Chapter 4 - Density-Functional Theory
Chapter 4 - Density-Functional Theory
Density‐Functional Theory
(a)
Quantum
Input Output
ESPRESSO
Structures Properties
of materials of materials
HΨ = Etot Ψ, (4.1)
where H is the Hamiltonian of the system, Etot is the total energy of the
electrons and nuclei, and Ψ(r1 , …, rNe , R1 , …, RNn ) is the wavefunction
of many particles (electrons and nuclei). By using the atomic units
listed in Table 4.1, a general Hamiltonian in Eq. (4.1) is given by
kinetic energies T and potential energies V of the electrons and
nuclei as
H = Tn + Vn + Te + Ve + Ven
ZI ZJ
Nn Nn
∇2RI 1
2MI 2 |RI − RJ |
=− +
I=1 I̸=J
(4.2)
nuclei
Ne Ne Ne Nn
∇2r 1 1 −ZI
.
i
2 2 |ri − rj | |ri − RI |
− + +
i=1 i̸=j i=1 I=1
electrons mixed
is expressed by
h̄2 ∇2ri e2
Ne Ne
1
. (4.3)
2me 2 4πϵ0 |ri − rj |
− +
i=1 i̸=j
e–
r1 |r1 – r2|
e–
r2
2e+
x y
1 1 1 2 2
− ∇2r1 − ∇2r2 + ψ(r1 , r2 )
2 2 |r1 − r2 | |r1 | |r2 | (4.5)
− −
= Eψ(r1 , r2 ).
Ne Ne
Nn
∇2r
ψ = Eψ.
−ZI
(4.6)
i
|ri − RI |
(Te + Ven )ψ =
2
− +
i=1 i=1 I=1
N
∇2r n
h(r) = −
−ZI
, (4.7)
2 |r − RI |
+
I=1
From Eq.(4.8) we can say that, for the case of the non‐interacting
electrons, the Hamiltonian of the system can be written as Ne
independent single‐particle Hamiltonian, and the wavefunction
is written by ψ(ri ). Eq. (4.8) is used to obtain the energy of
electrons for the following two cases: (1) distinguishable4 and (2)
indistinguishable5 electrons. For simplicity, we consider the two
electrons in a helium atom as follows:
(1) Two distinguishable electrons: Let us consider electron 1
in one state ϕa (r1 ) and electron 2 in another state ϕb (r2 ). Since two
electrons are independent of each other, the probability of finding
(electron 1 at r1 ) and (electron 2 at r2 ) can be given by the product
as |ψ(r1 , r2 )|2 = |ϕa (r1 )|2 |ϕb (r2 )|2 . In this case, the wavefunction can
4 If the electrons are distinguishable, the physical properties of the system are
[h(r1 ) + h(r2 )] ϕa (r1 )ϕb (r2 ) = Eϕa (r1 )ϕb (r2 ). (4.9)
where ϵα is the energy of one electron, the left‐hand side of Eq. (4.9)
is rewritten as
E = ϵa + ϵb . (4.12)
Thus, for the case of the non‐interacting system with the distinguish‐
able electrons, the total energy of the electrons is the sum of the
energy of each electron.
(2) Two indistinguishable electrons: In quantum mechanics,
two electrons are indistinguishable (i.e., ϕa (r1 )ϕb (r2 ) is equivalent to
ϕa (r2 )ϕb (r1 )). Thus, one possible shape of the wavefunction ψ(r1 , r2 )
is a linear combination of two states as [Kittel (1976)]
1
ψ(r1 , r2 ) = √ [ϕa (r1 )ϕb (r2 ) − ϕb (r1 )ϕa (r2 )], (4.13)
2
where P1 (r) = |ψ(r, r2 )|2 dr2 and P2 (r) = |ψ(r1 , r)|2 dr1 are prob‐
By substituting Eq. (4.13) into Eq. (4.16) and using the normalization
and orthogonality conditions, respectively, as
ϕ∗a (r)ϕa (r)dr = ϕ∗b (r)ϕb (r)dr = 1 (4.17)
and
ϕ∗b (r)ϕa (r)dr = ϕ∗a (r)ϕb (r)dr = 0, (4.18)
Z3/2 Z2
ϕ(r) = √ exp(−Z|r|) and ϵ = − , (4.20)
π 2
2Z3
n(r) = exp(−2Z|r|) and E = −Z2 , (4.21)
π
16
n(r) = exp(−4|r|) (4.22)
π
He+ + e− −
I I
He −
→1
→2
He++ + 2e− , (4.23)
1
h(r1 ) + h(r2 ) + ϕ (r1 )ϕb (r2 ) = Eϕa (r1 )ϕb (r2 ). (4.24)
|r1 − r2 | a
with
ϵHb = E − ϵa , (4.28)
n(r′ )
dr′ . (4.33)
|r − r′ |
VH (r) =
occupied states. We repeat the procedure until n(r) does not change
180 | Density‐Functional Theory
No End
Figure 4.4 Flow‐chart of the self‐consistent field method for the solution of
the single‐particle Schrödinger equation with the Hartree potential VH .
