Jump back to chapter selection.


Table of Contents

7.1 Stoner Instability
7.2 General Spin Susceptibility and Magnetic Instabilities
7.3 Stoner Excitations


7 Magnetism in Metals

Magnetic ordering in metals can be viewed as an instability of the Fermi liquid state. We introduce this new phase of metals through the description of the Stoner ferromagnetism. The discussion of antiferromagnetism and spin density wave phases will be only brief in this chapter. In Stoner ferromagnets the magnetic moment is provided by the spin of itinerant electrons. Magnetism due to localised magnetic moments will be considered in the context of Mott insulators, which are the subject of the next chapter.
Well-known examples of elemental ferromagnetic metals are iron (Fe), cobalt (Co), and nickel (Ni) belonging to the 3d transition metals, where the 3d-orbital character is dominant for the conduction electrons at the Fermi energy. These orbitals are rather tightly bound to the atomic cores such that the electron mobility is reduced, enhancing the effect of interaction which is essential for the formation of a magnetic state. Other forms of magnetism, such as antiferromagnetism and the spin density wave state are found in the 3d transition metals Cr and Mn. Note, 4d and 5d transition metals within the same columns of the periodic system are not magnetic. Their d-orbitals are more extended, leading to a higher mobility of the electrons, such that the mutual interaction is insufficient to trigger magnetism. It is, however, possible to find ferromagnetism in ZrZn2 where zinc (Zn) may act as a spacer reducing the mobility of the 4d-electrons of zirconium (Zr). The 4d-elements Pd and Rh, and the 5d-element Pt are, however, nearly ferromagnetic. Going further in the periodic table, the 4f-orbitals appearing in the lanthanides are nearly localised and can lead to ferromagnetism, as illustrated by the elements going from Gd through Tm in the periodic system.
Magnetism appears through a phase transition, meaning that the metal is non-magnetic at temperatures above a critical temperature Tc, the Curie temperature. In many cases, magnetism appears at Tc as a continuous, second order phase transition involving the spontaneous violation of symmetry. This transition is lacking latent heat (no discontinuity in entropy and volume) but instead features a discontinuity in the specific heat.


7.1 Stoner Instability

In the following section, we study the emergence of the metallic ferromagnetism originating from the Stoner mechanism. In close analogy to the first Hund's rule, the exchange interaction among the electrons plays a crucial role here. The alignment of the electronic spins in a favoured direction allows the system to reduce the energy contribution due to Coulomb repulsion. According to Landau's theory of Fermi liquids, the interaction between electrons renormalises the spin susceptibility χ0 to

χ=mmχ01+F0a,

which obviously diverges for F0a1 and leads to a ferromagnetic instability of the Fermi liquid. Then, F0aUN(ϵF)/2 provides a critical value for the interaction Uc=2/N(ϵF) such that F0a=1 and χ diverges. We will see below that this corresponds to a value we will derive also by a mean field theory.

7.1.1 Stoner Model Within the Mean Field Approximation

Consider the following model for conduction electrons with a repulsive contact interaction,

H=k,sϵkc^ksc^ks+Ud3rρ^(r)ρ^(r),

where we use the electron density ρ^s(r)=Ψ^s(r)Ψ^s(r) and the field operator Ψ^s(Ψ^s) follows from the previously established definition. The contact interaction is an approximation of the screened Coulomb interaction. Due to the Pauli exclusion principle, the contact interaction is only active between electrons with opposite spins. This is a consequence of the exchange hole in the two-particle correlation between electrons of identical spin. We obtain a useful insight into mechanisms leading to ferromagnetism by means of a mean field approximation. Note that the following mean field calculation is equivalent to a variational approach using simple many-body wave-function (Slater determinant) with different concentrations of up and down spins. We rewrite,

ρ^s(r)=ns+[ρ^s(r)ns],

where

ns=ρ^s(r),

and represents the thermal average. We stipulate that the deviation from the mean value ns is small in the sense that

[ρ^(r)n][ρ^(r)n]nn.

