Jump back to chapter selection.


Table of Contents

3.1 The Jellium Model of the Metallic State
3.2 Charge Excitations and the Dielectric Function
3.3 Lattice Vibrations - Phonons
3.4 Appendix: Linear Response Theory


3 Metals

The electronic states in a periodic atomic lattice are extended and have an energy spectrum forming energy bands. In the ground state these energy states are filled successively starting at the bottom of the electronic spectrum until the number of electrons is exhausted. Metallic behaviour occurs whenever in this way a band is only partially filled. The fundamental difference that distinguishes metals from insulators and semiconductors is the absence of a gap for electron-hole excitations. In metals, the ground state can be excited at arbitrarily small energies which has profound phenomenological consequences.
We will consider a basic model suitable for the description of simple metals like the alkali metals Li, Na, or K, where the (atomic) electron configuration consists of closed shell cores and one single valence electron in an ns-orbital. Neglecting the core electrons (completely filled bands), we consider the valence electrons only and apply the approximation of nearly free electrons. The lowest band around the Γ-point is then half-filled. First, we will also neglect the influence of the periodic lattice potential and consider the problem of a free electron gas subject to mutual (repulsive) Coulomb interaction.


3.1 The Jellium Model of the Metallic State

The Jellium model is probably the simplest possible model of a metal that is able to describe qualitative and to some extent even quantitative aspects of simple metals. The main simplification made is to replace the ionic lattice by a homogeneous positively charged background (Jellium). The uniform charge density enion is chosen such that the whole system - electrons and ionic background - is charge neutral, that is nion=n, where n is the electron density. In this fully translational invariant system, the plane waves

ψk,s(r)=1Ωeikr

represent the single-particle wave functions of the free electrons. Here Ω is the volume of the system, k and s{,} denote the wave vector and spin, respectively. Assuming a cubic system of side length L and volume Ω=L3 we impose periodic boundary conditions for the wave function

ψk,s(r+(L,0,0))=ψk,s(r+(0,L,0))=ψk,s(r+(0,0,L))=ψk,s(r)

such that the reciprocal space is discretised as

k=2πL(nx,ny,nz)

where (nx,ny,nz)Z3. The energy of a single-electron state is given by ϵk=2|k|2/2m (free particle). The ground state of non-interacting electrons is obtained by filling all single particle states up to the Fermi energy with two electrons. In the language of second quantisation the ground state is, thus, given by

|Ψ0=|k|kFsc^k,s|0

where the operators c^k,s(c^k,s) create (annihilate) an electron with wave vector k and spin s. We call this state also a filled Fermi sea. The Fermi wave vector kF with the corresponding Fermi energy ϵF=2kF2/2m is determined by equating the filled electronic states with the electron density n. We have

n=1Ω|k|kF,s1=2d3k(2π)31=24π3kF3(2π)3

which results in

kF=(3π2n)1/3.

Note kF is the radius of the Fermi sphere in k-space around k=0.

3.1.1 Theory of Metals - Sommerfeld and Pauli

In a first step we neglect the interaction among the electrons and consider the electrons in the metal simply as a Fermi gas. Then thermodynamic properties can be described by using the Fermi-Dirac distribution function,

f(ϵk)=1e(ϵkμ)/kBT+1

and the density of states

N(E)=1Ωk,sδ(Eϵk)=2d3k(2π)3δ(E2|k|22m)=14π3dΩkdkk2m2kδ(k2mE)=12π2(2m2)3/2E1/2=32nϵF(EϵF)1/2

for E>0. We first address the temperature dependence of the chemical potential up second power in T for fixed electron density n, by using the equation

n=1Ωk,sf(ϵk)=0+dEf(E)N(E)=0μdEN(E)+π26(kBT)2N(μ)+

where we used the Sommerfeld expansion (see next paragraph) assuming TTF=ϵF/kB.

Sommerfeld expansion: In the limit kBTϵF the derivative f(E)/E is well concentrated around E=μ. We consider

+dEg(E)(f(E)E)=+dE{g(μ)+(Eμ)g(μ)+(Eμ)22g(μ)+}(f(E)E)=g(μ)+g(μ)2+dE(Eμ)2(f(E)E)+=g(μ)+π26g(μ)(kBT)2+.

and analogous

+dE(g(E)E)f(E)=g(μ)+π26g(μ)(kBT)2+.

With the definition

g(E)=EdEΓ(E)+dEΓ(E)f(E)=μdEΓ(E)+π26(kBT)2Γ(μ)+

Note that

+dxβx2eβx(eβx+1)2=π23β2.

We now use

0μdEN(E)0ϵFdEN(E)+(μϵF)N(ϵF)=n+(μϵF)N(ϵF)

leading to

nn+(μϵF)N(ϵF)+π26(kBT)2N(ϵF)μ(T)=ϵFπ26(kBT)2N(ϵF)N(ϵF)+

with N(ϵF)/N(ϵF)=1/(2ϵF). Now we also determine the internal energy density,

u(T)=U(T)Ω=0dEEN(E)f(E)0μdEEN(E)+π26(kBT)2{μN(μ)+N(μ)}0ϵFdEEN(E)+ϵF{(μϵF)N(ϵF)+π26(kBT)2N(ϵF)}=0+π26(kBT)2N(ϵF)=u0+π26(kBT)2N(ϵF).

The specific heat is then given by

C=uT=π23ΩkB2TN(ϵF)=γT

and shows a T-linear behaviour where γ is the Sommerfeld coefficient, proportional to the density of states at the Fermi energy.
Next we consider the effect of a magnetic field coupling to the electron spin, so that ϵkϵk,s=ϵkμBsH with μB the Bohr magneton and s=±1. We consider the magnetisation due to the spin polarisation of the electrons,

M=μB(n+n)=μB2{0dEN(E)f(EμBH)0dEN(E)f(E+μBH)}μB20dEN(E)(f(E)E)2μBHμB2HN(ϵF)0dE(f(E)E)=μB2HN(ϵF).

By taking the derivative with respect to H we find for the susceptibility,

χp=MH=μB2N(ϵF).

This is the Pauli paramagnetic susceptibility which is to lowest order temperature independent and, like γ, is proportional to the density of states at the Fermi energy.
The temperature dependence of the χp can be found by going beyond the lowest order approximation:

MμB2H0dEN(E)(f(E)E)=μB2H{N(μ)+π26(kBT)2N(μ)}

We then write

N(μ)N(ϵFπ26(kBT)2N(ϵF)N(ϵF))N(ϵF)π26(kBT)2N(ϵF)2N(ϵF),

which leads to

MμB2HN(ϵF)[1π26(kBT)2{(N(ϵF)N(ϵF))2N(ϵF)N(ϵF)}]=χp(T)H,

and defines the temperature dependent spin susceptibility, which depends on details of the density of states.

3.1.2 Stability of Metals - a Hartree-Fock Approach

Now we would like to examine the stability of the Jellium model. For this purpose, we compute the ground state energy of the Jellium system variationally, using the density n as a variational parameter, which is equivalent to the variation of the lattice constant. In this way, we will obtain an understanding of the stability of a metal, that is, the cohesion of the ion lattice through the itinerant electrons (in contrast to semiconductors where the stability was due to covalent chemical bonding). The variational ground state shall be |Ψ0 for given kF. The Hamiltonian splits into four terms

H=Hkin+Hee+Hei+Hii

with

Hkin=k,sϵkc^ksc^ksHee=12s,sd3rd3rΨ^s(r)Ψ^s(r)e2|rr|Ψ^s(r)Ψ^s(r)Hei=sd3rd3rne2|rr|Ψ^s(r)Ψ^s(r)Hii=12d3rd3rn2e2|rr|,

where we have used in second quantisation language the electron field operators

Ψ^s(r)=1Ωkc^k,seikrΨ^s(r)=1Ωkc^k,seikr

The variational energy - which we want to minimise with respect to n - can be computed from Eg=Ψ0|H|Ψ0 and consists of four different contributions:
First we have the kinetic energy

Ekin=Ψ0|Hkin|Ψ0=k,sϵkΨ0|c^ksc^ks|Ψ0=nks=2Ωd3k(2π)3ϵknks=N35ϵF

where we used N=Ωn the number of valence electrons and

nks={1|k|kF0|k|>kF.

