Jump back to chapter selection.


Table of Contents

5.1 Lifetime of Quasi-Particles
5.2 Phenomenological Theory of Fermi Liquids
5.3 Microscopic Considerations


5 Landau's Theory of Fermi Liquids

In the previous chapters, we considered the electrons of the system as more or less independent particles. The effect of their mutual interactions only entered via the renormalisation of potentials and in the emergence of collective excitations. The underlying assumptions of our earlier discussions were that electrons in the presence of interactions can still be described as Fermionic particles with a well-defined energy-momentum relation, and that their ground state is a filled Fermi sea with a sharp Fermi surface. Since there is no guarantee that this hypothesis holds in general (and indeed, this is not always true), we have to show that in metals the description of electrons as quasi-particles is justified. This quasi-particle picture will lead us to Landau's phenomenological theory of Fermi liquids. Indeed, it is very surprising that a strongly interacting many-electron system does not end up in an extremely complex quantum state. What saves us are two important points:

  1. In a metal the long-ranged Coulomb interaction is screened and becomes short-ranged;
  2. The Pauli principle reduces the phase space of low-energy excitations of electrons dramatically, at least in a three-dimensional system (quantum protection).

5.1 Lifetime of Quasi-Particles

We first consider the lifetime of a state consisting of a filled Fermi sea to which one electron is added. Let k with |k|>kF (ϵk=2|k|2/(2m) with ϵk>ϵF) be the momentum (energy) of the additional electron. Due to interactions between the electrons, this state will decay into a many-body state. In momentum space such an interaction takes the form

Hee=k,k,qs,sV(q)c^kq,sc^k+q,sc^k,sc^k,s

where V(q) represents the electron-electron interaction in momentum space while q indicates the momentum transfer in the scattering process. Below, the short-ranged Yukawa potential

V(q)=4πe2|q|2ε(q,0)=4πe2|q|2+kTF2

derived in the discussion of Thomas-Fermi screening will be used. As we are only interested in very small energy transfers ωϵF, the static approximation is admissible.
In a perturbative treatment, the lowest order effect of the interaction is the creation of a particle-hole excitation in addition to the single electron above the Fermi energy. As the additional electron changes its momentum from k to kq, a hole appears at k and a second electron with wavevector k+q is created outside the Fermi sea. The transition is allowed whenever both energy and momentum are conserved. Momentum conservation for the process k+k(kq)+(k+q) is trivially satisfied. Energy conservation requires

ϵk+ϵk=ϵkq+ϵk+q.

Alternatively, for the decay of the particle k, its energy loss must equal the energy of the created particle-hole pair: ϵkϵkq=ϵk+qϵk.

We calculate the lifetime τk of the initial state with momentum k using Fermi's golden rule, yielding the transition rate from the initial state of a filled Fermi sea and one particle with momentum k to a state with two electrons above the Fermi sea, with momenta kq and k+q, and a hole with k:

Attachments/Script 51.webp|700

Since neither the momenta k and q, nor the spin of the created electron are fixed, a summation over the possible configuration has to be performed, leading to

1τk=2π1Ω2k,qs|V(q)|2n0,k(1n0,kq)(1n0,k+q)δ(ϵkq+ϵk+qϵkϵk).

Note that the term n0,k(1n0,kq)(1n0,k+q) takes care of the Pauli principle, by ensuring that the final state after the scattering process exists, that is the hole state k lies inside and the two particle states kq and k+q lie outside the Fermi sea. First the integral over k is performed under the condition that the energy ϵk+qϵk of the excitation is small. With that, the integral reduces to

S(ωq,k,q)=1Ωkn0,k(1n0,k+q)δ(ϵk+qϵkωq,k)=1(2π)3d3kn0,k(1n0,k+q)δ(ϵk+qϵkωq,k)=N(ϵF)4ωq,kqvF

where N(ϵF)=mkF/(π22) is the density of states (for one spin direction) of the electrons at the Fermi surface and ωq,k=ϵkϵkq=2(2kq|q|2)/(2m)>0 is the energy loss of the decaying electron. Here q=|q|. Small ωq,k are justified, because ωq,k2(2kFqq2)/(2m) for most allowed ωq,k. The integral may be computed using cylindrical coordinates, where the vector q points along the axis of the cylinder. It results in

S(q,ω)=1(2π)2k2k1dkk0kFdkδ(2q22m+2qkmω)=m4π22q(k12k22)

with k12=kF2k,02 and k22=kF2(k,0+q)2, where k,0=(2mωq2)/(2q) is enforced by the delta function.