Inserting this decomposition into the Hamiltonian for conduction electrons, we obtain

Hmf=k,sϵkc^ksc^ks+Ud3r[ρ^(r)n+ρ^(r)nnn]=k,s(ϵk+Uns¯)c^ksc^ksUΩnn

the mean field Hamiltonian, describing electrons which move in the uniform background of electrons of opposite spin coupling via the spin dependent exchange interaction (s¯ denotes the spin opposite to s). Fluctuations of the form [ρ^(r)n][ρ^(r)n] are neglected here. The advantage of this approximation is, that the many-body problem is now reduced to an effective one-particle problem, where only the mean electron interaction is taken into account. This is equivalent to a generalised Hartree-Fock approximation and enables us to calculate certain expectation values, such as the density of one spin species, for example;

n=1Ωkc^kc^k=1Ωkf(ϵk+Un)=dϵ1Ωkδ(ϵϵkUn)f(ϵ)=dϵ12N(ϵUn)f(ϵ)

where f(ϵ) is the Fermi-Dirac distribution function. An analogous result is found for the opposite spin direction. These mean densities are determined self-consistently, namely such that the insertion of ns into the mean-field Hamiltonian provides the correct output according to the expression for n. Furthermore, the constraint that the total number of electrons is conserved, must be implemented. The real magnetisation M=μBm is proportional to m which is defined via

ns=12[(n+n)+s(nn)]=n0+sm2

where n0 is the total particle density and s=±1. This leads to the two coupled equations

n0=12dϵ(N(ϵUn)+N(ϵUn))f(ϵ)m=12dϵ(N(ϵUn)N(ϵUn))f(ϵ)

or equivalently

n0=12s=±1dϵN(ϵUn02sUm2)f(ϵ)m=12s=±1sdϵN(ϵUn02sUm2)f(ϵ)

which usually can not be solved analytically and must be treated numerically.

7.1.2 Stoner Criterion

An approximate solution can be found if mn0. The two equations are solved by adapting the chemical potential μ. For low temperatures and small magnetisation we can expand μ as

μ(m,T)=ϵF+Δμ(m,T).

The constant energy shift Un0/2 appearing can be absorbed into ϵF. The Fermi-Dirac distribution takes the form

f(ϵ)=1eβ[ϵμ(m,T)]+1

where β=(kBT)1. After expanding for small m, one obtains using the Sommerfeld expansion,

n0dϵf(ϵ)[N(ϵ)+12(Um2)2N(ϵ)]0ϵFdϵN(ϵ)+N(ϵF)Δμ+π26(kBT)2N(ϵF)+12(Um2)2N(ϵF)

where we introduced the abbreviations N(ϵ)=dN(ϵ)/dϵ and N(ϵ)=d2N(ϵ)/dϵ2. Since the first term on the right side is identical to n0, Δμ(m,T) is immediately found to be given by

Δμ(m,T)N(ϵF)N(ϵF)[π26(kBT)2+12(Um2)2].

Analogously, the expansion of the equation for m in m and T, results in

mdϵf(ϵ)[N(ϵ)Um2+13!N(ϵ)(Um2)3][N(ϵF)+π26(kBT)2N(ϵF)+13!(Um2)2N(ϵF)+ΔμN(ϵF)](Um2)

and, finally, inserting the result for Δμ, we find

m=N(ϵF)[1π26(kBT)2Λ12(ϵF)](Um2)N(ϵF)Λ22(ϵF)(Um2)3

where

Λ12(ϵF)=(N(ϵF)N(ϵF))2N(ϵF)N(ϵF),

and

Λ22(ϵF)=12(N(ϵF)N(ϵF))2N(ϵF)3!N(ϵF).

The structure is m=am|b|m3 (assuming b<0 for stable ferromagnetism), where b=N(ϵF)Λ22(U/2)3. Thus, two types of solutions emerge

m2={0,a<1a1|b|,a1

With this, a=1 corresponds to a critical value.