Secondly, there is the energy resulting from the Coulomb repulsion between the electrons,

Eee=12d3rd3re2|rr|s,sΨ0|Ψ^s(r)Ψ^s(r)Ψ^s(r)Ψ^s(r)|Ψ0=12d3rd3re2|rr|(n2G(rr))=EHartree+EFock.

For this contribution we used the fact, that the two-particle correlation function may be expressed as

s,sΨ0|Ψ^s(r)Ψ^s(r)Ψ^s(r)Ψ^s(r)|Ψ0=n2G(rr)

where

G(r)=9n22(kF|r|cos(kF|r|)sin(kF|r|)(kF|r|)3)2.

The Coulomb repulsion Hee between the electrons leads to two terms, called the direct or Hartree term describing the Coulomb energy of a uniformly spread charge distribution, and the exchange or Fock term resulting from the exchange hole that follows from the Fermi-Dirac statistics (Pauli exclusion principle).
The third contribution originates in the attractive interaction between the (uniform) ionic background and the electrons,

Eei=d3rd3re2|rr|nsΨ0|Ψ^s(r)Ψ^s(r)|Ψ0=d3rd3re2|rr|n2,

where the expectation value Ψ0|Ψ^s(r)Ψ^s(r)|Ψ0 corresponds to the uniform density n, as is easily calculated.
Finally we have the repulsive ion-ion interaction

Eii=Ψ0|Hii|Ψ0=12d3rd3rn2e2|rr|.

Attachments/Script 27.webp|700

It is easy to verify that the three contributions EHartree,Eei, and Eii compensate each other to exactly zero. Note that these three terms are the only ones that would arise in a classical electrostatic calculation, implying that the stability of metals relies purely on quantum effect. The remaining terms are the kinetic energy and the Fock term. The latter is negative and reads

EFock=Ω9n24d3re2|r|(sin(kF|r|)kF|r|cos(kF|r|)(kF|r|)3)2=N3e24πkF.

Eventually, the total energy per electron is given by

EgN=352kF22m3e24πkF=(2.21rs20.916rs)Ry

where 1Ry=e2/2aB and the dimensionless quantity rs is defined via

n=34πd3

and

rs=daB=(9π4)1/3me22kF.

The length d is the average radius of the volume occupied by one electron. Minimising the energy per electron with respect to n is equivalent to minimise it with respect to rs, yielding rs,min=4.83,d2.5Å and a density of n01.5×1024cm3. This corresponds to a lattice constant of a=(4π/3)1/3d4Å. This estimate is roughly in agreement with the lattice constants of the alkali metals: rs,Li=3.22,rs,Na=3.96,rs,K=4.86. Note that in metals the delocalised electrons are responsible for the cohesion of the positive background yielding a stable solid.

Attachments/Script 28.webp|700

The good agreement of this simple estimate with the experimental values is due to the fact that the alkali metals have only one valence electron in an s-orbital that is delocalised, whereas the core electrons are in a noble gas configuration and, thus, relatively inert. In the variational approach outlined above correlation effects among the electrons due to the Coulomb repulsion have been neglected. In particular, electrons can be expected to 'avoid' each other not just because of the Pauli principle, but also as a result of the repulsive interaction. However, for the problem under consideration the correlation corrections turn out to be small for rsrs,min:

EtotN=(2.21rs20.916rs+0.062lnrs0.096+)Ry

which can be obtained from a more sophisticated quantum field theoretical analysis.


3.2 Charge Excitations and the Dielectric Function

In analogy to semiconductors, the elementary excitations of metallic systems are the electron-hole excitations, which for metals, however, can have arbitrarily small energies. One particularly drastic consequence of this behaviour is the strong screening of the long-ranged Coulomb potential. As we will see, a negative test charge in a metal reduces the electron density in its vicinity, and the induced cloud of positive charges, relative to the uniform charge density, weaken the Coulomb potential as,

V(r)1rV(r)er/lr

that is, the Coulomb potential is modified into the short-ranged Yukawa potential with screening length l. In contrast to metals, due to the finite energy gap for electron-hole excitations, the charge distribution in semiconductors reduces the adaption of the system to perturbations, so that the screened Coulomb potential remains long-ranged,

V(r)1rV(r)1εr

As mentioned earlier, the semiconductor acts as a dielectric medium and its screening effects are accounted for by the polarisation of localised electric dipoles, that is, the Coulomb potential inside a semiconductor is renormalised by the dielectric constant ε.

3.2.1 Dielectric Response and Lindhard Function

We will now investigate the response of an electron gas to a time- and position-dependent weak external potential Va(r,t) in more detail based on the equation of motion. We introduce the Hamiltonian

H=Hkin+HV=H0+HV=k,sϵkc^ksc^ks+sd3rVa(r,t)Ψ^s(r)Ψ^s(r)

where the second term is considered as a small perturbation to our system described by a time-independent Hamiltonian, H0=Hkin whose properties we know exactly. In a first step we consider the linear response of the system to the external potential. On this level we restrict ourselves to one Fourier component in the spatial and time dependence of the potential,

Va(r,t)=Va(q,ω)eiqriωteηt

where η0+includes the adiabatic switching on of the potential. To linear response (order), this potential induces a small modulation of the electron density of the form nind(r,t)=n0+δnind(r,t) with

δnind(r,t)=δnind(q,ω)eiqriωteηt.

We obtain for the density operator in momentum space,

ρ^q=sd3rΨ^s(r)Ψ^s(r)eiqr=k,sc^ksc^k+q,s=k,sρ^k,q,s,

where we define ρ^k,q,s=c^ksc^k+q,s. The perturbation term HV now reads

HV=sd3rVa(r,t)Ψ^s(r)Ψ^s(r)=ρ^qVa(q,ω)eiωteηt=k,sρ^k,q,sVa(q,ω)eiωteηt.

The density operator ρ^q(t) in Heisenberg representation is the relevant quantity needed to describe the electron density in the metal.

Linear response

We introduce the equation of motion for ρ^k,q,s(t):

iddtρ^k,q,s=[ρ^k,q,s,H]=[ρ^k,q,s,Hkin+HV]=(ϵk+qϵk)ρ^k,q,s+(c^ksc^ksc^k+q,sc^k+q,s)Va(q,ω)eiωteηt

We now take the thermal average A^0=Tr[A^eβH0]/Tr[eβH0] with respect to the unperturbed Hamiltonian and follow the linear response scheme by assuming the same time dependence for ρ^k,q,s(t)0=ρ^k,q,s0(ω)eiωt+ηtVa(q,ω)eiωt+ηt as for the potential, so that the equation of motion reads,

(ω+iη)ρ^k,q,s0(ω)=(ϵk+qϵk)ρ^k,q,s0(ω)+(n0k,sn0k+q,s)Va(q,ω)

where n0k,s=c^ksc^ks0. Note that we take here the approximation that for the limit Va0 we keep only terms linear Va, such that n0k,s is independent of Va and thus of time, while ρ^k,q,s0 is proportional to Va. This leads then consistently to

δnind(q,ω)=1Ωk,sρ^k,q,s0(ω)=1Ωk,sn0k+q,sn0k,sϵk+qϵkωiηVa(q,ω).

