Chap 6
Chap 6
Chap 6
How best to approximate the effects of electron–electron interactions in the exact Hamiltonian H within the Born–
Oppenheimer approximation ?
Very often, an effective one-electron Schrödinger-like equation of the form Eq. (6.1) is obtained by seeking the ground-state
energy and wavefunction of an interacting many-electron system using a variational principle approach.
In the Hartree–Fock (HF) treatment: the ground-state many electron wavefunction is of the form of a single Slater determinant
6 Introduction: Hartree-Fock Approximation
Here N is the number of electrons in the system, and ri and σi denote the spatial and spin coordinates of the ith electron,
respectively. φλ are single-particle orbitals to be determined in the calculation.
Applying the variational principle, i.e. demanding the variation of the ground-state energy with respect to any of the
single-particle orbitals φλ to be zero,
6 Introduction: Hartree-Fock Approximation
Here N is the number of electrons in the system, and ri and σi denote the spatial and spin coordinates of the ith electron,
respectively. φλ are single-particle orbitals to be determined in the calculation.
Applying the variational principle, i.e. demanding the variation of the ground-state energy with respect to any of the
single-particle orbitals φλ to be zero,
6 Introduction: Hartree-Fock Approximation
Here N is the number of electrons in the system, and ri and σi denote the spatial and spin coordinates of the ith electron,
respectively. φλ are single-particle orbitals to be determined in the calculation.
Applying the variational principle, i.e. demanding the variation of the ground-state energy with respect to any of the
single-particle orbitals φλ to be zero,
H. W. derivation
Hartree where
Fock (exchange)
6 Introduction: Hartree-Fock Approximation
Fock (exchange) term
The exchange term arises from the antisymmetric form of the many-fermion wavefunction, which introduces an effective
interaction between electrons owing to the Pauli exclusion principle.
It is a nonlocal operator of the integral form when acting on a single-particle orbital φi(r, σ), i.e.
The eigenvalues εi, for either occupied or empty orbitals, naturally seem to have the interpretation of eigenenergies of the
original Hamiltonian, namely the total energy of the system is approximated as ∑𝑖𝑖 𝜖𝜖𝑖𝑖 𝑛𝑛𝑖𝑖 , where ni is occupation number of
orbital i.
The eigenvalue εi in the Hartree–Fock equation (Eq. (6.5)) is a Lagrange multiplier in the variational procedure, constraining
the normalization of the single-particle orbitals. Thus, rigorously speaking, they are not electron excitation energies of the
system. However, through the use of Koopmans’ theorem, these eigenvalues may be roughly interpreted as electron
excitation energies to lowest order in the Coulomb interaction.
6.1 The homogeneous interacting electron gas
The uniform interacting electron gas (the jellium model)
In this model, we consider a system of interacting electrons in a uniform positive-charged background so that the total
system is neutral. (an idealized metal)
In the zero-order approximation, we neglect electron–electron interactions; the jellium model reduces to that of a non-
interacting charged Fermi gas, i.e. the Sommerfeld model.
The value of k is within the Fermi sphere, i.e. k ≤ kF, where the factor 2 is due to spin.
6.1 The homogeneous interacting electron gas
The uniform interacting electron gas (the jellium model)
𝑁𝑁
Only one variable physical parameter, the density of the electrons 𝑛𝑛 = or 𝑘𝑘𝐹𝐹 = 3𝜋𝜋 2 𝑛𝑛 1/3
H. W. derivation
Ω
the dimensionless electron gas parameter the effective radius of the volume occupied by
the electron in units of the Bohr radius
H. W. derivation
is a measure of the ratio of the average potential energy to the kinetic energy in the electron gas.
Coulomb energy
A system with rs < 1 corresponds to one with high density and dominant kinetic energy; a system with rs > 1 corresponds to
one with low density and dominant potential energy.
6.2 Hartree–Fock treatment of the interacting electron gas
The single-particle orbitals are planewaves times a spin function, i.e. the orbitals are specified by the quantum
numbers k and λ.
plane wave
Spin function
We note that planewave solutions result in a uniform electron charge density. This results in Vion(r) + VH = 0,
since the electron charge density cancels the positively charged background density.