Attachments/Script 69.webp|700

Here, this condition corresponds to

1=12UN(ϵF)[1π26(kBTC)2Λ12(ϵF)]

yielding

kBTC=6πΛ1(ϵF)12UN(ϵF)1UcU

for U>Uc=2/N(ϵF). This is an instability condition for the paramagnetic Fermi liquid state with m=0, and TC is the Curie temperature, below which the ferromagnetic state appears. The temperature dependence of the magnetisation M of the ferromagnetic state (T<TC) is given by

M(T)=μBm(T)TCT

close to the phase transition (TCTTC). Note that the Curie temperature TC is nonzero for UN(ϵF)>2, and TC0 in the limit UN(ϵF)2+. For UN(ϵF)<2 no phase transition occurs. This condition for a finite transition temperature TC is known as the Stoner criterion. This simple model also describes a so-called quantum phase transition, that is, a phase transition that appears at T=0 as a function of system parameters, which in our case are the density of states N(ϵF) and the Coulomb repulsion U. While thermal fluctuations destroy the ordered state at finite temperature via entropy increase, entropy is irrelevant at T=0. Here, the order is suppressed by quantum fluctuations (Heisenberg's uncertainty principle). At zero temperature we find for m the following dependence on U:

m(U)=1Λ2(ϵF)(2UN(ϵF))3/2(UN(ϵF)21)1/2(U2)3/2

for U>Uc and m=0 for U<Uc. The density of states as an internal parameter can, for example, be changed by applying a pressure. By reducing the lattice constant, pressure may facilitate the motion of the conduction electrons and increase the Fermi velocity. Consequently, the density of states is reduced:

Attachments/Script 70.webp|700

Indeed, pressure is able to destroy ferromagnetism in weakly ferromagnetic materials as ZrZn2, MnSi, and UGe2. In other materials, the Curie temperature is high enough, such that the technologically applicable pressure is insufficient to suppress magnetism. It is, however, possible, that pressure leads to other transitions, such as structural phase transitions, that eventually destroy magnetism. This is seen in iron (Fe), where a pressure of about 12GPa induces a transition from magnetic iron with body-centred crystal (bcc) structure to a nonmagnetic, hexagonal close packed (hcp) structure:

Attachments/Script 71.webp|700

While this structural transition is a quantum phase transition as well, it appears as a discontinuous, first order transition. Note that the Stoner instability is a simplification of the quantum phase transition. In most cases, a discontinuous phase transition originates in the band structure or in fluctuation effects, which were ignored here. In some cases, pressure can also induce an increase in N(ϵF), for example in metals with multiple bands, where compression leads to a redistribution of charge. One example is the ruthenate Sr3Ru2O7 for which uniaxial pressure along the z-axis leads to magnetism. Finally, let us turn to the question, why Cu, being a direct neighbour of Ni in the 3d-row of the periodic table, is not ferromagnetic, even though both elemental metals share the same fcc crystal structure. The answer is given by the Stoner criterion UN(ϵF)=2. While the conduction electrons at the Fermi level of Ni have 3d-character and belong to a narrow band with a large density of states, the Fermi energy of Cu is situated in the broad 4s-band and constitutes a much smaller density of states:

Attachments/Script 72.webp|700

With this, the Cu conduction electrons are much less localised and feature a weaker tendency towards ferromagnetic order.

7.1.3 Spin Susceptibility for T>TC

Next we study the response of the metallic system in the paramagnetic state when we apply a small magnetic field H along the z-axis, which induces a spin polarisation due to the Zeeman coupling,

HZ=gμBd3rH12{ρ^(r)ρ^(r)}.

From the self-consistency equations we obtain

m=12dϵf(ϵ)s=±1sN(ϵμBsHsUm2)dϵf(ϵ)N(ϵ)(Um2+μBH)=N(ϵF)[1π26(kBT)2Λ1(ϵF)2](Um2+μBH)

to lowest order in m and H. Solving this equation for m yields

M=μBm=χ0(T)1Uχ0(T)/(2μB2)H,

and, consequently, the magnetic susceptibility χ reads

χ=MH=χ0(T)1Uχ0(T)/(2μB2),

where the bare susceptibility χ0 is given by

χ0(T)=μB2N(ϵF)[1π26(kBT)2Λ1(ϵF)2].