Attachments/Script 52.webp|700

The wave vectors k2 and k1 are the upper and lower limits of integration determined from the condition n0,k(1n0,k+q)>0 and can be obtained by simple geometric considerations. The previous expression for S(ωq,k,q) follows immediately.

In order to compute the remaining integral over q, we assume that the matrix element |V(q)|2 depends only weakly on q when |q|kF. This is especially true when the interaction is short-ranged. In spherical coordinates, the integral reads (with k=|k|,q=|q|)

1τk=2πN(ϵF)4vFΩq,s|V(q)|2ωq,kq=N(ϵF)(2π)22vFd3q|V(q)|2ωq,kq=N(ϵF)(2π)4mvFdq|V(q)|2q2θ1θ2dθsinθ(2kcosθq)=N(ϵF)(2π)4mvFdq|V(q)|2q2[14k(2kcosθq)2]θ1θ2.

The restriction of the domain of integration of θ follows from the two conditions |k|2|kq|2kF2 and |kq|2=k22kqcosθ+q2. From the first condition, cosθ2=q/(2k), and from the second, cosθ1=(k2kF2+q2)/(2kq). Thus,

1τk=N(ϵF)(2π)4mvFdq|V(q)|214k(k2kF2)2N(ϵF)(2π)4vFmkF14(ϵkϵF)2dq|V(q)|2=18π3N(ϵF)vF2(ϵkϵF)2dq|V(q)|2.

Note that convergence of the last integral over q requires that the integrand does not diverge stronger than qα with α<1 for q0. This condition is fulfilled for the screened Yukawa potential, but would not be for the unscreened Coulomb potential. Essentially, the result states that

1τk(ϵkϵF)2

for k slightly above the Fermi surface. This implies that the state |ks occurs as a resonance of width /τk and features a quasi-particle, which can be observed in the spectral function A(E,k) as depicted here:

Attachments/Script 53.webp|700

The quasi-particle (coherent) part of the spectral function has a weight reduced from one (corresponding to the quasi-particle weight Zk). The remaining weight is shifted to higher energies as a so-called incoherent part (continuum without clear momentum-energy relation). The resonance becomes arbitrarily sharp as the Fermi surface is approached

lim|k|kF/τkϵkϵF=0

so that the quasi-particle concept is asymptotically valid. This equation can also be seen as a verification of Heisenberg's uncertainty principle. Consequently, the momentum of an electron is a good quantum number in the vicinity of the Fermi surface. Underlying this result is the Pauli exclusion principle, which restricts the phase space for decay processes of single particle states close to the Fermi surface. In addition, the assumption of short ranged interactions is crucial. Long ranged interactions can change the behaviour drastically due to the larger number of decay channels.


5.2 Phenomenological Theory of Fermi Liquids

The existence of well-defined Fermionic quasi-particles in spite of the underlying complex many-body physics inspired Landau to formulate the following phenomenological theory. Just like the states of independent electrons, quasi-particle states shall be characterised by their momentum k and spin σ. In fact, there is a one-to-one mapping between the free electrons and the quasi-particles. Consequently, the number of quasi-particles and the number of electrons coincide. The momentum distribution function of quasi-particles, defined as nσ(k), is subject to the condition

N=k,σnσ(k).

In analogy to the Fermi-Dirac distribution of free electrons, one demands, that the ground state distribution function nσ(0)(k) for the quasi-particles is described by a simple step function

nσ(0)(k)=Θ(kF|k|).

For a spherically symmetric electron system, the quasi-particle Fermi surface is a sphere with the same radius as the one for free electrons of the same density. For a general point group symmetry, the Fermi surface may be deformed by the interactions without changing the underlying symmetry. The volume enclosed by the Fermi surface is always conserved despite the deformation. Note that the distribution nσ(0)(k) of the quasi-particles in the ground state and that n0kσ=c^kσc^kσ of the real electrons in the ground state are not identical:

Attachments/Script 54.webp|700

Interestingly, n0kσ is still discontinuous at the Fermi surface, but the height of the jump is, in general, smaller than unity. The modification of the electron distribution function from a step function to a "smoother" Fermi surface indicates the involvement of electron-hole excitations and the renormalisation of the electronic properties, which deplete the Fermi sea and populate the states above the Fermi level. The reduced jump in n0kσ is a measure for the quasi-particle weight at the Fermi surface, ZkF, that is, the amplitude of the corresponding free electron state in the quasi-particle state.
In Landau's theory of Fermi liquids, the essential information on the low-energy physics of the system shall be contained in the deviation of the quasi-particle distribution nσ(k) from its ground state distribution nσ(0)(k),

δnσ(k)=nσ(k)nσ(0)(k).