6.2 Hartree–Fock treatment of the interacting electron gas
The planewaves φkλ are solutions to the Hartree–Fock equation in the case of the uniform electron gas.
with
6.2 Hartree–Fock treatment of the interacting electron gas
The planewaves φkλ are solutions to the Hartree–Fock equation in the case of the uniform electron gas.
with
𝑘𝑘𝐹𝐹 1
2𝜋𝜋 4𝜋𝜋𝑒𝑒 2
−� 𝑑𝑑𝑑𝑑𝑑 � 𝑑𝑑cos𝜃𝜃 3 ′ 2
𝑘𝑘𝑘2
0 −1 2𝜋𝜋 𝑘𝑘 − 𝑘𝑘
𝑒𝑒 2 𝑘𝑘𝐹𝐹 1
𝑘𝑘𝑘2
=− � 𝑑𝑑𝑑𝑑𝑑 � 𝑑𝑑cos𝜃𝜃 2 Let 𝑎𝑎 = 𝑘𝑘 2 + 𝑘𝑘𝑘2 and 𝑏𝑏 = 2𝑘𝑘𝑘𝑘 ′
𝜋𝜋 0 −1 𝑘𝑘 + 𝑘𝑘𝑘2 − 2𝑘𝑘𝑘𝑘 ′ cos𝜃𝜃
𝑒𝑒 2 𝑘𝑘𝐹𝐹 2
1
1 1 1 1 1 1 1
=− � 𝑑𝑑𝑑𝑑𝑑𝑘𝑘𝑘 � 𝑑𝑑cos𝜃𝜃 ∫−1 𝑑𝑑cos𝜃𝜃 𝑎𝑎−𝑏𝑏cos𝜃𝜃= −𝑏𝑏 ln(𝑎𝑎 − 𝑏𝑏cos𝜃𝜃) = ln(𝑎𝑎 − 𝑏𝑏) + ln(𝑎𝑎 + 𝑏𝑏)
𝜋𝜋 0 −1 𝑎𝑎 + 𝑏𝑏cos𝜃𝜃 −1 −𝑏𝑏 𝑏𝑏
𝑒𝑒 2 𝑘𝑘𝐹𝐹 2
1 2 + 𝑘𝑘 ′ 2 + 2𝑘𝑘𝑘𝑘 ′ ) −ln(𝑘𝑘 2 + 𝑘𝑘𝑘2 − 2𝑘𝑘𝑘𝑘 ′ )
=− � 𝑑𝑑𝑑𝑑𝑑𝑘𝑘𝑘 ln(𝑘𝑘
𝜋𝜋 0 2𝑘𝑘𝑘𝑘 ′
𝑒𝑒 2 𝑘𝑘𝐹𝐹 1
=− � 𝑑𝑑𝑑𝑑𝑑𝑘𝑘𝑘2 ′ ln 𝑘𝑘 + 𝑘𝑘 ′ 2 −ln
𝑘𝑘 − 𝑘𝑘 ′ 2 )
𝜋𝜋 0 2𝑘𝑘𝑘𝑘
6.2 Hartree–Fock treatment of the interacting electron gas
The planewaves φkλ are solutions to the Hartree–Fock equation in the case of the uniform electron gas.
with
𝑘𝑘𝐹𝐹 1
2𝜋𝜋 4𝜋𝜋𝑒𝑒 2
−� 𝑑𝑑𝑑𝑑𝑑 � 𝑑𝑑cos𝜃𝜃 3 ′ 2
𝑘𝑘𝑘2
0 −1 2𝜋𝜋 𝑘𝑘 − 𝑘𝑘
𝑒𝑒 2 𝑘𝑘𝐹𝐹 2
1
=− � 𝑑𝑑𝑑𝑑𝑑𝑘𝑘𝑘 ′
ln 𝑘𝑘 + 𝑘𝑘 ′ 2 −ln 𝑘𝑘 − 𝑘𝑘 ′ 2 )
𝜋𝜋 0 2𝑘𝑘𝑘𝑘
𝑒𝑒 2 𝑘𝑘𝐹𝐹 𝑘𝑘𝑘 𝑥𝑥 2 −𝑎𝑎2 1
= � 𝑑𝑑𝑑𝑑𝑑 − ln k + k′ +ln k − k′ using ∫ 𝑥𝑥ln 𝑥𝑥 + 𝑎𝑎 𝑑𝑑𝑑𝑑 = ln 𝑥𝑥 + 𝑎𝑎 − (𝑥𝑥 − 𝑎𝑎)2
𝜋𝜋 0 𝑘𝑘 2 4
𝑘𝑘𝐹𝐹
𝑒𝑒 2 𝑘𝑘 ′2 − 𝑘𝑘 2 𝑘𝑘 − 𝑘𝑘 ′ 1 ′ 2+
1
= ln| | − 𝑘𝑘 + 𝑘𝑘 𝑘𝑘 − 𝑘𝑘 ′ 2
𝜋𝜋𝜋𝜋 2 𝑘𝑘 + 𝑘𝑘 ′ 4 4 0
The planewaves φkλ are solutions to the Hartree–Fock equation in the case of the uniform electron gas.
with
where
If one invokes Koopmans’ theorem, then the eigenvalues are a rough estimate of the electron
(or quasiparticle) excitation energy of the system.