We see, that the denominator of the susceptibility χ(T) vanishes exactly when the Stoner instability criterion for finite temperatures is fulfilled. Thus, for UN(ϵF)>2 the susceptibility

χ(T)χ0(TC)(UN(ϵF)21)(T2TC21)=χ0(TC)UN(ϵF)π212(kBTC)2(12UN(ϵF))1Λ12(ϵF)kB2(T2TC2)

diverging at TC indicates the instability. Note that for TTC from the paramagnetic side, the susceptibility diverges as χ(T)|TCT|1 corresponding to the mean field behaviour, since the mean field critical exponent γ for the susceptibility takes the value γ=1.
For the case of UN(ϵF)<2 there is no instability down to T=0. In the zero-temperature limit we obtain

χ(0)=χ0(0)1UN(ϵF)2=μB2N(ϵF)1+F0a,

corresponding to the form found in the Landau Fermi liquid theory with F0a>1. Note that χ0(0)=μB2N(ϵF) is the Pauli spin susceptibility (assuming g=2).


7.2 General Spin Susceptibility and Magnetic Instabilities

The ferromagnetic state is characterised by a uniform magnetisation. There are, however, magnetically ordered states which do not feature a non-zero net magnetisation but specially modulated magnetic moments. Examples are spin density wave (SDW) states, antiferromagnets and spin spiral states. In this section, we analyse general instability conditions for metallic systems towards some magnetic ordering.


7.2.1 General Dynamic Spin Susceptibility

We consider a magnetic field, oscillating in time and with spatial modulation

H(r,t)=H0eiqriωteηt

where η0+ yields an adiabatic switching on of the field. We calculate the resulting magnetisation, for the corresponding Fourier component. For that, we proceed analogously to the previously discussed calculation of susceptibility and define the spin density operator S^(r) in real space,

S^(r)=2s,sΨ^s(r)σssΨ^s(r)=2(Ψ^(r)Ψ^(r)+Ψ^(r)Ψ^(r)iΨ^(r)Ψ^(r)+iΨ^(r)Ψ^(r)Ψ^(r)Ψ^(r)Ψ^(r)Ψ^(r))

with momentum space representation

S^q=d3rS^(r)eiqr=2k,s,sck,sσssck+q,s=kS^k,q

where S^k,q=s,s(/2)ck,sσssck+q,s. The Hamiltonian of the electronic system with contact interaction is given by

H=H0+HZ+Hint,

where

H0=k,sϵkc^ksc^ks,HZ=gμBd3rH(r,t)S^(r),Hint=Ud3rρ^(r)ρ^(r).

The operator HZ describes the Zeeman coupling between the electrons of the metal and the perturbing field. We investigate a magnetic field

H=H+(q,ω)eiqriωteηt(1i0)+ h.c. 

in the xy-plane. The Zeeman term then simplifies to

HZ=gμBH+(q,ω)S^qeiωteηt+ h.c. 

where S^q±=S^qx±iS^qy. In the following the Hermitian conjugate (h.c.) part will be ignored. We use

S^q=kc^kc^kq,

in the c-operator representation. In the framework of linear response theory, this coupling will induce a magnetisation Mind +(q,t)=(μB/Ω)S^q+(t)eiωt+ηt. Using a similar equation of motion formalism as used previously for charge susceptibility,

itS^k,q+=[S^k,q+,H]

with S^k,q+=c^kc^k+q, we can determine this induced magnetisation, first without the interaction term (U=0). We obtain for the given Fourier component,

itS^k,q+(t)=(ϵk+qϵk)S^k,q+(t)gμB(nknk+q)H+(q,ω)eiωt+ηt

Using the monochromatic time dependence of the field and the response (eiωt+ηt) and applying the thermal average we obtain,

(ϵk+qϵkωiη)Sk,q+(t)=gμB(nk+qnk)H+(q,ω)eiωt+ηt

which then leads to the induced spin density-magnetisation,

μBMind+(q,ω)eiωt+ηt=1ΩSind +(q,ω)eiωt+ηt=1ΩkSk,q+(t)=μBχ0(q,ω)H+(q,ω)eiωt+ηt

with (assuming g=2 for consistency with later expressions, particularly the RPA denominator and Pauli susceptibility limit)