The symbol δ is generally used in literature to denote this difference. Unfortunately this may suggest that the term δnσ(k) is small, which is not true in general. Indeed, δnσ(k) is concentrated on momenta k very close to the Fermi surface only, where the quasi-particle concept is valid. This distribution function, describing the deviation from the ground state, enters a phenomenological energy functional of the form

E=E0+k,σϵσ(k)δnσ(k)+12Ωk,kσ,σfσσ(k,k)δnσ(k)δnσ(k)+O(δn3)

where E0 denotes the energy of the ground state. Moreover, the phenomenological parameters ϵσ(k) and fσσ(k,k) have to be determined by experiments or by means of a microscopic theory. The variational derivative

ϵ~σ(k)=δEδnσ(k)=ϵσ(k)+1Ωk,σfσσ(k,k)δnσ(k)

yields an effective energy-momentum relation ϵ~σ(k), whose second term depends on the distribution of all quasi-particles. A quasi-particle moves in the "mean-field" of all other quasi-particles, so that changes δnσ(k) in the distribution affect ϵ~σ(k). The second variational derivative describes the coupling between the quasi-particles

δ2Eδnσ(k)δnσ(k)=1Ωfσσ(k,k).

We introduce a parametrisation for these couplings fσσ(k,k) by assuming spherical symmetry of the system. Furthermore, the radial dependence is ignored, as we only consider quasi-particles in the vicinity of the Fermi surface where |k|,|k|kF. Therefore the dependence of fσσ(k,k) on k,k can be reduced to the relative angle θk^,k^

fσσ(k,k)=fs(k^,k^)+σσfa(k^,k^)

where k^=k/|k|. The symmetric (s) and antisymmetric (a) part of fσσ(k,k) can be expanded in Legendre-polynomials Pl(z), leading to

fs,a(k^,k^)=l=0fls,aPl(cosθk^,k^).

The density of states at the Fermi surface is defined as (for both spins, per unit volume)

N(ϵF)=2Ωkδ(ϵ(k)ϵF)=kF2π2vF=mkFπ22

and follows from the dispersion ϵ(k) of the bare quasi-particle energy

kϵ(k)|k=kF=vF=kFm

where for a fully rotation symmetric system we may write ϵ(k)=2|k|2/(2m) with m as an "effective mass", although we will be only interested at the spectrum in the immediate vicinity of the Fermi energy. With this definition, we also introduce the so-called Landau parameters

Fls=N(ϵF)fls,Fla=N(ϵF)fla,

commonly used in the literature. In the following, we want to study the relation between the different phenomenological parameters of Landau's theory of Fermi liquids and the experimentally accessible quantities of a real system, such as specific heat, compressibility, spin susceptibility among others.

5.2.1 Specific Heat - Density of States

Since the quasi-particles are Fermions, they obey Fermi-Dirac statistics

nσ(T,k)=1e[ϵ~(k)μ]/kBT+1

with the chemical potential μ, leading to

δnσ(k)=nσ(T,k)nσ(0)(0,k).

We will only consider here the behaviour in lowest-order in temperature and restrict ourselves, therefore, to ϵ~(k)=ϵ(k). Furthermore, we replaced μ=ϵF+O(T2) by ϵF. In order to determine the specific heat, we use the expression for the entropy of a Fermion gas based on the momentum distribution function,

S=kBΩk,σ[nσ(k,T)ln(nσ(k,T))+(1nσ(k,T))ln(1nσ(k,T))].

Taking the derivative of the entropy S with respect to T, the specific heat

C(T)=TST=kBTΩk,σeξ(k)/kBT(eξ(k)/kBT+1)2ξ(k)kBT2ln(nσ(k,T)1nσ(k,T))=kBTΩk,σ14cosh2(ξ(k)/(2kBT))ξ(k)kBT2ξ(k)kBT

is obtained, where we introduced ξ(k)=ϵ(k)ϵF. In the limit T0 we find

C(T)TN(ϵF)4kBT3dξξ2cosh2(ξ/(2kBT))kB2N(ϵF)4+dyy2cosh2(y/2)=π2kB2N(ϵF)3

which is the well-known linear behaviour C(T)=γT for the specific heat at low temperatures, with γ=π2kB2N(ϵF)/3. Since N(ϵF)=mkF/(π22) (total DOS per unit volume), the effective mass m of the quasi-particles can directly determined by measuring the specific heat of the system.