With this, we define the dynamical linear response function as

χ0(q,ω)=1Ωk,sn0k+q,sn0k,sϵk+qϵkωiη

such that δnind(q,ω)=χ0(q,ω)Va(q,ω), where χ0(q,ω) is known to be the Lindhard function. Note the Fourier transform in real space and time yields the relation,

δnind(rr,tt)=d3rdtχ0(rr,tt)Va(r,t)

The adiabatic switching on of the perturbation has mathematically the nice regularisation feature that χ0(r,t) is causal, that is χ0(r,t)=0 for t<0. The potential Va can only influence the response into the future. Note, that for n0k,s we can use the Fermi-Dirac distribution function f(ϵk) for spin-independent densities.

As a simple example we consider here a metal at T=0 exposed to a uniform static potential, which corresponds simply to a shift of the chemical potential: Va(r,t)=δμ. Thus, we use ω=0 and take in the Lindhard function the limit q0. Using Bernoulli-L'Hôpital rule for the q-limit we find

χ0(q0,0)1Ωk,sqkϵkn0(ϵk)ϵkqkϵk=1Ωk,sf(ϵk)ϵk=1Ωk,sδ(ϵkϵF)=N(ϵF)Ω.

The electron density shift is then given by

δnind=N(ϵF)ΩVa=N(ϵF)Ωδμ

which agrees with the Sommerfeld approximation,

δn=1ΩϵFϵF+δμdEN(E)N(ϵF)Ωδμ.

This expression is in linear order exact. Note that this is connected with the compressibility κ of the electrons,

κ=1n2nμ=χ0(0,0)n2=1n2ΩN(ϵF)=32nϵF

where the last equality is the expression for free electrons.

Coulomb interaction - Random phase approximation

So far we treated the linear response of the system to an external perturbation without considering "feedback effects" due to the interaction among electrons. In fact, the density fluctuation δn(r,t) can be thought as a source for an additional Coulomb potential Vδn which can be determined by means of the Poisson equation,

2Vδn(r,t)=4πe2δn(r,t)

or in Fourier space

Vδn(q,ω)=4πe2|q|2δn(q,ω).

If we allow feedback effects in our system with external perturbation Va(q,ω), the effective potential V felt by the electrons is determined self-consistently via

V(q,ω)=Va(q,ω)+Vδn(q,ω)=Va(q,ω)+4πe2|q|2δn(q,ω),

where

δn(q,ω)=χ0(q,ω)V(q,ω).

The relation between V and Va may then be written as

V(q,ω)=Va(q,ω)ε(q,ω)

with

ε(q,ω)=14πe2|q|2χ0(q,ω)

where ε(q,ω) is termed the dynamical dielectric function and describes the renormalisation of the external potential due to the dynamical response of the electrons in the metal. Extending

δn(q,ω)=χ0(q,ω)V(q,ω)=χ(q,ω)Va(q,ω),

we define the response function χ(q,ω) within "random phase approximation" to be

χ(q,ω)=χ0(q,ω)ε(q,ω)=χ0(q,ω)14πe2|q|2χ0(q,ω).

This response function χ(q,ω) contains also effects of electron-electron interaction and comprises information not only about the renormalisation of potentials, but also on the excitation spectrum of the metal.

3.2.2 Electron-Hole Excitation

The most simple excitation in a metal is the electron-hole excitation which resembles in some way that discussed for the semiconductor. Neglecting the Coulomb interaction we remove an electron from an occupied state and place it into a state which is unoccupied:

Attachments/Script 29.webp|700

Since the ground-state |Ψ0 is the filled Fermi sea, we remove the electron with an energy ϵkϵF (|k|kF) and place into an energy state outside the Fermi sea, ϵk+q>ϵF. Thus, the excited state is given by

|k+q,s;k,s=c^k+q,sc^k,s|Ψ0=ρ^k,q,s|Ψ0

with the constraint that n0,k(1n0,k+q)=1 (we assume that the spin s remains unchanged in the excitation, what we call a pure charge excitation). Note that n0,k=Θ(ϵFϵk) at zero temperature. Analogous to the semiconductors the excitation energy is given by

Ek,q=ϵk+qϵk>0.

Also here we find a continuum of electron-hole excitation spectrum in the energy-momentum plane - sketched in a conceptual diagram. In contrast to semiconductors electron-hole excitations are possible to arbitrarily low energies. The possible momentum transfer is dictated by the geometry of the Fermi sea. For Ek,q0 the momentum transfer ranges from q0 to q2kF as the electron has to be removed just below and be place just above the Fermi energy. For increasing the excitation energy this momentum range is gradually shift as depicted as the blue area:

Attachments/Script 30.webp|700

Interestingly |k+q,s;k,s can be generated through the operator ρ^k,q,s which also couples to the external potential which we used the derive the linear response theory. The linear response function is actually built upon the properties of electron-hole excitation. Indeed χ(q,ω) contains information about the excitation spectrum. Without Coulomb it is sufficient to consider χ0(q,ω), the Lindhard function.
We may separate χ0 into its real and imaginary part, χ0(q,ω)=χ0(q,ω)+iχ0(q,ω). Using the relation

limη0+1ziη=P(1z)+iπδ(z)

where the Cauchy principal value P of the first term has to be taken, we separate the Lindhard function into

χ0(q,ω)=1Ωk,sP(n0,k+qn0,kϵk+qϵkω)andχ0(q,ω)=πΩk,s(n0,k+qn0,k)δ(ϵk+qϵkω).

The real part will be important later in the context of instabilities of metals. The excitation spectrum is visible in the imaginary part which relates to the absorption of energy by the electrons subject to a time-dependent external perturbation. Note that χ0(q,ω) corresponds to Fermi's golden rule of time-dependent perturbation theory, that is, the transition rate from the ground state to an excited state of energy ω and momentum q for a perturbation. The sum over k yields the density of possible electron-hole excitations with the given excitation energy.

3.2.3 Collective Excitation

Similar to semiconductors, for metals also the Coulomb interaction yields a further type of excitation which, however, cannot be described by a simple superposition of electron-hole excitations, unlike excitons. Here the renormalised response function χ(q,ω), including the Coulomb interaction, serves as a means to uncover collective excitations beyond the level of independent electrons. It is the long-range nature of the Coulomb interaction which is responsible for the so-called plasma resonance which appears at rather high energies for small momenta q. We analyse χ(q,ω) in the small |q|-limit, that is |q|kF where we expand χ0(q,ω) in q, starting with

ϵk+q=ϵk+qkϵk+n0,k+q=n0,k+n0ϵkqkϵk+

Note that n0/ϵk=δ(ϵkϵF) at T=0 and kϵk=vk gives the velocity. Since only states located at the Fermi energy are relevant here, vk=vFk/|k| is the Fermi velocity. This leads to the approximation (with q=|q|)

χ0(q,ω)2d3k(2π)3qvkδ(ϵkμ)qvkωiη=2(2π)21+1dcosθkF2vF[qvFcosθω+iη+(qvFcosθω+iη)2+(qvFcosθω+iη)3+]kF3q23π2m(ω+iη)2(1+35vF2q2(ω+iη)2)=n0q2m(ω+iη)2(1+35vF2q2(ω+iη)2).