χ0(q,ω)=2μB2Ωknk+qnkϵk+qϵkωiη

Note that the form of the bare susceptibility χ0(q,ω) is similar to the Lindhard function, as discussed earlier in the context of charge susceptibility, actually identical, if there is no spin polarisation. This result for the induced spin density describes the induced spin density within linear response approximation.
In a next step, we want to include the effects of the interaction. Analogously to the induced charge modulation, discussed previously, the induced spin density generates an effective field on the spin of the electrons ("mean field"). The induced spin polarisation may be represented as an effective magnetic field through the exchange interaction. To implement this feature let us rewrite the contact interaction term in the form

Hint=UΩk,k,qc^k+qc^kc^kqc^k=U2Ωk,k,q{c^kc^k+qc^kc^kq+c^kc^k+qc^kc^kq}+UΩk(n^k+n^k)=U2Ω2q{S^q+S^q+S^qS^q+}+UΩkn^k.

The last term proportional to n^k=n^k+n^k can be absorbed into the term of the chemical potential. The induced spin polarisation Sind +(q,ω) acts through the exchange interaction as an effective (local) field, as can be seen by replacing S^q+Sind +(q,ω)δq,q,

U2Ω2q{S^q+S^q+S^qS^q+}UΩ2S+(q,ω)S^q=gμBΩHind+(q,ω)S^q

where the effective magnetic field Hind +(q,ω) finally reads (assuming g=2 for consistency with χ0)

Hind+(q,ω)=UΩgμBS+(q,ω)=U2ΩμBS+(q,ω).

with the same monochromatic time dependence as above. This induced field acts on the spins as well, such that the total response of the spin density on the external field becomes

M+(q,ω)=μBΩS+(q,ω)=χ0(q,ω)[H+(q,ω)+Hind+(q,ω)]=χ0(q,ω)H+(q,ω)+χ0(q,ω)U2ΩμBS+(q,ω)=χ0(q,ω)H+(q,ω)+χ0(q,ω)U2μB2M+(q,ω).

In the last step we introduce self-consistency taking the induced magnetisation as the real magnetisation. With the definition

M+(q,ω)=χ(q,ω)H+(q,ω)

of the susceptibility we find

χ(q,ω)=χ0(q,ω)1U2μB2χ0(q,ω).

which corresponds to the random phase approximation (RPA), as discussed earlier. This form of the susceptibility is found to be valid for all field directions, as long as spin-orbit coupling is neglected. Within the random phase approximation, the generalisation of the Stoner criterion for the appearance of an instability of the system at finite temperature reads

1=U2μB2χ0(q,ω).

For the limiting case (q,ω)(0,0) corresponding to a uniform, static external field, we obtain for the bare susceptibility

χ0(q,0)=2μB2Ωknk+qnkϵk+qϵkq02μB2Ωkf(ϵk)ϵk=χ0(T),

which corresponds to the Pauli susceptibility (with g=2). Then, χ(T) (the full susceptibility) is again cast into the form previously discussed for the simpler ferromagnetic case and describes the instability of the metal with respect to ferromagnetic spin polarisation, when the denominator vanishes. Similar to the charge density wave, the isotropic deformation for q=0 is not the leading instability, when χ0(Q,0)>χ0(0,0) for a finite Q. Then, another form of magnetic order is more favoured.


7.2.2 Instability with Finite Wave Vector Q

In order to show that, indeed, the Stoner instability does not always prevail among all possible magnetic instabilities, we first go through a simple argument based on the local susceptibility. For that, we define the local magnetic moment along the z-axis, M(r)=μBρ^(r)ρ^(r), and consider the non-local relation

M(r)=d3rχ~0(rr)Hz(r)

within the linear response approximation. In Fourier space, the same relation reads

Mq=χ0(q)Hq,

with

χ0(q)=d3rχ~0(r)eiqr

Now, compare χ0(q=0) with χ0(q) defined as

χ0(q)=1Ωqχ0(q)=1Ωqd3rχ~0(r)eiqr=d3rχ~0(r)δ(r)=χ~0(r=0)

This q-averaged susceptibility corresponds to the local susceptibility. For a paramagnetic metal at T=0 we may write (using a "per spin" density of states N(ϵ) and corresponding susceptibility definition consistent with χ0(q=0)=μB2N(ϵF) for this subsection's comparison):

