Inside the superconductor , the magnetic field exponetially decay, which well explains the
Meissner effect. With the magnetic field distribution, we can use Ampere's law again to
see that the supercurrent also flows near the surface of superconductor, as expected from the requirement
for interpreting as physical current.

While the above rationale holds for superconductor, one may also argue in the same way for a perfect
conductor. However, one important fact that distinguishes the superconductor from perfect conductor is that
perfect conductor does not exhibit Meissner effect for . In fact, the postulation

does not hold for a perfect conductor. Instead, the time derivative must be kept and
cannot be simply removed. This results in the fact that the time derivative of field (instead of field)
obeys:

For , deep inside a perfect conductor we have rather than as the superconductor.
Consequently, whether the magnetic flux inside a perfect conductor will vanish depends on the initial
condition (whether it's zero-field cooled or not).

Canonical momentum arguments

It is also possible to justify the London equations by other means.[13][14] Current density is defined
according to the equation

Taking this expression from a classical description to a quantum mechanical one, we must replace values
and by the expectation values of their operators. The velocity operator

is defined by dividing the gauge-invariant, kinematic momentum operator by the particle mass m.[15] Note
we are using as the electron charge. We may then make this replacement in the equation above.
However, an important assumption from the microscopic theory of superconductivity is that the
superconducting state of a system is the ground state, and according to a theorem of Bloch's,[16] in such a
state the canonical momentum p is zero. This leaves

which is the London equation according to the second formulation above.

References
1. London, F.; London, H. (1935). "The Electromagnetic Equations of the Supraconductor" (http
s://doi.org/10.1098%2Frspa.1935.0048). Proceedings of the Royal Society A: Mathematical,
Physical and Engineering Sciences. 149 (866): 71. Bibcode:1935RSPSA.149...71L (https://u
i.adsabs.harvard.edu/abs/1935RSPSA.149...71L). doi:10.1098/rspa.1935.0048 (https://doi.o
rg/10.1098%2Frspa.1935.0048).
2. Michael Tinkham (1996). Introduction to Superconductivity. McGraw-Hill. ISBN 0-07-
064878-6.
3. Neil Ashcroft; David Mermin (1976). Solid State Physics (https://archive.org/details/solidstate
physic00ashc/page/738). Saunders College. p. 738 (https://archive.org/details/solidstatephy
sic00ashc/page/738). ISBN 0-03-083993-9.
4. Charles Kittel (2005). Introduction to Solid State Physics (8th ed.). Wiley. ISBN 0-471-41526-
X.
5. Meissner, W.; R. Ochsenfeld (1933). "Ein neuer Effekt bei Eintritt der Supraleitfähigkeit".
Naturwissenschaften. 21 (44): 787. Bibcode:1933NW.....21..787M (https://ui.adsabs.harvard.
edu/abs/1933NW.....21..787M). doi:10.1007/BF01504252 (https://doi.org/10.1007%2FBF015
04252). S2CID 37842752 (https://api.semanticscholar.org/CorpusID:37842752).
6. James F. Annett (2004). Superconductivity, Superfluids and Condensates (https://archive.or
g/details/superconductivit00anne_663). Oxford. p. 58 (https://archive.org/details/supercondu
ctivit00anne_663/page/n60). ISBN 0-19-850756-9.
7. John David Jackson (1999). Classical Electrodynamics (https://archive.org/details/classicale
lectro00jack_697). John Wiley & Sons. p. 604 (https://archive.org/details/classicalelectro00j
ack_697/page/n627). ISBN 0-19-850756-9.
8. London, F. (September 1, 1948). "On the Problem of the Molecular Theory of
Superconductivity" (https://journals.aps.org/pr/abstract/10.1103/PhysRev.74.562). Physical
Review. 74 (5): 562–573. Bibcode:1948PhRv...74..562L (https://ui.adsabs.harvard.edu/abs/1
948PhRv...74..562L). doi:10.1103/PhysRev.74.562 (https://doi.org/10.1103%2FPhysRev.74.
562).
9. Michael Tinkham (1996). Introduction to Superconductivity (https://archive.org/details/introdu
ctiontosu00tink_948). McGraw-Hill. p. 6 (https://archive.org/details/introductiontosu00tink_94
8/page/n23). ISBN 0-07-064878-6.
10. Bardeen, J. (February 1, 1951). "Choice of Gauge in London's Approach to the Theory of
Superconductivity" (https://link.aps.org/doi/10.1103/PhysRev.81.469.2). Physical Review. 81
(3): 469–470. Bibcode:1951PhRv...81..469B (https://ui.adsabs.harvard.edu/abs/1951PhRv...
81..469B). doi:10.1103/PhysRev.81.469.2 (https://doi.org/10.1103%2FPhysRev.81.469.2).
11. (The displacement is ignored because it is assumed that electric field only varies slowly with
respect to time, and the term is already suppressed by a factor of c.)
12. Michael Tinkham (1996). Introduction to Superconductivity (https://archive.org/details/introdu
ctiontosu00tink_948). McGraw-Hill. p. 5 (https://archive.org/details/introductiontosu00tink_94
8/page/n22). ISBN 0-07-064878-6.
13. John David Jackson (1999). Classical Electrodynamics (https://archive.org/details/classicale
lectro00jack_449). John Wiley & Sons. pp. 603 (https://archive.org/details/classicalelectro00
jack_449/page/n602)–604. ISBN 0-19-850756-9.
14. Michael Tinkham (1996). Introduction to Superconductivity (https://archive.org/details/introdu
ctiontosu00tink_948). McGraw-Hill. pp. 5 (https://archive.org/details/introductiontosu00tink_9
48/page/n22)–6. ISBN 0-07-064878-6.
15. L. D. Landau and E. M. Lifshitz (1977). Quantum Mechanics- Non-relativistic Theory.
Butterworth-Heinemann. pp. 455–458. ISBN 0-7506-3539-8.
16. Tinkham p.5: "This theorem is apparently unpublished, though famous."

Retrieved from "https://en.wikipedia.org/w/index.php?title=London_equations&oldid=1183215790"

You might also like