The eigenvalues describe the energy for exciting a single-particle-like excitation within the
Hartree–Fock approximation, including only the exchange interaction.
the self energy of the excited particle
with
6.2 Hartree–Fock treatment of the interacting electron gas
The planewaves φkλ are solutions to the Hartree–Fock equation in the case of the uniform electron gas.
with
where
with
with
F′(1)= ∞
This divergence leads to behaviors for the electrons that are peculiar to the Hartree–Fock results.
∗
the effective mass 𝑚𝑚HF → 0 as k → kF.
6.2 Hartree–Fock treatment of the interacting electron gas
Limitation of the HF approximation
with
F′(1)= ∞
This divergence leads to behaviors for the electrons that are peculiar to the Hartree–Fock results.
- An electronic specific heat that behaves like Cel(T) ∼ T/ln|T| instead of a linear T
dependence at low temperature T
6.2 Hartree–Fock treatment of the interacting electron gas
Limitation of the HF approximation
with
F′(1)= ∞
This divergence leads to behaviors for the electrons that are peculiar to the Hartree–Fock results.
The above unphysical behavior can be traced back to the divergence of the Fourier
4𝜋𝜋𝑒𝑒 2 𝑒𝑒 2
transform of the Coulomb interaction , as k goes to zero.
Ω𝑘𝑘 2 𝑟𝑟
𝑒𝑒 2
Physically, in a metallic system, will be screened by the other electrons, especially at
𝑟𝑟
large values of r. A screened Coulomb potential would eliminate the divergence as k → 0.
Thomas–Fermi screening
(see Chapter 8)
The ground-state energy per particle EKE: the kinetic energy per particle
𝑘𝑘
using � sin𝜃𝜃𝜃𝜃𝜃𝜃𝜃𝜃𝜃𝜃 = 4𝜋𝜋, Let = 𝑥𝑥
𝑘𝑘𝐹𝐹
𝑛𝑛 = 𝑘𝑘𝐹𝐹3 /3𝜋𝜋 2
=1/4
6.3 Ground-state energy: Hartree–Fock and beyond
The Hartree–Fock approximation, being a variational approach for the ground-state wavefunction, is expected to
be more accurate for the ground-state energy as opposed to individual single-particle excitation energies.
The ground-state energy per particle EKE: the kinetic energy per particle
1 1 2)
1 (1 − 𝑥𝑥 1 + 𝑥𝑥 𝐼𝐼
� 𝑥𝑥 2 𝐹𝐹 𝑥𝑥 𝑑𝑑𝑑𝑑 = � 𝑥𝑥 2 + 𝑥𝑥ln 𝑑𝑑𝑑𝑑
0 0 2 4 1 − 𝑥𝑥
1 1 1 2
1 + 𝑥𝑥
= + � (1 − 𝑥𝑥 )𝑥𝑥ln 𝑑𝑑𝑑𝑑
6 4 0 1 − 𝑥𝑥
1 1
= + 𝐼𝐼
6 4
using 1
1
� 𝑥𝑥 2 𝐹𝐹 𝑥𝑥 𝑑𝑑𝑑𝑑 =
0 4
6.3 Ground-state energy: Hartree–Fock and beyond
The Hartree–Fock approximation, being a variational approach for the ground-state wavefunction, is expected to
be more accurate for the ground-state energy as opposed to individual single-particle excitation energies.
The ground-state energy per particle EKE: the kinetic energy per particle
We expect that a more accurate theoretical treatment of the many-body wavefunction, going
beyond Hartree–Fock, would result in more terms.
In the limit of small rs (i.e. high density), the kinetic energy dominates, and we may expect the ground-state energy per particle
(and other quantities) to be expandable in a series in powers of rs.
The ground-state energy per particle EKE: the kinetic energy per particle
Many-body perturbation theory is accurate in the limit of rs < 1. Correlation energy of the interacting electron gas as
a function of rs from different calculations.
6.3 Ground-state energy: Hartree–Fock and beyond
One of the simplest and yet highly accurate expressions for EC is the Wigner interpolation formula. (low density, i.e. rs >>1 .)