χ0(q)=2μB2Ω2k,qnk+qnkϵkϵk+q=μB22dϵN(ϵ)dϵN(ϵ)f(ϵ)f(ϵ)ϵϵ,

and must be compared to χ0(q=0)=μB2N(ϵF)(f(ϵ)=Θ(ϵFϵ)). The local susceptibility depends on the density of states and the Fermi energy of the system. A very good qualitative understanding can be obtained by a very simple form

N(ϵ)={1D,DϵD0,|ϵ|>D

for the density of states which does not correspond to a particular band structure but mimics a band of width 2D. With this rough approximation, the integral for χ0(q) is easily evaluated. The ratio between χ(q) and χ0(q=0) is then found to be

R0=χ0(q)χ0(q=0)=ln(41η2)+ηln(1η1+η)

with η=ϵF/D where D<ϵF<+D. For both small and large band fillings (ϵF close to the band edges), the tendency towards ferromagnetism dominates, whereas when ϵF lies in the centre of the band, the susceptibility χ0(q) is not maximal at q=0 anymore, and magnetic ordering with a well-defined finite q=Q becomes more probable.

Attachments/Script 73.webp|700


7.2.3 Influence of the Band Structure

Whether magnetic order arises at finite q or not depends strongly on the details of the band structure. The argument given above, comparing the local (r=0) to the uniform (q=0) susceptibility is nothing more than a vague indicator for a possible instability at non-zero q. A crucial ingredient for the appearance of magnetic order at a given q=Q is the so-called nesting of the Fermi surface. Within extended areas of the Fermi surface the energy dispersion satisfies the nesting condition,

ξk+Q=ξk

where ξk=ϵkϵF and Q is some fixed vector. The nesting conditions connects for given k an electron- and hole-like band states (at T=0 filled and empty states, respectively). If the Fermi surface of a material features such a nesting trait, the susceptibility will be dominated by the contribution from this vector Q. In order to see this, let us investigate the static susceptibility χ0(q) for q=Q under the assumption that the nesting condition holds for all k (see tight-binding example below). Thus,

χ0(Q;T)=2μB2Ωknk+Qnkξkξk+Q=μB2d3k(2π)3f~(ξk)f~(ξk)ξk,

where f~(ξ)=f(ξ+ϵF)=f(ϵ)=[exp(ξ/kBT)+1]1 and f is the Fermi-Dirac distribution. Under the further assumption that ξk is weakly angle dependent, we find

χ0(Q;T)=μB2d3k(2π)3tanh(ξk/2kBT)ξk=μB22dξN(ξ+ϵF)tanh(ξ/2kBT)ξ.

In order to approximate this integral properly, we notice that the integral has a logarithmic divergence at infinite energies ξ. The band width gives a natural cutoff. Let us, therefore, take the density of states with ϵF=0,

χ0(Q;T)μB2N(ϵF)0ϵ0dξtanh(ξ/2kBT)ξ=μB2N(ϵF)(ln(ϵ02kBT)+ln(4eγπ))μB2N(ϵF)ln(1.14ϵ0kBT),

where we assumed ϵ0kBT, cutoff energy of the order of the band width, and where γ0.57721 is the Euler-Mascheroni constant. The bare susceptibility χ0 diverges logarithmically at zero temperatures. Inserting this result for χ0(Q;T) into the generalised Stoner relation, results in

0=1UN(ϵF)2ln(1.14ϵ0kBTc),

with the critical temperature

kBTc=1.14ϵ0e2/UN(ϵF).