5.2.2 Compressibility - Landau parameter F0s

A Fermi gas has a finite compressibility because each Fermion occupies a finite amount of space due to the Pauli principle. The compressibility κ is defined as

κ=1Ω(Ωp)T,N

where p is the uniform hydrostatic pressure. The indices T,N mean, that the temperature T and the particle number N are kept fixed. We consider the response of the Fermi liquid upon application of uniform pressure p. The shift of the bare quasi-particle energies is given by

δϵσ(k)=ϵσ(k)pδp=ϵ(k)kkΩΩpδp=κ(0)3vkkδp=γkκ(0)δp

with n=N/Ω. We use the fact that

k=2πLn=2πnΩ1/3kΩ=13Ω2πnΩ1/3=k3Ω

and denote γk=vkk/3=2ϵσ(k)/3 and κ(0) is the unrenormalised compressibility derived below. Analogously we introduce the shift of the renormalised quasi-particle energies with the renormalised compressibility κ,

δϵ~σ(k)=γkκδp=δϵσ(k)+1Ωk,σfσ,σ(k,k)δnσ(k)=γkκ(0)δp+1Ωk,σfσ,σ(k,k)nσ(k)ϵ~σ(k)δϵ~σ(k)=γkκ(0)δp1Ωk,σfσ,σ(k,k)δ(ϵ~σ(k)ϵF)γkκδp.

Changes are concentrated on the Fermi surface such that we can replace γk2ϵF/3 so that

κ=κ(0)κN(ϵF)dΩk^4πfs(k^,k^)=κ(0)κF0s.

Therefore we find

κ=κ(0)1+F0s.

Now we determine κ(0) from the volume dependence of the energy

E(0)=k,σϵσ(k)=35NϵF=35N2kF22m=3102Nm(3π2NΩ)2/3.

Then we determine the pressure

p=(E(0)Ω)N=152Nm(3π2NΩ)2/31Ω

and

1κ(0)=ΩpΩ=132NmΩ(3π2n)2/3=23nϵF.

5.2.3 Spin Susceptibility - Landau Parameter F0

In a magnetic field H coupling to the electron spins the bare quasi-particle energy is supplemented by the Zeeman term,

ϵσ(k)=2|k|22mgμBHσ2

where σ=±1 denotes the spin component parallel to the applied field. The shift of the renormalised quasi-particle energy due to the applied field is

δϵ~σ(k)=ϵ~σ(H,k)ϵ~(H=0,k)=gμBHσ2+1Ωk,σfσ,σ(k,k)δnσ(k)=g~μBHσ2.

Attachments/Script 55.webp|700

Note that by symmetry, δnσ(k)=δnσ(k). Due to interactions, the renormalised gyromagnetic factor g~ differs from the value of g2 for free electrons. We focus on the second term, which can be expressed as

1Ωk,σfσσ(k,k)δnσ(k)=1Ωk,σfσσ(k,k)nσ(0)(k)ϵ~σ(k)δϵ~σ(k)=1Ωk,σfσσ(k,k)(δ(ϵ~σ(k)ϵF))(g~μBHσ2).

We derive

g~=gg~N(ϵF)dΩk^4πfa(k^,k^)=gg~F0a

or equivalently

g~=g1+F0a.

The magnetisation of the system can be computed from the distribution function,

M=gμBk,σσ2δnσ(k)=gμBk,σσ2nσ(0)(k)ϵ~σ(k)δϵ~σ(k)=gμBk,σσ2(δ(ϵ~σ(k)ϵF))(g~μBHσ2)

from which the susceptibility is immediately found to be (assuming g2)

χ=MHΩ=μB2N(ϵF)1+F0a.

The changes in the distribution function induced by the magnetic field feed back into the susceptibility, so that the latter may be either weakened (F0a>0) or enhanced (F0a<0). For the magnetic susceptibility, the Landau parameter F0a and the effective mass m (through N(ϵF)) lead to a renormalisation compared to the free electron susceptibility.

5.2.4 Galilei Invariance - Effective Mass and F1s

We initially introduced by hand the effective mass of quasi-particles in ϵσ(k). In this section we show that overall consistency of the phenomenological theory requires a relation between the effective mass and one Landau parameter (F1s). The reason is that the effective mass is the result of the interactions among the electrons. This self-consistency is connected with the Galilean invariance of the system. When the momenta of all particles are shifted by q (|q| shall be very small compared to the Fermi momentum kF in order to remain within the assumption-range of the Fermi liquid theory) the distribution function given by

δnσ(k)=nσ(0)(kq)nσ(0)(k)qknσ(0)(k).

This function is strongly concentrated around the Fermi energy:

Attachments/Script 56.webp|700

The current density can now be calculated, using the distribution function nσ(k)=nσ(0)(k)+δnσ(k). Within the Fermi liquid theory this yields,

jq=1Ωk,σv(k)nσ(k)=1Ωk,σv(k)δnσ(k)

with

v(k)=1kϵ~σ(k)=1(kϵσ(k)+1Ωk,σkfσσ(k,k)δnσ(k)).

