You are on page 1of 36

Opleidingsonderdeel G0J16A Solid State Physics I

Beyond the independent electron approximation


[see chapter 17 in Solid State Physics by N.W. Ashcroft and N.D. Mermin] 06 + 13 October 2011

Born-Oppenheimer approximation I

Born - Oppenheimer or adiabatic approximation (1928): the crystal ions (nuclei + core electrons) and the conduction electrons (valence electrons) can be treated separately.

Born-Oppenheimer approximation II
Hamiltonian for conduction electrons with density ~ 1023/cm3:

free electrons

Fermi liquid Bloch electrons describe superconductivity

Born-Oppenheimer approximation III


The sum 1 + 2 violates charge neutrality add uniform positive background jellium Hamiltonian

Hamiltonian for ions with density ~ 1023/cm3:

represents a uniform negative charge background charge neutrality

Free electron gas


High density of states in k-space for macroscopic sample size High density of states as a function of energy:

We have to deal with a liquid rather than with a gas ! Metals have a high heat capacity: cV = 2 kB2T g(F) / 3 = 2 (kBT/F) nkB / 2 Metals have a high electrical conductivity: = ne2 / m = e2 g(F) vF2 / 3

Hartree approximation I
In 1928 Hartree proposed to approximate the electron-electron interaction by considering the interaction of one electron (independent electron approximation) with the charge cloud that is caused by all the other electrons:

Hartree approximation II
This way one obtains a Hartree (Schrdinger) equation for each of the electrons that move into the periodic potential caused by the positive ions:

The Hartree approximation assumes that the many-electron wave function is the product of the one-electron wave functions

The Hartree approximation is a classical approximation that neglects the Pauli principle, i.e. the fact that the total wave function needs to be anti-symmetric

Hartree-Fock approximation I
In 1930 Fock and Slater extended the Hartree approximation by using an anti-symmetric wave function. Such an anti-symmetric wave function, which includes the spatial as well as the spin coordinates, is the Slater determinant:

In most of the cases the Slater determinant will not be an eigen function of the Hamiltonian a variational approach is needed

Hartree-Fock approximation II
Look for a minimum of the expectation value of the Hamiltonian (minimum of the total electron energy):

Calculate the expectation value for the Hamiltonian:

Hartree-Fock approximation III


Minimalize the expectation value of the Hamiltonian using a variational approach (Lagrange multiplicators) to obtain a set of equations of the form:

In the above Hartree-Fock equations the i are the Lagrange multiplicators. It is, however, very tempting to treat the above set of equations as one-electron Schrdinger equations.

Hartree-Fock approximation IV
The theorema of Koopmans implies that the energy needed to transfer an electron from the state with label i to the state with label j is given by the difference j i

The Hartree-Fock equations can be considered as a set of Schrdinger equations that need to be solved self-consistently Introducing the Slater determinant for the many-electron wave function causes the appearance of an extra exchange term in the Schrdinger equation

The exchange term results in a lowering of the electron energy !


The exchange term involves only electrons having the same spin !

Hartree-Fock approximation for free electrons I


Look for normalized plane wave solutions (r) = exp(ikr) / V1/2 that are plugged into the Hartree-Fock equations:

Hartree-Fock approximation for free electrons II


To obtain this result we rely on the fact that for plane waves Uel (Hartree term) is exactly canceled by the uniform positive background that we need to introduce to conserve the charge neutrality! We then proceed by expanding the Coulomb interaction potential VC(r) 1/r into a Fourier series with Fourier components VC(q) 1/q2:

Hartree-Fock approximation for free electrons III

Hartree-Fock approximation for free electrons IV


F(x) is a continuous function of x = k/kF and 0 < F(x) 1

Hartree-Fock approximation for free electrons V

The derivative F(x) has a logarithmic singularity for x = 1, i.e. for k = kF the Fermi velocity vF = /k(=F)/ The heat capacity cV T/|lnT| will deviate from a linear temperature dependence, which clearly is in conflict with experiment! On the other hand, the concept of a Fermi sea, which is densely filled with electrons up to the Fermi level, remains intact

Hartree-Fock approximation for free electrons VI

BANDWIDTH

The energy levels are considerably lowered The width of the band increases by a factor 2.3

Hartree-Fock approximation for free electrons VII


Within the Hartree-Fock approximation the total energy E of the free electron gas is given by

The summations over the allowed k-vectors can be replaced with an integral. When taking properly into account the logarithmic divergence of the second term, the total energy is given by

Hartree-Fock approximation for free electrons VIII


The Hartree-Fock equation for the electron with wave function i(r) can be interpreted in terms of two electron charge densities. The first density el is the uniform negative density that follows from the Hartree approximation:

The second charge density is a positive density that can be associated with the exchange term. For the electron with wave function i(r) it is given by

Hartree-Fock approximation for free electrons IX


The extra charge density resulting from the exchange can be simply visualized if we average over all electrons i and replace the summation with an integration. We then plot the absolute value of the difference between el and the averaged density related to the exchange:

charge density

Each electron drags along an exchange hole with the same spin!