First we consider the limit q0 for the dielectric function,

ε(q0,ω)=14πe2|q|2n|q|2mω2=1ωp2ω2

with the plasma frequency defined as

ωp2=4πe2n0m.

Note that the long-range nature of the Coulomb interaction (manifest in 4πe2/|q|2) yields a finite plasma frequency, as this cancels the |q|2-dependence of χ0.

We now approximate χ(q,ω),

χ(q,ω)n0|q|2R(|q|,ω)2m[(ω+iη)2ωp2R(|q|,ω)2]=n0|q|2R(|q|,ω)2mωp{1ω+iηωpR(|q|,ω)1ω+iη+ωpR(|q|,ω)}

where we introduced

R(|q|,ω)2=(1+3vF2|q|25ω2).

Note that we keep η only in lowest order and ignore it where ever it would appear in higher than linear power. Thus, η is dropped in the definition of R. We obtain for the imaginary part of χ(q,ω),

χ(q,ω)πn0|q|2R(|q|,ωp)2mωp[δ(ωωpR(|q|,ωp))δ(ω+ωpR(|q|,ωp))]

which yields a sharp excitation mode,

ω(q)=ωpR(|q|,ωp)=ωp{1+3vF2|q|210ωp2+},

which is called plasma resonance with ωp as the plasma frequency. Similar to the exciton, the plasma excitation has a well-defined energy-momentum relation and may consequently be viewed as a quasi-particle (plasmon) which has bosonic character. When the plasmon dispersion merges with the electron-hole continuum it is damped (Landau damping) because of the allowed decay into electron-hole excitations. This results in a finite life-time of the plasmons within the electron-hole continuum corresponding to a finite width of the resonance of the collective excitation.

It is possible to understand the plasma excitation in a classical picture. Consider negatively charged electrons in a positively charged ionic background. When the electrons are shifted uniformly by r with respect to the ions, a polarisation P=n0er results. The polarisation causes an electric field E=4πP which acts as a restoring force. The equation of motion for an individual electron describes harmonic oscillations

md2dt2r=eE=4πe2n0r

with the same oscillation frequency as the plasma frequency,

ωp2=4πe2n0m.

Classically, the plasma resonance can therefore be thought as an oscillation of the whole electron gas cloud on top of a positively charged background.

Attachments/Script 31.webp|700

3.2.4 Screening

Thomas-Fermi screening

Next, we analyze the potential V felt by the electrons exposed to a static field (ω0). Using the expansion of the Lindhard function previously derived we obtain

χ0(q,0)=1Ωk,sδ(ϵkϵF)=1π2kF2vF=3n02ϵF

and thus

ε(q,0)=1+kTF2|q|2

with the so-called Thomas-Fermi wave vector kTF2=6πe2n0/ϵF. The effect of the renormalised |q|-dependence of the dielectric function can best be understood by considering a bare point charge Va(r)=e2/|r| (or Va(q)=4πe2/|q|2) and its renormalisation in momentum space

V(q)=Va(q)ϵ(q,0)=4πe2|q|2+kTF2

or in real space

V(r)=e2|r|ekTF|r|.

The potential is screened by a rearrangement of the electrons and this turns the long-ranged Coulomb potential into a Yukawa potential with exponential decay. The new length scale is kTF1, the so-called Thomas-Fermi screening length. In ordinary metals kTF is typically of the same order of magnitude as kF, that is the screening length is of order 5Å comparable to the distance between neighboring atoms. As a consequence also external electric fields cannot penetrate a metal, but are screened on this length 1/kTF. This legitimates one of the basic assumptions used in electrostatics with metals.

Friedel oscillations

The static dielectric function can be evaluated exactly for a system of free electrons, resulting for 3 dimensions in (with q=|q|)

ε(q,0)=1+4e2mkFπ2q2{12+4kF2q28kFqln|2kF+q2kFq|}.

Noticeably the dielectric function varies little for small |q|kF. At |q|=2kF there is, however, a logarithmic singularity. This is a consequence of the sharpness of the Fermi surface in k-space. Consider the induced charge of a point charge at the origin: ena(r)=ena0δ(r) which Fourier transformed is na(q)=na0.

δn(r)=d3q(2π)3{1ε(|q|)1}na(q)eiqr=1|r|0g(q)na(q)sin(q|r|)dq

with (q=|q|)

g(q)=q2π2ε(q)1ε(q).

Note that g(q) vanishes for both q0 and q. Using partial integration twice, we find (with r=|r|)

δn(r)=na0r30g(q)sin(qr)dq

where

g(q)Aln|q2kF|

and

g(q)Aq2kF

dominate around q2kF. Hence, for kFr1,

δn(r)Ana0r32kFΛ2kF+Λsin[(q2kF)r]cos(2kFr)+cos[(q2kF)r]sin(2kFr)q2kFdqπAna0cos(2kFr)r3

with a cutoff Λ. The induced charge distribution exhibits so-called Friedel oscillations. Finally we may ask what is the total electron charge displaced around the point charge ena0δ(r). We integrate over r:

δQ=ed3rδn(r)=limq0{1ε(|q|)1}na(q)=ena0

where we used 1/ϵ(q)0 for q0. The charge displacement corresponds to the exact opposite amount of charge of the point charge. Thus we find a perfect compensation which corresponds to perfect screening.

Attachments/Script 32.webp|700

Dielectric function in various dimensions

Above we have treated the dielectric function for a three-dimensional parabolic band. Similar calculations can be performed for one- and two-dimensional systems. In general, the static susceptibility is given by (q=|q|)

χ0(q,ω=0)={12πqvFln|s+2s2|,1Dm2π2{1(14s2)1/2θ(s2)},2DmkF2π22{1s4(14s2)ln|s+2s2|},3D

where s=q/kF. (Note: Pre-factors for 1D and 2D χ0 were adjusted to standard forms; original 1D prefactor was missing vF, 2D prefactor 1/(2π) implies units where 2/m=1 or similar simplification. I've put in more standard forms with m,,vF for dimensional correctness, assuming parabolic bands. If the original expressions were specifically derived with certain simplifications, they might differ. The 3D form kF/(2π2) also implies 2/m=1 if compared to N(ϵF) forms, or means mkF/(2π22) if explicit. I have used the latter convention for the 3D prefactor and tried to make 1D/2D consistent, assuming the user is working in a system where m, are explicit).

Interestingly χ0(q,0) has a singularity at q=2kF in all dimensions:

Attachments/Script 33.webp|700

The singularity becomes weaker as the dimensionality is increased. In one dimension, there is a logarithmic divergence, in two dimensions there is a kink, and in three dimensions only the derivative diverges. Later we will see that these singularities may lead to instabilities of the metallic state, in particular for the one-dimensional case.


3.3 Lattice Vibrations - Phonons

The atoms in a lattice of a solid are not immobile but vibrate around their equilibrium positions. We will describe this new degree of freedom by treating the lattice as a continuous elastic medium (Jellium with elastic modulus λ). This approximation is sufficient to obtain some essential features of the interaction between lattice vibrations and electrons. In particular, renormalised screening effects will be found. Our approach here is, however, limited to monoatomic unit cells because the internal structure of a unit cell is neglected.