Thus we obtain for the current density,

jq=1Ωk,σkmnσ(k)+1Ω2k,σk,σ[nσ(0)(k)+δnσ(k)]1kfσσ(k,k)δnσ(k)=1Ωk,σkmδnσ(k)1Ω2k,σk,σ1[knσ(0)(k)]fσσ(k,k)δnσ(k)+O(|q|2)=1Ωk,σkmδnσ(k)+1Ω2k,σk,σfσσ(k,k)(δ(ϵσ(k)ϵF)2km)(q2kmδ(ϵσ(k)ϵF))+O(|q|2)j1+j2.

where, for the second line, we performed an integration by parts and neglect terms quadratic in δn and, in the third line, used fσσ(k,k)=fσσ(k,k) and

knσ(0)(k)=nσ(0)(k)ϵσ(k)kϵσ(k)=δ(ϵσ(k)ϵF)kϵσ(k)=δ(ϵσ(k)ϵF)2km.

The first term denotes quasi-particle current, j1, while the second term can be interpreted as a drag current, j2, an induced motion (backflow) of the other particles due to interactions.
From a different viewpoint, we consider the system as being in the inertial frame with a velocity q/m, as all particles received the same momentum. Here m is the bare electron mass. The current density is then given by

jq=NΩqm=1Ωk,σkmnσ(k)=1Ωk,σkmδnσ(k).

Since these two viewpoints have to be equivalent, the resulting currents should be the same. Thus, we compare the expressions for current density and obtain the equation,

km=km+1Ωk,σfσσ(k,k)(δ(ϵσ(k)ϵF)km)

which with k^=k/kF then leads to

1m=1m+N(ϵF)dΩk^4πfs(k^,k^)k^k^m=1m+1mdΩk^4πN(ϵF)fs(k^,k^)l=0FlsPl(cosθk^k^)cosθk^k^P1(cosθk^k^)

or by using the orthogonality of the Legendre polynomials,

mm=1+13F1s

where 1/3=1/(2l+1) for l=1 originates from the orthogonality relation of Legendre polynomials, as shown above. Therefore, this relation has to couple m to F1s in order for Landau's theory of Fermi liquids to be self-consistent. Generally, we find that F1s>0 so that quasi-particles in a Fermi liquid are effectively heavier than bare electrons.

5.2.5 Stability of the Fermi Liquid

Upon inspection of the renormalisation of the quantities treated previously

γγ0=mmκκ0=mm11+F0sχχ0=mm11+F0a

with

mm=1+13F1s

and the response functions of the non-interacting system are given by

γ0=kB2mkF32,κ0=3mn2kF2andχ0=μB2mkFπ22

one notes that the compressibility κ (susceptibility χ) diverges for F0s1 (F0a1), indicating an instability of the system. A diverging spin susceptibility, for example, leads to a ferromagnetic state with a split Fermi surface, one for each spin direction. On the other hand, a diverging compressibility leads to a spontaneous contraction of the system. More generally, the deformation of the quasi-particle distribution function may vary over the Fermi surface, so that arbitrary deviations of the Fermi liquid ground state may be classified by the deformation

δnσ(k^)=l=0m=l+lδnσ,l,mYlm(θk,ϕk).

Note that we allow here formally for complex distribution functions. For pure charge density deformations we have δn+,l,m(k^)=δn,l,m(k^), while pure spin density deformations are described by δn+,l,m(k^)=δn,l,m(k^). The general response function for a redistribution δnσ(k^) with the anisotropy Ylm(θk,ϕk) is given by

χl,m=χl,m(0)1+Fls,a2l+1.

This comes from here:
General response and distribution deformations: We consider a force field F with conjugate "polarisation" P which yields a modification of the quasi-particle dispersion,

δϵσ(k)=αλσ(k)F and δϵ~σ(k)=α~λσ(k)F

where we assume that λσ(k)=Yl,m(θk^,ϕk^)=(1)mYl,m(θk^,ϕk^) without spin dependence. Then we can write

δϵ~σ(k)=δϵσ(k)+1Ωk,σfσσ(k^,k^)δnσ(k)=δϵσ(k)+1Ωk,σfσσ(k^,k^)nσϵ~σ(k)δϵ~σ(k).

