London equations
   HOME

TheInfoList



OR:

The London equations, developed by brothers
Fritz Fritz originated as a German nickname for Friedrich, or Frederick (''Der Alte Fritz'', and ''Stary Fryc'' were common nicknames for King Frederick II of Prussia and Frederick III, German Emperor) as well as for similar names including Fridolin a ...
and
Heinz London Heinz London (Bonn, Germany 7 November 1907 – 3 August 1970) was a German-British physicist. Together with his brother Fritz London he was a pioneer in the field of superconductivity. Biography London was born in Bonn in a liberal Jewish-Ge ...
in 1935, are
constitutive relation In physics and engineering, a constitutive equation or constitutive relation is a relation between two physical quantities (especially kinetic quantities as related to kinematic quantities) that is specific to a material or substance, and app ...
s 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. A major triumph of the equations is their ability to explain the
Meissner effect The Meissner effect (or Meissner–Ochsenfeld effect) is the expulsion of a magnetic field from a superconductor during its transition to the superconducting state when it is cooled below the critical temperature. This expulsion will repel a ne ...
, wherein a material exponentially expels all internal magnetic fields as it crosses the superconducting threshold.


Description

There are two London equations when expressed in terms of measurable fields: :\frac = \frac\mathbf, \qquad \mathbf\times\mathbf_ =-\frac\mathbf. Here _s is the (superconducting) current density, E and B are respectively the electric and magnetic fields within the superconductor, e\, is the charge of an electron or proton, m\, is electron mass, and n_s\, is a phenomenological constant loosely associated with a number density of superconducting carriers. The two equations can be combined into a single "London Equation" in terms of a specific
vector potential In vector calculus, a vector potential is a vector field whose curl is a given vector field. This is analogous to a ''scalar potential'', which is a scalar field whose gradient is a given vector field. Formally, given a vector field v, a ''vecto ...
\mathbf_s which has been gauge fixed to the "London gauge", giving: :\mathbf_ =-\frac\mathbf_s. In the London gauge, the vector potential obeys the following requirements, ensuring that it can be interpreted as a current density: * \nabla\cdot \mathbf_s = 0, * \mathbf_s = 0 in the superconductor bulk, * \mathbf_s \cdot \hat = 0, where \hat is the normal vector at the surface of the superconductor. The first requirment, also known as
Coulomb gauge In the physics of gauge theories, gauge fixing (also called choosing a gauge) denotes a mathematical procedure for coping with redundant degrees of freedom in field variables. By definition, a gauge theory represents each physically distinct co ...
condition, leads to the constant superconducting electron density \dot \rho_s = 0 as expected from the continuity equation. The second requirment is consistent with the fact that supercurrent flows near the surface. The third requirment 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 \mathbf by simply defining \mathbf_s = (\mathbf + \nabla \phi), where \phi is a scalar function and \nabla \phi 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.


London penetration depth

If the second of London's equations is manipulated by applying Ampere's law, :\nabla \times \mathbf = \mu_0 \mathbf, then it can be turned into the Helmholtz equation for magnetic field: :\nabla^2 \mathbf = \frac\mathbf where the inverse of the laplacian eigenvalue: :\lambda_s \equiv \sqrt is the characteristic length scale, \lambda_s, over which external magnetic fields are exponentially suppressed: it is called the
London penetration depth In superconductors, the London penetration depth (usually denoted as \lambda or \lambda_L) characterizes the distance to which a magnetic field penetrates into a superconductor and becomes equal to e^ times that of the magnetic field at the surface ...
: 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 :B_z(x) = B_0 e^. \, From here the physical meaning of the London penetration depth can perhaps most easily be discerned.


Rationale for the London equations


Original arguments

While it is important to note that the above equations cannot be formally derived, 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 :\mathbf=m \dot=-e\mathbf + e\mathbf \times \mathbf 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 dentisy \mathbf_s = -n_s e \mathbf_s we should have \frac = -n_s e \frac =\frac\mathbf This is the first London equation. To obtain the second equation, take the curl of the first London equation and apply Faraday's law, :\nabla \times \mathbf = -\frac, to obtain : \frac\left( \nabla \times \mathbf_s + \frac \mathbf \right) = 0. 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: \nabla \times \mathbf_s + \frac \mathbf = 0 This results in the second London equation and \mathbf_ =-\frac\mathbf_s (up to a gauge tranformation which is fixed by choosing "London gauge") since the magnetic field is defined through B=\nabla \times A_s. Additionally, accroding to Ampere's law \nabla \times \mathbf = \mu_0 \mathbf_s , one may derive that: \nabla \times (\nabla\times \mathbf) =\nabla \times \mu_0 \mathbf_s =-\frac \mathbf. On the other hand, since \nabla \cdot \mathbf = 0 , we have \nabla \times (\nabla\times \mathbf) = -\nabla^2\mathbf , which leads to the spatial distribution of magnetic field obeys : \nabla^2 \mathbf = \frac\mathbf with penetration depth \lambda_s=\sqrt. In one dimesion, such Helmholtz equation has the solution form B_z(x) = B_0 e^. \, Inside the superconductor (x>0), the magnetic field exponetially decay, which well explains the Meissner effect. With the magnetic field distribution, we can use Ampere's law \nabla \times \mathbf = \mu_0 \mathbf_s again to see that the supercurrent \mathbf_salso flows near the surface of superconductor, as expected from the requirment for intepreting \mathbf_sas 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 pefect conductor is that pefect conductor does not exhibit Meissner effect for T. In fact, the postulation \nabla \times \mathbf_s + \frac \mathbf = 0 does not hold for a pefect conductor. Instead, the time derivative must be kept and cannot be simply removed. This results in the fact that the time derivative of \mathbf field (instead of \mathbf field) obeys: \nabla^2 \frac = \frac\frac. For T, deep inside a perfect conductor we have \dot \mathbf = 0 rather than \mathbf=0 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. Current density is defined according to the equation :\mathbf_s =- n_s e \mathbf_s. Taking this expression from a classical description to a quantum mechanical one, we must replace values \mathbf_s and \mathbf_s by the expectation values of their operators. The velocity operator :\mathbf_s = \frac \left( \mathbf + e \mathbf_s \right) is defined by dividing the gauge-invariant, kinematic momentum operator by the particle mass ''m''. Note we are using -e 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,Tinkham p.5: "This theorem is apparently unpublished, though famous." in such a state the canonical momentum p is zero. This leaves :\mathbf =-\frac\mathbf_s, which is the London equation according to the second formulation above.


References

{{Reflist Superconductivity Equations