A finite critical temperature persists for arbitrarily small positive values of UN(ϵF). The nesting condition for a given Q leads to a maximum of χ0(q,0;T) at q=Q and triggers the relevant instability in the system. The latter finally stabilises in a magnetic ordered phase with wave vector Q, the so-called spin density wave. The spin density distribution takes, for example, the form

S(r)=z^Scos(Qr),

without a uniform component. In comparison, the charge density wave was a modulation of the charge density with a much smaller amplitude than the height of the uniform density,

ρ(r)=ρ0+δρcos(Qr),

with δρρ0. The spin density states frequently appear in low-dimensional systems like organic conductors, or in transition metals such as chromium (Cr) for example. In all cases, nesting plays an important role:

Attachments/Script 74.webp|700

In quasi-one-dimensional electron systems, a main direction of motion dominates over two other directions with weak dispersion. In this case, the nesting condition is very probable to be fulfilled, as it is schematically shown in the centre panel of the figure above. Chromium is a three-dimensional metal, where nesting occurs between an electron-like Fermi surface around the Γ-point and a hole-like Fermi surface at the Brillouin zone boundary (H-point). These Fermi surfaces originate in different bands (right panel in the figure shown above). Chromium has a body-centred cubic crystal structure, where the H-point at (π/a,0,0) leads to the nesting vector Qx(1,0,0) and equivalent vectors in y- and z-direction, which are incommensurable with the lattice.

The textbook example of nesting is found in a tight-binding model of a simple cubic lattice with nearest-neighbour hopping at half filling. The band structure is given by

ϵk=2t[cos(kxa)+cos(kya)+cos(kza)],

where a is the lattice constant and t the hopping term. Because of half filling, the chemical potential μ=ϵF lies at μ=0 such that ξk=ϵk. Obviously, ϵk+Q=ϵk holds for all k, for the nesting vector Q=(π/a)(1,1,1). This full nesting trait is a signature of the total particle-hole symmetry, meaning in the ground state there are as many occupied as empty states. Analogously to the Peierls instability, the spin density wave induces the opening of a gap at the Fermi surface. This is another example of a Fermi surface instability. In this situation, the gap is confined to the areas of the Fermi surface obeying the nesting condition. Contrary to the ferromagnetic order, the material can become insulating when forming the spin density wave state.


7.3 Stoner Excitations

In this last section, we discuss the elementary excitations of the ferromagnetic ground state with n>n, including both particle-hole excitations and collective modes. For this purpose we use the Stoner model Hamiltonian, introduced earlier, which we write here entirely in momentum space operators,

H=k,sϵkc^ksc^ks+UΩk,k,qc^k+qc^kc^kqc^k.

The spin polarised ground state |ψg can be written on the mean field level as

|ψg=|k|kFc^k|k|kFc^k|o

with kFs=(6π2ns)1/3.
We now consider spin excitations, for which we make the Ansatz

|ψq=kfkc^k+qc^k|ψg=kfk|ψk,q.

This is a superposition of states where a spin up electron is removed from the ground state |ψg and placed back with opposite spin and a fixed momentum transfer q. The simple electron-hole excitation with such a spin flip corresponds to the state |ψk,q=c^k+qc^k|ψg and has the energy

Ek,q=ϵk+qϵk=ϵk+qϵk+U(nn)=ϵk+qϵk+Δ.

We have to ensure that an electron with (k,) is available to be removed, and that the state (k+q,) is unoccupied. The independent electron-hole part of the spectrum constitutes a continuum of excitations and is depicted by the shaded region:

Attachments/Script 75.webp|700

Note that the spin splitting of the spectrum opens a window in the low-energy low-momentum transfer sector of the excitations.
The excitation energy of |ψq can be obtained by solving the following Schrödinger equation

H|ψq=(Eg+ωq)|ψq,

with Eg as the ground state energy (H|ψg=Eg|ψg). Actually it is more convenient to rearrange this equation into the following form to eliminate Eg,

ωq|ψq=kfk[H,c^k+qc^k]|ψg.

Evaluating the commutator we obtain

ωq|ψq=(ϵk+qϵk)|ψq+UΩkfkk,q{c^k+qc^kc^k+qqc^kc^k+qc^kqc^kqc^k}|ψg.