In the last step we take for δnσ(k) the self-consistent value taking the feedback of the quasi-particle coupling into account. We now use the relation

fs,a(k^,k^)=l=0fls,aPl(k^k^)=4πl=0fls,a2l+1m=l+lYlm(θk^,ϕk^)Ylm(θk^,ϕk^)

and find,

α~λσ(k)F=αλσ(k)FN(ϵF)l=0fls2l+1m=l+lYlm(θk^,ϕk^)α~dΩk^Ylm(θk^,ϕk^)λσ(k)δllδmmF

which leads straightforwardly to

α~=αα~Fls2l+1α~=α1+Fls2l+1.

Now the polarisation is calculated which we may define as

P=1Ωk,σαλσ(k)δnσ(k)=1Ωk,σαλσ(k)nσϵ~σ(k)δϵ~σ(k)=1Ωk,σα(δ(ϵσ(k)ϵF))(α~|λ(k)|2F)=αα~N(ϵF)FdΩk4π|λ(k)|2=αα~N(ϵF)F4π

such that the linear response is given by

χ=PF=α2N(ϵF)4π(1+Fls2l+1).

Stability of the Fermi liquid against any of these deformations requires

1+Fls,a2l+1>0.

If for any deformation channel l this conditions is violated one talks about a "Pomeranchuk instability".Generally, the renormalisation of the Fermi liquid leads to a change in the Wilson ratio, defined as

RR0=χ/χ0γ/γ0=1+F1s/31+F0a

where R0=χ0/γ0=(g/2)23μB2/(π2kB2). Note that the Wilson ratio does not depend on the effective mass m/m directly through F1s in this form, but rather through the χ0 and γ0 if m is considered part of the "bare" parameters for them.
A remarkable feature of the Fermi liquid theory is that even very strongly interacting Fermions remain Fermi liquids, notably the quantum liquid 3He and so-called heavy Fermion systems, which are compounds of transition metals and rare earths. Both are strongly renormalised Fermi liquids. For 3He we give some of the parameters in Table 5.1 both for zero pressure and for pressures just below the critical pressure at which He solidifies (pc2.5MPa=25bar).

Pressure m/m F0s F0a F1s κ/κ0 χ/χ0
p=0 3.0 10.1 -0.52 6.0 0.27 6.3
p<pc 6.2 94 -0.74 15.7 0.065 24

The trends show obviously, that the higher the applied pressure is, the denser the liquid becomes and the stronger the quasi-particles interact. Approaching the solidification the compressibility is reduced, the quasi-particles become heavier (slower) and the magnetic response increases drastically. Finally the heavy Fermion systems are characterised by the extraordinary enhancements of the effective mass which for many of these compounds lie between 100 and 1000 times higher than the bare electron mass (for example CeAl3,UBe13). This large masses lead the notion of almost localised Fermi liquids, since the large effective mass is induced by the hybridisation of itinerant conduction electrons with strongly interacting (localised) electron states in partially filled 4f- or 5f-orbitals of Lanthanide and Actinide atoms, respectively.


5.3 Microscopic Considerations

A rigorous derivation of Landau's Fermi liquid theory requires methods of quantum field theory and would go beyond the scope of these lectures. However, plain Rayleigh-Schrödinger theory applied to a simple model allows to gain some insights into the microscopic fundament of this phenomenological theory. In the following, we consider a model of Fermions with contact interaction Uδ(rr), described by the Hamiltonian

H=k,sϵkc^ksc^ks+d3rd3rΨ^(r)Ψ^(r)Uδ(rr)Ψ^(r)Ψ^(r)=k,sϵkc^ksc^ks+UΩk,k,qc^k+qc^kqc^kc^k

where ϵk=2|k|2/(2m) is a parabolic dispersion of non-interacting electrons. We previously noticed that, in order to find well-defined quasi-particles, the interaction between the Fermions has to be short-ranged. This specially holds for the contact interaction.

5.3.1 Landau Parameters

Starting from the previously defined Hamiltonian, we will determine Landau parameters for a corresponding Fermi liquid theory. For a given momentum distribution nks=c^ksc^ks=nks(0)+δnks, we can expand the energy resulting from the Hamiltonian following the Rayleigh-Schrödinger perturbation method,

E=E(0)+E(1)+E(2)+

with

E(0)=k,sϵknks,E(1)=UΩk,knknk,E(2)=U2Ω2k,k,qnknk(1nk+q)(1nkq)ϵk+ϵkϵk+qϵkq.

The second order term E(2) describes virtual processes corresponding to a pair of particle-hole excitations. The numerator of this term can be split into four different contributions.
We first consider the term quadratic in nk and combine it with the first order term E(1), which has the same structure,

E~(1)=E(1)+U2Ω2k,k,qnknkϵk+ϵkϵk+qϵkqU~Ωk,knknk.