Hartree-Fock approximation for free electrons X


The non-locality (dependence on r and r) of the exchange term makes a self-consistent solution of the Hartree-Fock equations impossible for electrons that move in a periodic potential ! While it is straightforward to write down appropriate equations, it is unfortunately not possible to solve them for large numbers of electrons. This was already noted in 1929 by Dirac.

Even with the most powerful computers numerical solutions become impossible for N around 10.
While for atoms and molecules one tries to do better than Hartree-Fock, even more crude approximations are required for treating the many electrons in a solid. We can get a first idea of how crude the approximations are from the discussion of equations (17.25) and (17.26) in chapter 17 in the book of Ashcroft and Mermin.

Screening I
We introduce an external charge density ext(r) in a gas of free electrons The free electron gas reacts by creating a charge density ind(r) and the total charge density (r) is given by

(r) and ext(r) are linked to potentials (r) and ext(r) via Poissons equation:

Screening results in a reduction of the potential that can be described in terms of a (relative) dielectric function r(r,r):

Screening II
For a homogeneous medium we have that r(r,r) = r(r-r) Switch to k-space:

This way, we have a wave vector dependent dielectric constant of the metal that is related to the wave vector dependent potentials:

From a computational point of view it is more convenient to calculate the induced charge density ind(r) (use a linear approximation):

Screening III
Introduce the Fourier transforms of the Poisson equations to link the wave vector dependent dielectric constant to the wave vector dependent induced charge density:

Screening IV
We obtain the following equation for the dielectric function that needs to be solved self-consistently:

The Thomas-Fermi approximation I


In principle we need to solve the Schrdinger equation:

determines the charge density (r)

Provided (r) varies very slowly on a scale corresponding to F, we can according to Ehrenfests theorem rely on a classical approach:

The spatial variation of the electron density n(r) is given by

The Thomas-Fermi approximation II


We need to take into account the positive charge background:

We can formally rewrite ind(r) as

We assume that (r) is a small perturbation Use a series expansion for n0:

The susceptibility (q) is then a constant that is independent of q and is given by

The Thomas-Fermi approximation III


Finally, the dependence on q of the dielectric constant is given by

The Fourier components of the Coulomb potential need to be adapted according to

There no longer occurs a divergence for q 0 !

The Thomas-Fermi approximation IV


We can then investigate how a point charge is screened:

The screening of the Coulomb potential can be compared to the screened Yukawa potential in nuclear physics !

The Thomas-Fermi approximation V


There occurs a very effective screening of an external charge at a distance r >> k0-1:

a0 represents the Bohr radius ( 0.5 )

The Lindhard approximation I


For this approximation we need to solve the Schrdinger equation ! We assume a linear relation between ind and :

When q 0, the Fermi-Dirac distribution function can be approximated by a series expansion:

The linear term reproduces the Thomas-Fermi approximation. For arbitrary q we need to solve the integral. For T 0 the solution is

The Lindhard approximation II


In the above expression for the susceptibility we have defined x = q/2kF.

For r >> kF-1 we can calculate how a point charge is screened by the electron gas:

The long-range charge density oscillations that surround a point charge impurity are know as Friedel oscillations The physical origin of the Friedel oscillations can be directly linked to the sharpness of the Fermi surface: Fourier components of the Coulomb potential with wave vector q > kF cannot be screened since there are no electrons with k > kF ! The extra factor of 2 in the cosine dependence results from the fact that the charge density is determined by the square of the wave function

Fourier-transform STM I
Friedel charge density oscillations will also be present for the twodimensional (2D) electron gas at the (111) surface of noble metals (surface states that are decoupled from the bulk states) STM images the electron charge density From a Fourier analysis of the charge density oscillations the Fermi circle can be reconstructed !

Fermi surface of bulk (3D) copper


(measured by the de Haas van Alphen effect)

There are no allowed bulk states within the necks along the [111] directions! [See chapter 9 in the book of Kittel on Solid State Physics] Surface states, which are decoupled from the bulk states, can be formed within the energy gaps of the projected bulk states

Fourier-transform STM II
Look at surface electrons for more complicated, highly deformed Fermi surfaces

Fourier-transform STM III

From the Fourier transform of the charge density oscillations at a step edge on the Be(10-10) surface the allowed states in the surface Brillouin zone (SBZ) can be reconstructed A pronounced anisotropy of the screening is present, consistent with the theoretical expectations

You might also like