1 ∂ ∂
2
r 2
r ∂r ∂r
∇ = 2
(4.34)
1 ∂ ∂ 1 ∂2
.
r2 sin θ ∂θ ∂θ r2 sin θ ∂φ2
+ sin θ + 2
1 ∂ 2 ∂VH
r (4.35)
r2 ∂r ∂r
= −4πn(r).
r
1 r′
C
r′′ n(r′′ )dr′ dr′′ −
. (4.36)
2
r′2 r
VH (r) = −4π
0 0
It is noted that C/r in Eq. (4.36) is a solution of Eq. (4.35) for any value
of C. Then by substituting the electron density of the helium atom in
Eq. (4.21) into Eq. (4.36) and using
x
1 x′
2
x′′ exp(x′′ )dx′ dx′′ = (1 −
(4.37)
2
x′2 x
) exp(x),
0 0
1 C
VH (r) = −2 Z + . (4.38)
r r
exp(−2Zr) −
1
r n(r′ )dr′ = Zr . Applying this boundary condition for Eq. (4.38), we
obtain C = −Z, and Eq. (4.38) can be rewritten for the helium atom
182 | Density‐Functional Theory
Figure 4.5 The Hartree potential VH (r) (solid line), the Coulomb attraction
between electrons and nuclei Ven (dotted line), and the total potential Ven +
VH (r) (dashed line) of helium atom are plotted in atomic unit as functions of
r.
as
1 Z
VH (r) = −2 Z + exp(−2Zr) + . (4.39)
r r
1
h(r1 ) + h(r2 ) +
|r1 − r2 |
[ϕa (r1 )ϕb (r2 ) − ϕb (r1 )ϕa (r2 )]
(4.40)
= E[ϕa (r1 )ϕb (r2 ) − ϕb (r1 )ϕa (r2 )].
By multiplying ϕ∗b (r2 ) in Eq. (4.40) and integrating over r2 with the
normalization and orthogonality conditions (Eqs. (4.17) and (4.18)),
we obtain:
where ϵHF
a is defined by
γ(r2 , r1 )
a ϕa (r1 ), (4.43)
ϕ (r2 )dr2 = ϵHF
|r1 − r2 | a
[h(r1 ) + VH (r1 )] ϕa (r1 ) −
γ(r2 , r1 ) = ϕ∗a (r2 )ϕa (r1 ) + ϕ∗b (r2 )ϕb (r1 ). (4.44)
γ(r1 , r2 )
b ϕb (r2 ), (4.45)
ϕ (r1 )dr1 = ϵHF
|r1 − r2 | b
[h(r2 ) + VH (r2 )] ϕb (r2 ) −
in which ϵHF
b is symmetrically given by
b = E − ϵa .
ϵHF (4.46)
γ(r2 , r1 )
Vx (r1 , r2 ) = − . (4.47)
|r1 − r2 |
Vx exists only for two electrons at r1 and r2 , which have the same
quantum state ϕa or ϕb including spin, i.e., the exchange potential
is the effective Coulomb interaction between the electrons with
parallel spins in the system. The origin of Vx is that two‐electrons
with the same spin can not exist at the same position (r1 = r2 ), in
which VH should be small by the restriction of the Pauli principle. By
considering both VH and Vx in Eq. (4.39), two electrons are interacted
not only by their electronic charge but also by their spins. Therefore,
the quantum behavior of electrons is included by Vx term in the
Hartree‐Fock equation. It is important to note that Vx is a non‑local
potential on r1 since it depends on an integration over the additional
variable r2 . Therefore, the practical solution of the Hartree‐Fock
equation becomes more complicated and time‐consuming.
Quantum ESPRESSO Course for Solid‐State Physics | 185
Let us discuss the energy that is obtained by the Hartree‐Fock
equation. From Eq. (4.43), we can write the energy by integrating the
wavefunctions as
ϵHF
a = ϕ∗a (r1 )[h(r1 ) + VH (r1 )]ϕa (r1 )dr1
γ(r2 , r1 ) (4.48)
ϕ (r2 )dr1 r2
|r1 − r2 | a
+ ϕ∗a (r1 )
=ϵa + J − K,
and
K=
dr1 dr2 . (4.50)
ϕ∗a (r1 )ϕb (r1 )ϕ∗b (r2 )ϕa (r2 )
|r1 − r2 |
E = ϵa + ϵb + J − K. (4.51)
186 | Density‐Functional Theory
On the other hand, from Eq. (4.40), the total electron energy can be
obtained, too, as
1
ψ∗ (r1 , r2 ) ψ(r1 , r2 )dr1 r2
(4.52)
|r1 − r2 |
+
|ψ(r2 , r1 )|2
dr1 r2 .