In the last step, we defined the renormalised interaction U~ through,

U~=U+U2Ωq1ϵk+ϵkϵk+qϵkq.

In principle, U~ depends on the wave vectors k and k. However, when the wave vectors are restricted to the Fermi surface (|k|=|k|=kF), and if the range of the interaction is small compared to the mean electron spacing, that is, kF1, this dependency may be neglected.
We should be careful with our choice of a contact interaction, since it would lead to a divergence in the large-q range. A cutoff for q of order Qc1 would regularise the integral which is dominated by the large-q part. Thus we may use the following expansion (with q=|q|),

1Ωq1ϵk+ϵkϵk+qϵkq=1(2π)320Qcdqq2dΩqm(kk)q|q|2=m(2π)220Qcdqq1+1dcosθKcosθq=m(2π)220Qcdqq1Kln|qKq+K|=m(2π)22(Qc+K2Qc22Kln|QcKQc+K|)mQc(2π)22(1K22Qc2+O(K4Qc4))=N(ϵF)Qc2kFπ22mkFmQc(2π)22N(ϵF)πQc4kF2

where we use K=|kk|2kFQc. From this we conclude that the momentum dependence of U~ is indeed weak.
Since the term quartic in nk vanishes due to symmetry, the remaining contribution to E(2) is cubic in nk and reads

E~(2)=U~2Ω2k,k,qnknk(nk+q+nkq)ϵk+ϵkϵk+qϵkq.

We replaced U2 by U~2, which is admissible at this order of the perturbative expansion. The variation of the energy E with respect to δnk can easily be calculated,

ϵ~(k)=ϵk+U~ΩknkU~2Ω2k,qnk(nk+q+nkq)nk+qnkqϵk+ϵkϵk+qϵkq,

and an analogous expression is found for ϵ(k). The coupling parameters fσσ(k,k) may be determined using the definition fσσ=Ωδ2Eδnσδnσ. Starting with f↑↑(kF,kF) with wave-vectors on the Fermi surface (|kF|=|kF|=kF), the terms contributing to the coupling can be written as

U~2Ω2k,qnk+qnkqnkϵk+ϵkϵk+qϵkqk+qkF1ΩkFnkFU~2Ωknkq(0)nk(0)ϵkϵkq|q=kFkF=1ΩkFnkFU~22χ0(kFkF),

where we consider nkF=nkF(0)+δnkF. Note that the part in this term which depends on nkF(0) will contribute the ground state energy in Landau's energy functional. Here, χ0(q) is the static Lindhard susceptibility. With the help of the definition fσσ=Ωδ2Eδnσδnσ, it follows immediately, that

f↑↑(kF,kF)=f↓↓(kF,kF)=U~22χ0(kFkF).

The other couplings are obtained in a similar way, resulting in

f↑↓(kF,kF)=f↓↑(kF,kF)=U~U~22[2χ~0(kF+kF)χ0(kFkF)],

where the function χ~0(q) is defined as

χ~0(q)=1Ωknk+q(0)+nk(0)2ϵFϵk+qϵk.

If the couplings are parameterised by the angle θ between kF and kF, they can be expressed as

fσσ(θ)=U~2[(1+U~N(ϵF)4(2+cosθ2sin(θ/2)ln1+sin(θ/2)1sin(θ/2)))δσσ(1+U~N(ϵF)4(1sin(θ/2)2ln1+sin(θ/2)1sin(θ/2)))(1δσσ)].

Finally, we are in the position to determine the most important Landau parameters by matching these expressions to the parametrisation,

F0s=u~[1+u~(1+16(2+ln2))]F0a=u~[1+u~(123(1ln2))]F1s=u~2215(7ln21)

where u~=U~N(ϵF)/2 has been introduced for better readability. Since the Landau parameter F1s is responsible for the modification of the effective mass m compared to the bare mass m, m is enhanced compared to m for both attractive (U<0) and repulsive (U>0) interactions. Obviously, the sign of the interaction U does not affect the renormalisation of the effective mass m. This is so, because the existence of an interaction (whatever sign it has) between the particles enforces the motion of many particles whenever one is moved. The behaviour of the susceptibility and the compressibility depends on the sign of the interaction. If the interaction is repulsive (u~>0), the compressibility decreases (F0s>0), implying that it is harder to compress the Fermi liquid. The susceptibility is enhanced (F0a<0) in this case, so that it is easier to polarise the spins of the electrons. Conversely, for attractive interactions (u~<0), the compressibility is enhanced due to a negative Landau parameter F0s, whereas the susceptibility is suppressed with a factor 1/(1+F0a), with F0a>0. The attractive case is more subtle because the Fermi liquid becomes unstable at low temperatures, turning into a superfluid or superconductor, by forming so-called Cooper pairs. This represents another non-trivial Fermi surface instability.