In this limit, the potential energy dominates over the kinetic energy and the electrons condense into an electron crystal in three
dimensions at rs ≥ 100.
To calculate the ground-state energy of the electron crystal, let us divide the crystal into Wigner–Seitz cells.
The energy per electron is the energy of the electron in one of the Wigner–Seitz spheres, which consists of a kinetic energy term
EKE and a potential energy term EPE.
- EKE comes from the zero-point motion of the electron and scales like rs−3/2, which may be neglected in the rs → ∞ limit.
Ee−p is the interaction energy between the electron at the center of the Wigner–Seitz sphere and the positive
neutralizing background density np = 3/(4π rs3),
6.3 Ground-state energy: Hartree–Fock and beyond
One of the simplest and yet highly accurate expressions for EC is the Wigner interpolation formula. (low density, i.e. rs >>1 .)
In this limit, the potential energy dominates over the kinetic energy and the electrons condense into an electron crystal in three
dimensions at rs ≥ 100.
To calculate the ground-state energy of the electron crystal, let us divide the crystal into Wigner–Seitz cells.
The energy per electron is the energy of the electron in one of the Wigner–Seitz spheres, which consists of a kinetic energy term
EKE and a potential energy term EPE.
- EKE comes from the zero-point motion of the electron and scales like rs−3/2, which may be neglected in the rs → ∞ limit.
Ep−p, is the self energy of the positive neutralizing background charge, given by
6.3 Ground-state energy: Hartree–Fock and beyond
One of the simplest and yet highly accurate expressions for EC is the Wigner interpolation formula. (low density, i.e. rs >>1 .)
In this limit, the potential energy dominates over the kinetic energy and the electrons condense into an electron crystal in three
dimensions at rs ≥ 100.
To calculate the ground-state energy of the electron crystal, let us divide the crystal into Wigner–Seitz cells.
The energy per electron is the energy of the electron in one of the Wigner–Seitz spheres, which consists of a kinetic energy term
EKE and a potential energy term EPE.
- EKE comes from the zero-point motion of the electron and scales like rs−3/2, which may be neglected in the rs → ∞ limit.
Using the fact that EPE = EX + EC and the Hartree–Fock result of EX = −0.916/rs,
6.3 Ground-state energy: Hartree–Fock and beyond
One of the simplest and yet highly accurate expressions for EC is the Wigner interpolation formula. (low density, i.e. rs >>1 .)
In this limit, the potential energy dominates over the kinetic energy and the electrons condense into an electron crystal in three
dimensions at rs ≥ 100.
To calculate the ground-state energy of the electron crystal, let us divide the crystal into Wigner–Seitz cells.
The energy per electron is the energy of the electron in one of the Wigner–Seitz spheres, which consists of a kinetic energy term
EKE and a potential energy term EPE.
- EKE comes from the zero-point motion of the electron and scales like rs−3/2, which may be neglected in the rs → ∞ limit.
Using the fact that EPE = EX + EC and the Hartree–Fock result of EX = −0.916/rs,
Using an estimate of EC (rs = 1)=−0.10 Ry from experiment, Wigner arrived at a simple interpolation formula of
6.3 Ground-state energy: Hartree–Fock and beyond
One of the simplest and yet highly accurate expressions for EC is the Wigner interpolation formula. (low density, i.e. rs >>1 .)
In this limit, the potential energy dominates over the kinetic energy and the electrons condense into an electron crystal in three
dimensions at rs ≥ 100.
To calculate the ground-state energy of the electron crystal, let us divide the crystal into Wigner–Seitz cells.
The energy per electron is the energy of the electron in one of the Wigner–Seitz spheres, which consists of a kinetic energy term
EKE and a potential energy term EPE.
Using an estimate of EC (rs = 1)=−0.10 Ry from experiment, Wigner arrived at a simple interpolation formula of
The Wigner interpolation formula has been proved to be accurate and useful in many applications, and although many other
correlation energy formulas have been derived over the years, the Wigner interpolation formula remains among the best.
6.4 Electron density and pair-correlation functions
An N-body fermion wavefunction
g(r, r')ρ(r') is the density of electrons at r', given that there is one electron at r.
The sum rule: Integrating it over all volume gives the total number of
pairs in the system, which is N(N − 1)/2
the integrand is the distribution of particle pairs
(N − 1) − N = −1
6.5 g (r, r') of the interacting electron gas
For a homogeneous system like an electron gas, the pair-correlation function g(r, r') depends only on the distance
between r and r' and not on the positions separately:
We now evaluate these correlation functions for the interacting electron gas within the Hartree–Fock approximation.