3.3.1 Vibration of an Isotropic Continuous Medium

The deformation of an elastic medium can be described by the displacement of the infinitesimal volume element d3r around a point r to a different point r(r). We can introduce here the so-called displacement field u(r)=r(r)r as function of r. In general, u is also a function of time. In the simplest form of an isotropic medium the elastic energy for small deformations is given by

Eel=λ2d3r(u(r,t))2

where λ is the elastic modulus (note that there is no deformation energy, if the medium is just shifted uniformly). This energy term produces a restoring force trying to bring the system back to the undeformed state. In this model we are neglecting the shear contributions. The continuum form above is valid for deformation wavelengths that are much longer than the lattice constant, so that details of the arrangement of atoms in the lattice can be neglected. The kinetic energy of the motion of the medium is given by

Ekin=ρ02d3r(u(r,t)t)2

where ρ0=Mini is the mass density with the ionic mass Mi and the ionic density ni. Variation of the Lagrangian functional L[u]=EkinEel with respect to u(r,t) leads to the equation of motion

1cs22t2u(r,t)(u(r,t))=0

which is a wave equation with sound velocity cs2=λ/ρ0. The resulting displacement field can be expanded into normal modes,

u(r,t)=1Ωkek(qk(t)eikr+qk(t)eikr)

where every qk(t) satisfies the equation

d2dt2qk+ωk2qk=0

with the frequency ωk=cs|k| and the polarisation vector ek has unit length. Note that within our simplification for the elastic energy defined previously, all modes correspond to longitudinal waves, that is ×u(r,t)=0 and ekk. The total energy expressed in terms of the normal modes reads

E=kρ0ωk2[qk(t)qk(t)+qk(t)qk(t)].

Next, we switch from a Lagrangian to a Hamiltonian description by defining the new variables

Qk=ρ0(qk+qk)Pk=ddtQk=iωkρ0(qkqk)

in terms of which the energy is given by

E=12k(Pk2+ωk2Qk2).

Thus, the system is equivalent to an ensemble of independent harmonic oscillators, one for each normal mode k. Consequently, the system may be quantised by defining the canonical conjugate operators PkP^k and QkQ^k which obey, by definition, the commutation relation,

[Q^k,P^k]=iδk,k.

As it is usually done for quantum harmonic oscillators, we define the raising and lowering operators

b^k=12ωk(ωkQ^k+iP^k)b^k=12ωk(ωkQ^kiP^k),

satisfying the commutation relations

[b^k,b^k]=δk,k[b^k,b^k]=0,[b^k,b^k]=0.

These relations can be interpreted in a way that these operators create and annihilate quasi-particles following the Bose-Einstein statistics. According to the correspondence principle, the quantum mechanical Hamiltonian corresponding to the previously defined energy is

H=kωk(b^kb^k+12).

In analogy to the treatment of the electrons in second quantisation we say that the operators b^k (b^k) create (annihilate) a phonon, a quasi-particle with well-defined energy-momentum relation, ωk=cs|k|. Using the previously defined relations for normal modes and operators the displacement field operator u^(r) can now be defined as

u^(r)=1Ωkek2ρ0ωk[b^keikr+b^keikr].

As mentioned above, the continuum approximation is valid for long wavelengths (small |k|) only. For wavevectors with |k|π/a the discreteness of the lattice appears in the form of corrections to the linear dispersion ωk|k|. Since the number of degrees of freedom is limited to 3Ni (Ni number of atoms), there is a maximal wave vector called the Debye wavevector kD. We can now define the corresponding Debye frequency ωD=cskD and the Debye temperature ΘD=ωD/kB. In the continuous medium approximation there are only acoustic phonons. For the inclusion of optical phonons, the arrangement of the atoms within a unit cell has to be considered, which goes beyond this simple picture.

3.3.2 Phonons in Metals

The consideration above is certainly valid for semiconductors, where ionic interactions are mediated via covalent chemical bonds and oscillations around the equilibrium position may be approximated by a harmonic potential, so that the form of the elastic energy above is well motivated. The situation is more subtle for metals, where the ions interact through the long-ranged Coulomb interaction and are held to together through an intricate interplay with the mobile conduction electrons.

First, neglecting the gluey effect of the electrons, the positively charged background can itself be treated as an ionic gas. Similar to the electronic plasma frequency derived earlier, the background exhibits a well-defined collective plasma excitation at the ionic plasma frequency

Ωp2=4πni(Zie)2Mi.

For the ionic plasma frequency we used the previously derived plasma frequency formula with n0ni=n0/Zi the density of ions with charge number Zi, eZie, and mMi the atomic mass. Apparently the excitation energy does not vanish as |k|0. So far, the background of the metallic system can not be described as an elastic medium where the excitation spectrum is expected to be linear in |k|,ωk|k|.

The shortcoming in this discussion is that we neglected the feedback effects of the electrons that react nearly instantaneously to the slow ionic motion, due to their much smaller mass. The finite plasma frequency is a consequence of the long-range nature of the Coulomb potential (as mentioned earlier), but as we have seen above the electrons tend to screen these potentials, in particular for small wavevectors k. The "bare" ionic plasma frequency Ωp is thus renormalised to

ωk2=Ωp2ε(k,0)=|k|2Ωp2|k|2+kTF2(cs|k|)2,

where the presence of the electrons leads to a renormalisation of the Coulomb potential by a factor 1/ε(k,ω). Having included the back-reaction of the electrons, a linear dispersion of a sound wave (ωk=cs|k|) is finally recovered, and the renormalised velocity of sound cs reads

cs2Ωp2kTF2=Zmωp2MikTF2=13ZmMivF2.

For the comparison of the energy scales we find,

ΘDTFcsvF=13ZmMi1,

where we used kBTF=ϵF and kDkF.

Kohn anomaly

Notice that phonon frequencies are much smaller than the (electronic) plasma frequency, so that the approximation

ωk2=Ωp2ε(k,0)

is valid even for larger wavevectors. Employing the Lindhard form of ε(k,0), we deduce that the phonon frequency is singular at |k|=2kF. More explicitly we find

ωk|k|

in the limit |k|2kF. This behaviour is called the Kohn anomaly and results from the interaction between electrons and phonons. This effect is not contained in the previous elastic medium model that neglected ion-electron interactions.

Attachments/Script 34.webp|700

3.3.3 Peierls Instability in one Dimension

The Kohn anomaly has particularly drastic effects in (quasi-)one-dimensional electron systems, where the electron-phonon coupling leads to an instability of the metallic state. We consider a one-dimensional Jellium model where the ionic background is treated as an elastic medium with a displacement field u along the extended direction (x-axis). We neglect both the electron-electron interaction and the slow time evolution of the background modulation so that the Hamiltonian reads,

H=Hisol+Hint

where contributions of the isolated electronic and ionic systems are included in

Hisol=k,s2k22mckscks+λ2dx(dudx(x))2

whereas the interactions between the system comes in via the coupling

Hint=n0sdxdxV(xx)ddxu(x)Ψ^s(x)Ψ^s(x).

In the general theory of elastic media u=δn/n0 describes density modulations, so that the interaction term couples electrons to charge density fluctuations of the positively charged background mediated by the screened Coulomb interaction V(xx) (contact interaction). Note that only phonon modes with a finite value of u couple in lowest order to the electrons. This is only possible of longitudinal modes. Transverse modes are defined by the condition u=0 and do not couple to electrons in lowest order.

Consider the ground state of N electrons in a system of length L, leading to an electronic density n0=N/L. For a uniform background u(x)=const, the Fermi wavevector of free electrons is readily determined to be

N=sL2πkF+kFdk1=2L2π2kF

leading to

kF=π2n0.

