23.05.2014 Views

Carsten Timm: Theory of superconductivity

Carsten Timm: Theory of superconductivity

Carsten Timm: Theory of superconductivity

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

6<br />

Ginzburg-Landau theory<br />

Within London and Pippard theory, the superfluid density n s is treated as given. There is no way to understand<br />

the dependence <strong>of</strong> n s on, for example, temperature or applied magnetic field within these theories. Moreover, n s<br />

has been assumed to be constant in time and uniform and in space—an assumption that is expected to fail close<br />

to the surface <strong>of</strong> a superconductor.<br />

These deficiencies are cured by the Ginzburg-Landau theory put forward in 1950. Like Ginzburg and Landau<br />

we ignore complications due to the nonlocal electromagnetic response. Ginzburg-Landau theory is developed as<br />

a generalization <strong>of</strong> London theory, not <strong>of</strong> Pippard theory. The starting point is the much more general and very<br />

powerful Landau theory <strong>of</strong> phase transitions, which we will review first.<br />

6.1 Landau theory <strong>of</strong> phase transitions<br />

Landau introduced the concept <strong>of</strong> the order parameter to describe phase transitions. In this context, an order<br />

parameter is a thermodynamic variable that is zero on one side <strong>of</strong> the transition and non-zero on the other. In<br />

ferromagnets, the magnetization M is the order parameter. The theory neglects fluctuations, which means that<br />

the order parameter is assumed to be constant in time and space. Landau theory is thus a mean-field theory. Now<br />

the appropriate thermodynamic potential can be written as a function <strong>of</strong> the order parameter, which we call ∆,<br />

and certain other thermodynamic quantities such as pressure or volume, magnetic field, etc. We will always call<br />

the potential the free energy F , but whether it really is a free energy, a free enthalpy, or something else depends<br />

on which quantities are given (pressure vs. volume etc.). Hence, we write<br />

F = F (∆, T ), (6.1)<br />

where T is the temperature, and further variables have been suppressed. The equilibrium state at temperature<br />

T is the one that minimizes the free energy. Generally, we do not know F (∆, T ) explicitly. Landau’s idea was to<br />

expand F in terms <strong>of</strong> ∆, including only those terms that are allowed by the symmetry <strong>of</strong> the system and keeping<br />

the minimum number <strong>of</strong> the simplest terms required to get non-trivial results.<br />

For example, in an isotropic ferromagnet, the order parameter is the three-component vector M. The free<br />

energy must be invariant under rotations <strong>of</strong> M because <strong>of</strong> isotropy. Furthermore, since we want to minimize F<br />

as a function <strong>of</strong> M, F should be differentiable in M. Then the leading terms, apart from a trivial constant, are<br />

F ∼ = α M · M + β 2 (M · M)2 + O ( (M · M) 3) . (6.2)<br />

Denoting the coefficients by α and β/2 is just convention. α and β are functions <strong>of</strong> temperature (and pressure<br />

etc.).<br />

What is the corresponding expansion for a superconductor or superfluid? Lacking a microscopic theory,<br />

Ginzburg and Landau assumed based on the analogy with Bose-Einstein condensation that the superfluid part is<br />

described by a single one-particle wave function Ψ s (r). They imposed the plausible normalization<br />

∫<br />

d 3 r |Ψ s (r)| 2 = N s = n s V ; (6.3)<br />

30

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!