5.3.2 Distribution Function

Finally, we examine the effect of interactions on the ground state properties, using again Rayleigh-Schrödinger perturbation theory. The calculation of the corrections to the ground state |Ψ0, the filled Fermi sea can be expressed as

|Ψ=|Ψ(0)+|Ψ(1)+

where

|Ψ(0)=|Ψ0|Ψ(1)=UΩk,k,qs,sc^k+q,sc^kq,sc^k,sc^k,sϵk+ϵkϵk+qϵkq|Ψ0.

The state |Ψ0 represents the ground state of non-interacting Fermions. The lowest order correction involves particle-hole excitations, depleting the Fermi sea by lifting particles virtually above the Fermi energy. How this correction affects the distribution function, will be discussed next. The momentum distribution nks=c^ksc^ks is obtain as the expectation value,

nks=Ψ|c^ksc^ks|ΨΨΨ=nks(0)+δnks(2)+

where nks(0) is the unperturbed distribution Θ(kF|k|), and

δnks(2)={U2Ω2k1,k2,k3(1nk1)(1nk2)nk3(ϵk+ϵk3ϵk1ϵk2)2δk+k3,k1+k2|k|<kFU2Ω2k1,k2,k3nk1nk2(1nk3)(ϵk1+ϵk2ϵkϵk3)2δk+k3,k1+k2|k|>kF.

This yields the modification of the distribution functions as shown here:

Attachments/Script 57.webp|700

It allows us also to determine the size of the discontinuity of the distribution function at the Fermi surface,

nkFnkF+=1(UN(ϵF)2)2ln(2),

where

nkF±=lim|k|kF0±(nk(0)+δnk(2)).

The jump of nk at the Fermi surface is reduced independently of the sign of the interaction. The reduction is quadratic in the perturbation parameter UN(ϵF). This jump is also a measure for the weight of the quasi-particle state at the Fermi surface.

5.3.3 Fermi Liquid in One Dimension?

Within a perturbative approach the Fermi liquid theory can be justified for a three-dimensional system, and we recognise the one-to-one correspondence between bare electrons and quasi-particles renormalised by (short-ranged) interactions. Now we would like to show that within the same approach problems appear in one-dimensional systems, which are of a conceptual nature and hint that interacting Fermions in one dimension would not form a Fermi liquid, but a Luttinger liquid, as we will motivate briefly below.
The Landau parameters have been expressed above in terms of the response functions χ0(q=kFkF) and χ~0(q=kFkF). For the one-dimensional system, the relevant contributions come from two configurations, since there are two Fermi points only (instead of a two-dimensional Fermi surface),

(kF,kF)q=kFkF=0,±2kF.

We find that the response functions show singularities for some of these momenta, and we obtain

f↑↑(±kF,±kF)=f↓↓(±kF,kF)

as well as

f↑↓(kF,±kF)=f↓↑(kF,kF)

giving rise to the divergence of all Landau parameters. Therefore the perturbative approach to a Landau Fermi liquid is not allowed for the one-dimensional Fermi system.

The same message is obtained when looking at the momentum distribution form which had in three dimensions a step giving a measure for the (reduced but finite) quasi-particle weight. The analogous calculation as in the previous subsection leads here to (for scalar k in 1D)

nks(2){18π2U22vF2lnk+kkFk>kF18π2U22vF2lnkkFkk<kF.

Here, k±are cutoff parameters of the order of the Fermi wave vector kF. Apparently the quality of the perturbative calculation deteriorates as kkF±, since we encounter a logarithmic divergence from both sides.

Attachments/Script 58.webp|700

Indeed, a more elaborated approach shows that the distribution function is continuous at k=kF in one dimension, without any jump. Correspondingly, the quasi-particle weight vanishes and the elementary excitations cannot be described by Fermionic quasi-particles but rather by collective modes. Landau's Theory of Fermi liquids is inappropriate for such systems. This kind of behaviour, where the quasi-particle weight vanishes, can be described by the so-called bosonisation of Fermions in one dimension, a topic that is beyond the scope of these lectures. However, a result worth mentioning, shows that the Fermionic excitations in one dimensions decay into independent charge and spin excitations, the so-called spin-charge separation. This behaviour can be understood with the naive picture of a half-filled lattice with predominantly antiferromagnetic spin correlations. In this case both charge excitations (empty or doubly occupied lattice site) and spin excitations (two parallel neighbouring spins) represent different kinds of domain walls, and are free to move at different velocities.