Perturbative approach - instability

Now we consider the Kohn anomaly of this one-dimensional system. Note that here we calculated concretely in one-dimension. In real quasi-one-dimensional systems only the electron dispersion would be one-dimensional - no perpendicular motion - while the overall calculations would done in three dimensions. For a small background modulation u(x)const, the interaction term Hint can be treated perturbatively and will lead to a renormalisation of the elastic modulus λ in the elastic energy term. For that it will be useful to express the full Hamiltonian in momentum space,

Hisol=k,s2k22mckscks+ρ02qωq2uquqHint=in02Lk,q,sq[V~quqc^k+q,sc^k,sV~quqc^k,sc^k+q,s]

where we used from previous considerations λq2=ρ0ωq2. Furthermore we defined

u(x)=1LquqeiqxV(x)=1LqV~qeiqx

with V~qconst. - we consider an effectively short-ranged potential. We compute the second order correction to the ground state energy using Rayleigh-Schrödinger perturbation theory (note that the linear energy shift vanishes)

ΔE(2)=n024Lk,q,sq2|V~q|2uquqn|Ψ0|c^k,sc^k+q,s|n|2+|Ψ0|c^k+q,sc^k,s|n|2E0En=n024qq2|V~q|2uquq1Lk,snk+qnkϵk+qϵk=n024q|V~q|2q2χ0(q,0)uquq

where the virtual states |n are electron-hole excitations of the filled Fermi sea. This term gives a correction to the elastic term in the momentum-space Hamiltonian for isolated systems. In other words, the elastic modulus λ and, thus, the phonon frequency ωq2=q2λ/ρ0=cs2q2 is renormalised according to

(ωqren)2ωq2+n02|V~q|2q24ρ0χ0(q,0)=ωq2n02|V~q|2mq4πρ02ln|q+2kFq2kF|.

From the behaviour for q0 we infer that the velocity of sound is renormalised. However, a much more drastic modification occurs at q=2kF. Here the phonon spectrum is 'softened', that is the frequency vanishes and even becomes negative. The latter effect is an artifact of the perturbation theory. This hints at an instability triggered by the Bose-Einstein condensation of phonons with a wave vector of 2kF. This coherent superposition of many phonons corresponds classically to a static periodic deformation of the ionic background with wave vector 2kF. The unphysical behaviour of the frequency indicates that in the vicinity of 2kF, the current problem can not be treated with the help of perturbation theory around the uniform state.

Peierls instability at Q=2kF

Instead of the perturbative approach, we assume that the background shows a periodic density modulation (coherent phonon state)

u(x)=u0cos(Qx)

where Q=2kF and u0 remains to be determined variationally. We investigate the effect of this modulation on the electron-phonon system. To this end we show that such a modulation lowers the energy of the electrons. Assuming that u0 is small we can evaluate the electronic energy using the approximation of nearly free electrons, where Q appears as a reciprocal lattice vector. The electronic spectrum for 0kQ is then approximately determined by the secular equation

det(2k22mEΔΔ2(kQ)22mE)=0

where Δ derives from the Fourier transform of the potential V(x),

Δ=iQu0nV~Q

with

V~Q=dxeiQxV(x).

The preceding secular equation leads to the energy eigenstates

Ek±=24m[(kQ)2+k2±{(kQ)2k2}2+16m2|Δ|2/4].

The total energy of the electronic and ionic system is then given by

Etot(u0)=20k<QEk+λLQ24u02

where all electronic states of the lower band (Ek) are occupied and all states of the upper band (Ek+) are empty. The amplitude u0 of the modulation is found by minimising Etot with respect to u0:

0=1LdEtotdu0=22m16Q2m2n2V~Q24u00Qdk2π1{(kQ)2k2}2+16m2Q2n2V~Q2u02/4+λ2Q2u0=u02Qmn2V~Q22πkF+kFdq1q2+4m2n2V~Q2u02/4+λ2Q2u0=u04Qmn2V~Q22πarsinh(2kF2mnV~Qu0)+λ2Q2u0.

We solve this equation for u0 using arsinh(x)ln(2x) when x1.

u0=2kFmnV~Qexp[2kFπλ4mn2V~Q2]=2kFϵFnV~Qe1/N(ϵF)g

where ϵF=2kF2/2m is the Fermi energy and N(ϵF)=2m/(π2kF) is the density of states at the Fermi energy in the one-dimensional system. We introduced the coupling constant g=2n2V~Q2/λ that describes the phonon-induced effective electron-electron interaction. The coupling is the stronger the more polarisable (softer) ionic background, that is when the elastic modulus λ is small. Note that the static displacement u0 depends exponentially on the coupling and on the density of states. The underlying reason for this so-called Peierls instability to happen lies in the opening of an energy gap,

ΔE=EkF+EkF=2|Δ|=8ϵFexp(1N(ϵF)g)

at k=±kF, that is at the Fermi energy. The gap is associated with a lowering of the energy of the electron states in the lower band in the vicinity of the Fermi energy. For this reason this kind of instability is called a Fermi surface instability. Due to the gap the metal has turned into a semiconductor with a finite energy gap for all electron-hole excitations.

Attachments/Script 35.webp|700

The modulation of the electron density follows the charge modulation due to the ionic lattice deformation, which can be seen by expressing the wave function of the electronic states,

ψk(x)=1LΔeikx+(Ekϵk)ei(kQ)x(Ekϵk)2+|Δ|2

which is a superposition of two plane waves with wave vectors k and kQ, respectively. Hence the charge density reads

ρk(x)=e|ψk(x)|2=eL[12(ϵkEk)|Δ|(Ekϵk)2+|Δ|2sinQx]

and its modulation from the homogeneous distribution en is given by

δρ(x)=2kρk(x)(en)=2e0kFdk2πm|Δ|sinQx4kF2k2+m2|Δ|2=en|Δ|4ϵFln|4ϵF|Δ||sin(2kFx).

Such a state, with a spatially modulated electronic charge density, is called a charge density wave (CDW) state. This instability is important in quasi-one-dimensional metals which are, for example, realised in organic conductors such as TTF-TCNQ (tetrathiafulvalene tetracyanoquinomethane). In higher dimensions the effect of the Kohn anomaly is generally less pronounced, so that in this case spontaneous deformations rarely occur. As we will see later, a charge density wave instability can nevertheless be observed in multi-dimensional (d>1) systems with a so-called nested Fermi surface. These systems resemble in some respects one-dimensional systems. Finally, notice that the electron-phonon interaction strongly contributes to another kind of Fermi surface instability, when metals exhibit superconductivity.

3.3.4 Dynamics of Phonons and the Dielectric Function

We have seen that an external potential Va is screened by the polarisation of the electrons. As the positively charged ionic background is also polarisable, it should be included in the renormalisation of the external potential. In general, the fully renormalised potential Vren may be expressed via

εVren=Va

with the full dielectric function ε. In order to determine Vren and ε, we define the 'bare' (unrenormalised) electronic (ionic) dielectric function εel(εion). The renormalised potential can be expressed considering three other points of view. First, if the ionic potential Vion is added to the external potential Va, the remaining screening is due to the electrons only, that is,

εelVren=Va+Vion.

Secondly, the electronic potential Vel may be added to the external potential Va, so that the ions exclusively renormalise the new potential Vel+Va, resulting in

εionVren=Va+Vel.

Note that all effects of electron polarisation are included in Vel, so that the dielectric function results from the 'bare' ions. Finally we use the fact that Vren may be expressed as

Vren=Va+Vel+Vion.

We obtain

(εel+εionε)Vren=Va+Vel+Vion

which simplifies with the definition Vren=Va+Vel+Vion to

ε=εel+εion1.