For simplicity, we work with a paramagnetic electron gas, i.e. the number of spin-up electrons is equal to the number
of spin-down electrons.
(N-2)!
Use of one-electron orthonormal orbitals in the Hartree–Fock Slater determinant
Thus,
Four combinations for (si, sj): (↑, ↑), (↓, ↓), (↑, ↓), and (↓, ↑)
Thus,
𝐤𝐤 � 𝐫𝐫 = 𝑘𝑘𝑘𝑘cos𝜃𝜃
the first Bessel function
6.5 g (r, r') of the interacting electron gas
where
𝐻𝐻𝐻𝐻
The extent over which 𝑔𝑔↑↑ deviates from 1/2 is dictated by 1/kF.
The difference between the exact g and gHF leads to a change in the ground-state energy and gives rise to
the correlation energy.
6.6 The exchange-correlation hole
The electron density at position r' as seen by an electron at position r is given by
The difference between the electron density seen by an electron at r, ρr(r'), and that of the electron density ρ(r') measured,
for example, in an X-ray scattering experiment, is given by
Integrating over volume and making use of the sum rule yields
The electron density ρr(r') seen by a given electron at r is always different from that of ρ(r'). For example, in an
interacting homogeneous electron gas, although ρ(r') is a constant, ρr(r') is highly inhomogeneous.
The evaluation of the pair-correlation function and the exchange-correlation hole for a real material is non-trivial
since it requires knowing the many-electron ground-state wavefunction. reasonable description of g↑↓(r, r') means
incorporation of correlation effects going beyond Hartree–Fock. →variational quantum Monte Carlo simulations
r r
A suppression from the value of ½ for g↑↑ , as r → r' The suppression of g↑↓ is not nearly as pronounced because
showing a deep exchange hole diamond is only a moderately correlated electron system.
6.7 The exchange-correlation energy
The exchange-correlation energy of an electron is the amount of lowering of the energy of the electron, as the many-body
wavefunction describes more correlations of the motions of the electron as we improve on the Hartree approximation (that is,
the energy gain due to both exchange and correlation effects).
We may calculate this energy difference by evaluating the potential energy using a charge density around an electron at r,
with and without the exchange-correlation hole contribution.
The energy difference, which is the exchange-correlation energy for the electron at r,
the interaction energy of the electron with its exchange-correlation hole which can be considered positively
charged with an integrated charge of −e
6.7 The exchange-correlation energy
The quantity εxc(r) is also called the exchange-correlation energy density of the system.
We may apply this concept to the case of the interacting electron gas within the Hartree– Fock approximation. Since the
system is homogeneous, εxc(r) is the same everywhere and is the exchange-correlation energy per particle.
𝑘𝑘𝐹𝐹3 /3𝜋𝜋 2
6.7 The exchange-correlation energy
The quantity εxc(r) is also called the exchange-correlation energy density of the system.
We may apply this concept to the case of the interacting electron gas within the Hartree– Fock approximation. Since the
system is homogeneous, εxc(r) is the same everywhere and is the exchange-correlation energy per particle.
𝑘𝑘𝐹𝐹3 /3𝜋𝜋 2
6.7 The exchange-correlation energy
The quantity εxc(r) is also called the exchange-correlation energy density of the system.
We may apply this concept to the case of the interacting electron gas within the Hartree– Fock approximation. Since the
system is homogeneous, εxc(r) is the same everywhere and is the exchange-correlation energy per particle.
𝑘𝑘𝐹𝐹3 /3𝜋𝜋 2
A particularly interesting situation is the case in which the electron leaves the crystal and stays near the surface. The
exchange-correlation hole by necessity stays behind in the crystal
The Coulomb interaction between the positively charged exchange-correlation hole and the electron that leaves the crystal
results in an attraction that scales like the inverse of the distance from the surface, giving rise to the classical image force.
6.7 The exchange-correlation energy
The exchange-correlation energy plays an important role in determining conceptually and quantitatively many other
properties of solids.
Eatom: the binding energy of the outer active electrons in the atom for the isolated case
EGS: the energy per atom in the solid state
i)
the conduction band effective mass
Eatom: the binding energy of the outer active electrons in the atom for the isolated case
EGS: the energy per atom in the solid state
We see that the alkali metals are unbounded within the Hartree only approximation.
Both exchange and correlation effects are needed to give a reasonable description of
the cohesive properties of this simplest class of metals.