The terms with four c^-operators are not so easy to handle. Therefore we adopt a mean field point of view in the analogous way as before and decouple these terms through the following approximation,

c^k+qc^kc^k+qqc^kc^k+qc^kqc^kqc^kδq,0(nknk)c^k+q,c^k+(nk+qδk,k+qnkδk,k+q)c^kq+qc^kq

The basic scheme is to combine c^-operators in the four-operator expression so as to obtain a density-operator-like expression c^ksc^ks (this introduces some constraints on the momenta implemented) which then is replaced by its mean value nks. In this process it turns out that the remaining two operators combine to c^k+qc^k-like factors to construct |ψq. This helps then to "close" the Schrödinger equation for the wave function fk. We replace nks with the Fermi-Dirac distribution for electrons of spin s with the energy ϵks.
We insert this approximation into the Schrödinger equation (written above) and multiply from the left with ψk~,q|=ψg|c^k~c^k~+q to obtain the projected equation,

fk~{ωqϵk~+q+ϵk~U(nn)}=UΩkfk(nk+qnk)=Rq

where we used that the density of electrons of spin s is given by

ns=1Ωknks.

and we defined Rq. We solve now the equation for the wave function fk and obtain

fk=Rqωqϵk+q+ϵk

with ϵks=ϵk+Uns¯. We use now fk to re-express Rq which yields the equation,

Rq=URq1Ωknk+qnkωqϵk+q+ϵk

which can be used to determine ωq for Rq0. The solutions contain, in principle, also the continuum of the electron-hole excitations discussed above. We focus, however, now on the non-trivial collective mode. It is easy to see that in the q0 limit we find

1=U(nn)ω0U(nn)ω0=0

indicating that there are excitations for small q. Note that for q=0 we find fk= const. independent of k which we will use later.
We calculate now the dispersion for small q(kF). For the concrete calculation we will assume that ϵk=2k2/2m to obtain analytical results. Using the fact that ωqΔ we can expand the equation leading to

ωq[1+UΩknk+qnkϵk+qϵk][UΩknk+qnk(ϵk+qϵk)2]1.

For small q we expand up to order q2 (actually q-linear contributions cancel),

UΩknk+qnkϵk+qϵk+ΔUΩknk+qnkΔ{1ϵk+qϵkΔ+(ϵk+qϵkΔ)2}1+UΩk{nk+nkΔ22q22mnknkΔ2(22kq)24m2Δ}1+2q22mΔ2(Un04ϵF3)

and analogously

UΩknk+qnk(ϵk+qϵk+Δ)21Δ

such that we obtain

ωq2q22mΔ2ϵF3(UN(ϵF)2)2q22m(UUc1)1/2.

Where we use that Δ(UUc)1/2 for UUc. Hence, if U>Uc=2/N(ϵF) we find ωq>0. Since the elementary excitations have an energy gap of the order of Δ at small q, the collective excitations, which are termed magnons, are well-defined quasiparticles describing propagating spin waves. When these modes enter the electron-hole continuum, they are damped in the same way as plasmons decay into the electron-hole continuum.

Being a bound state composed of an electron and a hole, magnons are bosonic quasiparticles. This collective excitation features a q2-dependent dispersion with a vanishing excitation energy in the limit q0. This property represents a case of the so-called Goldstone theorem which states that there is a gapless q=0 excitation in every ordered phase which originates from the spontaneous breaking of a continuous symmetry. Remark: It is important that the symmetry is broken spontaneously and not by an external field, because we need a continuous set of ground states, here through spin rotation. With an external field we would have only one ground state and all other orientations of the magnetisation would have higher energy. Then the q=0 mode would have a finite energy. In our case the continuous symmetry is spin rotation and the ferromagnetic state breaks this symmetry by choosing spontaneously the direction of magnetisation. We defined this direction as the z-axis such that the total spin z-component is Stot z=12(NN). Any of the 2Stot +1 degenerate ground states can be reached by continued application of the spin lowering operator S^tot which lowers Stot z by 1,

S^totz|ψg=Stotz|ψgS^tot2|ψg=2Stot(Stot+1)|ψg}{[S^tot2,S^tot]=0,|ψg=S^tot|ψgS^tot2|ψg=2Stot(Stot+1)|ψg

The spin lowering operator can be written as

S^tot =Ωkc^kc^k

which notably is the operator generating |ψq=0, where, as mentioned, fk=1/Ω is independent of k. From the equation before, we understand that |ψq=0 is simply another ground state |ψg and has, thus, the same energy as |ψg. We conclude, therefore, that ωq=0=0.