In order to find an alternative expression relating the renormalised potential Vren to the external potential Va, we make the Ansatz

Vren=1εVa=1εeffion1εelVa

that is the potential Va/εel that results from bare screening of the polarisable electrons is additionally screened by an effective ionic dielectric function εeffion which includes electron-phonon interactions. Using the relation ε=εel+εion1 and the definition of εeffion we obtain

εeffion=1+1εel(εion1),

or using ε=1Vχ0 which implies (1ε)/V=χ0, we can write

χ0,effion=χ0ionεel.

Taking into account the discussion of the plasma excitation of the bare ions, and considering the long wave-length excitations (k0), we approximate (with k=|k|)

εion=1Ωp2ω2εel=1+kTF2k2.

For the electrons we used the result from the quasi-static limit derived for Thomas-Fermi screening. The full dielectric function now reads (with k=|k|)

ε=1+kTF2k2Ωp2ω2=(1+kTF2k2)(1ωk2ω2).

The time-independent Coulomb interaction (with q=|q|)

Va(q)=4πe2q2

between the electrons is replaced in a metal by an effective interaction (with q=|q|)

Vren(q,ω)=4πe2q2ε(q,ω)=4πe2kTF2+q2(ω2ω2ωq2).

This interaction corresponds to the matrix element for a scattering process of two electrons with momentum exchange q and energy exchange ω. The phonon frequency ωq is always less than the Debye frequency ωD. Hence the effect of the phonons is almost irrelevant for energy exchanges ω that are much larger than ωD. The time scale for such energies would be too short for the slow ions to move and influence the interaction. Interestingly, the repulsive bare Coulomb potential is renormalised to an interaction with an attractive channel for ω<ωD because of overcompensation by the ions. This aspect of the electron-phonon interaction is most important for superconductivity.

Attachments/Script 36.webp|700


3.4 Appendix: Linear Response Theory

Much information about a macroscopic system can be gained through the observation of its response to a small external perturbation. If the perturbation is sufficiently small we can consider the response of the system in lowest order, linear in the perturbing field. This is the approach pursued by the linear response theory.
If we knew all stationary states of a macroscopic quantum system with many degrees of freedom we could calculate essentially any desired quantity. However, the full information of such system is hard to store and is also unnecessary in view of our real experimental interests. The linear response functions are an efficient way to provide in a condensed form some of the most important and relevant information accessible in an experiment. The linear response function is one element of quantum field theory of solid-state physics. We will introduce it here on an elementary level.

3.4.1 Linear Response Function

Previously, we analysed the influence of an external potential on the distribution of electrons. The position- and time-dependent potential V(r,t) induces the density modulation δn(r,t) which would be measured as a response to the potential. In real space and time the connection between the two is given by the linear relation,

δn(r,t)=d3rdtχ(rr,tt)V(r,t)

where we assume that the system is homogeneous and isotropic. The response function χ(rr,tt) describes how the potential at a position r and time t influences the behaviour of the system at the position r and at a later time t (causality). Causality actually requires that χ(r,t)=0 for t<0. The response functions are non-local in space and time. The convolution can be converted into a simple product by going to momentum-frequency space,

δn(q,ω)=χ(q,ω)V(q,ω)

where the Fourier transformation is performed as follows,

f(r,t)=1Ωq+dω2πf(q,ω)ei(ωtqr).

We now determine the response function for a general external field and response quantity.

Kubo formula - retarded Green's function

We consider a quantum system described by the Hamiltonian H0 and analyze its response to an external field ϕ(r,t) which couples to the field operator A^(r),

H=H0+H(t)=H0+d3rA^(r)ϕ(r,t)eηt

where η=0+is a small positive parameter allowing to switch the perturbation adiabatically on, that is, at time t there is no perturbation. The behaviour of the system can be determined by the density matrix ρ^(t). Knowing ρ^ we can calculate interesting mean values of operators, B^(t)=tr(ρ^(t)B^). A way to find the density matrix goes via the equation of motion,

iρ^t=[ρ^,H]=[ρ^,H0+H].

We proceed using the concept of time-dependent perturbation theory, ρ^=ρ^0+δρ^(t), with

ρ^0=1ZeβH0 with Z=treβH0

as the partition function and insert this form into the equation of motion, which we then truncate at linear order in H,

itδρ^=[δρ^,H0][ρ^0,H]+.

We turn to the interaction representation (time-dependent perturbation theory),

δρ^(t)=eiH0t/y^(t)eiH0t/itδρ^=[δρ^,H0]+eiH0t/{iy^(t)t}eiH0t/.

Comparing the preceding equations for the density matrix and using the definition of the Hamiltonian we arrive at the equation for y^,

iy^(t)t=[ρ^0,Hint(t)] with Hint(t)=eiH0t/HeiH0t/

which is formally solved by

y^(t)=itdt[ρ^0,Hint(t)].

We now look at the mean value of the observable B^(r). For simplicity we assume that the expectation value of B^ vanishes, if there is no perturbation, that is B^0=tr{ρ^0B^}=0. We determine

B^(r)(t)=tr{δρ^(r,t)B^(r)}=tr{ieiH0t/tdt[ρ^0,Hint(t)]eiH0t/B^(r)}.

By means of cyclic permutation of the operators in {}, which does not affect the trace, we arrive at the form

B^(r)(t)=itdtd3rtr{ρ^0[B^int(r,t),A^int(r,t)]}ϕ(r,t)eηt=dtd3rχBA(rr,tt)ϕ(r,t)

which defines the response function χBA. Notably, it is entirely determined by the properties of the unperturbed system.

Kubo-Formula - Recipe for the linear response function: We arrive at the following recipe to obtain a general linear response function: We denote the Hamiltonian of the (unperturbed) system as H0. Then the linear response function of the pair of field operators A^(r),B^(r) (they are often in practice conjugate operators, A^=B^) is given by

χBA(rr,tt)=iΘ(tt)[B^H(r,t),A^H(r,t)]H0

where H0 is the thermal mean value with respect to the Hamiltonian H0,

C^H0=tr{C^eβH0}tr{eβH0}=tr{C^eβH0}Z,

A^H(t)=eiH0t/A^eiH0t/ is the Heisenberg representation of the operator A^ (analogous for B^). Note that the temporal step function Θ(tt) ensures the causality, that is there is no response for the system before there is a perturbation. This form is often called Kubo formula or retarded Green's function.

3.4.2 Information in the Response Function

The information stored in the response function can be most easily visualised by assuming that we know the complete set of stationary states of the system Hamiltonian H0:H0|n=ϵn|n. For simplicity we will from now on assume that A^=B^ which is the case in many practical examples, and will simplify our notation. We can then rewrite the response function χ as

χ(rr,tt)=iΘ(tt)neβϵnZ{n|eiH0t/B^(r)eiH0t/eiH0t/B^(r)eiH0t/|nn|eiH0t/B^(r)eiH0t/eiH0t/B^(r)eiH0t/|n}=iΘ(tt)n,neβϵnZ{n|B^(r)|nn|B^(r)|nei(ϵnϵn)(tt)/n|B^(r)|nn|B^(r)|nei(ϵnϵn)(tt)/},

where we inserted 1=n|nn|. It is convenient to work in q- and ω-space, reached via Fourier transformation,

χ(q,ω)=d3r~+dt~χ(r~,t~)eiωt~iqr~=iΩn,neβϵnZ|n|B^q|n|20dt~{ei(ϵnϵn+ω)t~/ei(ϵnϵn+ω)t~/}eηt~

where we introduce

B^q=d3r~B^(r~)eiqr~ and B^q=d3r~B^(r~)eiqr~.

After the integration over time we obtain

χ(q,ω)=1Ωn,neβϵnZ|n|B^q|n|2{1ωϵn+ϵn+iη1ωϵn+ϵn+iη}=0dωS(q,ω){1ωω+iη1ω+ω+iη}.

In the last line we write the response function in a spectral form with S(q,ω) as the spectral function,

S(q,ω)=1Ωn,neβϵnZ|n|B^q|n|2δ(ωϵn+ϵn).

We call S(q,ω) also dynamical structure factor which comprises information about the excitation spectrum associated with B^. It represents a correlation function,

S(rr,tt)=1B^H(r,t)B^H(r,t)H0,

and contains the spectrum of the excitations which can be coupled to by the external perturbation.

Lindhard function

We show now how the Lindhard function is derived, where we restrict to independent electrons. Thus, we choose, B^(r)=sΨ^s(r)Ψ^s(r) which leads to

B^q=ρ^q=k,sc^ksc^k+q,s=k,sρ^k,q,s.

Now consider the matrix elements in

n,neβϵnZn|c^ksc^k+q,s|nn|c^k+q,sc^ks|n=n0k(1n0k+q)δk,k

for which

ϵn=ϵnϵk+ϵk+qϵnϵn=ϵk+qϵk.

The matrix element probes probability of whether the electron state with momentum k is occupied, by nk, and the state k+q, by 1nk+q, is empty. Thus we obtain,

χ(q,ω)=1Ωkn0k(1n0k+q){1ωϵk+q+ϵk+iη1ω+ϵk+qϵk+iη}.

The second term can be rewritten by substitution kkq and in a next step switch the sign of q which leaves the result invariant, as our system has inversion symmetry. This leads then to

χ(q,ω)=1Ωkn0k(1n0k+q)n0k+q(1n0k)ωϵk+q+ϵk+iη=1Ωkn0k+qn0kϵk+qϵkωiη,

which is consistent with the Lindhard function defined previously. The Kubo formalism can be used to derive intensive response functions like the Lindhard function when appropriate normalization by volume is included.

3.4.3 Analytical Properties

The spectral representation of the linear response function shows that χBA(q,ω) has poles only in the lower half of the complex ω-plane. This property reflects causality (χ(r,t)=0 for t<0). We separate now χ=χ+iχ in real and imaginary part and use the relation

limη0+1x+iη=P1xiπδ(x).

with P denoting the principal part. This relation leads to

χ(q,ω)=0dωS(q,ω){P1ωωP1ω+ω},χ(q,ω)=π{S(q,ω)S(q,ω)}.

Therefore the imaginary part of χ corresponds to the excitation spectrum of the system. Finally, it has to be noted that χ(q,ω) follows the Kramers-Kronig relations:

χ(q,ω)=1π+dωPχ(q,ω)ωωχ(q,ω)=1π+dωPχ(q,ω)ωω.

Note to the Kramers-Kronig relation: This relation results from the analytic structure of χ. Consider a contour in the upper half-plane of ω where χ(q,ω) has no poles due to causality.

Cdωχ(q,ω)ωωiη=0+dωχ(q,ω)P1ωω+iπχ(q,ω)=0.

Separating this equation into real and imaginary part yields the Kramers-Kronig relation.

Attachments/Script 37.webp|700

3.4.4 Fluctuation-Dissipation Theorem

First we consider the aspect of dissipation incorporated in the response function. For this purpose we ignore for simplicity the spatial dependence and consider a perturbative part of the Hamiltonian which only depends on time.

H=d3rB^(r)ϕ(r,t)B^(r)(t)=0dtd3rχ(rr,tt)ϕ(r,t)

with B^=B^. We assume now a monochromatic external field,

ϕ(r,t)=12(ϕ0(r)eiωt+ϕ0(r)eiωt)B^(r)(t)=0dtd3rχ(rr,tt)12(ϕ0(r)eiωt+ϕ0(r)eiωt)=12d3r{χ(rr,ω)ϕ0(r)eiωt+χ(rr,ω)ϕ0(r)eiωt}.

The energy dissipation rate is determined by

dEdt=tH=d3rB^(r)(t)ϕ(r,t)t=12d3rd3rϕ(r,t)t{χ(rr,ω)ϕ0(r)eiωt+χ(rr,ω)ϕ0(r)eiωt}=12d3rd3riω{ϕ0(r)eiωt+ϕ0(r)eiωt}{χ(rr,ω)ϕ0(r)eiωt+χ(rr,ω)ϕ0(r)eiωt}dEdt=ω2d3rd3rϕ0(r)χ(rr,ω)ϕ0(r)=qω2χ(q,ω)|ϕ(q,ω)|2<0

where for the time averaged rate we drop oscillating terms with the time dependence e±i2ωt.
The result before used the following: The time-derivative of the Hamiltonian is given by

dHdt=Ht+i[H,H]=Ht

for a quantum mechanical problem. The analogous relation is obtained for classical systems.

The imaginary part of the dynamical susceptibility describes the dissipation of the system and gives information about the spectrum and the density of excitations for given q and ω. From the definition of the dynamical structure factor it follows that

S(q,ω)=eβωS(q,ω)

because

S(q,ω)=1Ωn,neβϵnZ|n|B^q|n|2δ(ωϵn+ϵn)=1Ωn,neβϵnβωZ|n|B^q|n|2δ(ωϵn+ϵn)=eβωS(q,ω).

This is a statement of detailed balance. The transition matrix element between two states is the same whether the energy is absorbed or emitted. For emitting, however, the thermal occupation of the initial state has to be taken into account.
Using the relation S(q,ω)=eβωS(q,ω) we can derive the following relation

χ(q,ω)=π[S(q,ω)S(q,ω)]=π[1eβω]S(q,ω)

which is known as the fluctuation-dissipation theorem. Let us consider here some consequences and find the relation to our earlier simplified formulations.

+dωS(q,ω)=+dω1Ωn,neβϵnZ|n|B^q|n|2δ(ωϵn+ϵn)=1Ωn,neβϵnZn|B^q|nn|B^q|n=1B^q(0)B^q(0)=1π+dωχ(q,ω)1eβω

This corresponds to the equal-time correlation function (assuming B^=0).

Now we turn to the classical case of the fluctuation-dissipation theorem and consider the limit kBTω. Then we may approximate this equation by

|B^q|2kBTπ+dωχ(q,ω)ω=kBTχ(q,0)=kBTχ(q,0).

This is valid, if χ(q,ω) essentially vanishes for frequencies comparable and larger than the temperature.
Static response function: We consider a system with

H=H0+H=H0+d3rh(r)B^(r)=H0+1VqhqB^q=H0+qHqB^q

where we assume for the following B^q=B^q. The mean value

B^q=FHq=HqkBTlnZ=tr(ρ^B^q)

with ρ^=exp[β(FH)] and B^q=0 for Hq=0. The static response function is obtain from

χ(q)=B^qHq|Hq=0=tr{B^qHqeβ(FH)}|Hq=0=tr{B^qβ(B^qB^q)ρ^}|Hq=0=βB^qB^q

which is the classical form of the fluctuation dissipation theorem for spatially modulated perturbative fields.
For a uniform field we find

d3rd3rB^(r,t=0)B^(r,t=0)=B^q=0B^q=0=kBTχ(q=0)=kBTχ

that is the static uniform susceptibility is related to the integration of the equal-time correlation function as we had used previously several times. Note the minus sign results from the sign of coupling to the external field.