|r1 − r2 |
= ϵa + ϵb +
|ψ(r1 , r2 )|2
J−K=
dr1 dr2 ≥ 0. (4.53)
|r1 − r2 |
Total energy of helium atom: In the case of the helium atom, two
electrons at the 1s state ϕ must have different spin directions, up ↑
and down ↓ because of the Pauli principle. Let us consider the spin
states of two electrons, respectively, as
1 0
and ϕ↓ (r) = ϕ(r) , (4.54)
0 1
ϕ↑ (r) = ϕ(r)
1
ϕ∗↑ (r)ϕ↑ (r) = ϕ2 (r) 1 0 = ϕ2 (r), (4.55)
0
and
0
ϕ∗↑ (r)ϕ↓ (r) = ϕ2 (r) 1 0 = 0. (4.56)
1
Quantum ESPRESSO Course for Solid‐State Physics | 187
By substituting Eqs. (4.55) and (4.56) into Eqs. (4.49) and (4.50),
respectively, we obtain:
2 2
J=
drdr′ ,
ϕ∗↑ (r)ϕ↑ (r)ϕ∗↓ (r′ )ϕ↓ (r′ ) |ϕ(r)| |ϕ(r′ )|
|r − r′ |2 |r − r′ |2
drdr′ =
(4.57)
and
K=
drdr′ = 0. (4.58)
ϕ∗↑ (r)ϕ↓ (r)ϕ∗↓ (r′ )ϕ↑ (r′ )
|r − r′ |2
1 |ϕ(r)|2 n(r′ ) 1
J=
|ϕ(r)|2 VH (r)dr. (4.59)
2 |r − r′ |2 2
drdr′ =
1
E = 2ϵ +
|ϕ(r)|2 VH (r)dr. (4.60)
2
In order to apply Eq. (4.60) for the helium atom, we define an effective
atomic number Z, i.e., ϕ(r, Z), which gives minimized E when subject
to the Hartree potential VH (r). Eq. (4.60) can be obtained by the
following steps:
• Step 1: Using a beginning value Z = 2, from Eq. (4.39), the
Hartree potential VH (r) is given by
1 2
VH (r) = −2 2 + exp(−4r) + . (4.61)
r r
188 | Density‐Functional Theory
E = 2ϵ + J − K
Eexp = − 2.9 Ha
E = 2ϵ
6 The variational principle states that the energy of any approximate wavefunction is
higher than or equal to the energy of ground‐state wavefunction. The lower the
energy is the better the approximation to the ground state.
Quantum ESPRESSO Course for Solid‐State Physics | 189
could be done by recalculating the Hartree potential with new state
ϕ(r, Zeff = 1.826) in Step 1, then we recalculate Steps 2 and 3 until
the value of Zeff does not change significantly. Finally, we can obtain
the Eeff = −2.85 Ha, which is not too far from the exact value from the
experiment about −2.9 Ha. Slater derived a formula to determine Zeff
as Zeff = Z − S, where S is the shielding constant, which is known as
Slater’s rule [Slater (1930a)]. For the 1s state in the helium atom,
S = 0.3 and Zeff = 1.7, which is close to the result of the SCF method
(Zeff = 1.826).
In Sec. 4.6, we show that due to the Pauli exclusion principle, the
electrons with parallel spins have an exchange term of the Coulomb
interaction Vx . The two electrons with the parallel spin move to avoid
each other because the electrons can not overlap. Avoiding picture
should exist for the two electrons with anti‐parallel spins because
these electrons also move with keeping apart to lower the Coulomb
repulsion. This behavior of the electrons with anti‐parallel spins is
missing in the Hartree approximation as well as in the Hartree‐Fock
approximation. A complete picture of the motion of an electron in a
system is shown in Fig. 4.7. In general, the correlation interaction,
Vc , is defined by the remaining term that is missing in the Hartree‐
Fock method.7 Thus, an effective potential in the single‐particle
Schrödinger equation is defined by
∇2r
(4.64)
2
− + Veff [n(r)] ϕ(r) = ϵϕ(r).
Vx
VH Vc
Vc
When we know the value of Vxc [n(r)], Eq. (4.64) can be solved by
using the SCF method as discussed in Sec. 4.5. However, we do not
know the exact shape of Vxc [n(r)]. Many people proposed the many
functionals for Vxc [n(r)] over the past few decades. The efforts to
develop the accurate approximations for Vxc [n(r)] led to a theory,
the so‐called density‐functional theory (DFT). We will discuss these
approximations in the next section.
Let us consider how to obtain Vxc [n(r)] for a free‐electron gas, where
the contribution of ions is treated as a uniform‐positive‐charge
background, which is called the Jellium model [Brack (1993)], as
shown in Fig. 4.8. Since the potential is uniform, the electronic states
are expressed by plane waves as (see Sec. 5.3)
1
ϕk,σ (r) = √ exp(ikr)χ σ , (4.65)
V
Quantum ESPRESSO Course for Solid‐State Physics | 191
Positive ion charges Uniform background
+ + + + + + +
+
+ + + + + + + +
+ + + + + + + +
+ + + + + + +
+
Since the unit cell can not be defined by the uniform background,
the value of k is taken not from the Brillouin zone but all the k
space. Since ϕk,σ (r) is a one‐electron wavefunction, ϕk,σ (r) can be
a solution of the Hartree‐Fock equation in Eq. (4.43). Moreover, in
the free‐electron gas, the electron density of the ground state is
also a uniform‐negative‐charge background. Therefore, the electron
density cancels the positive charge background, i.e., Ven + VH = 0
and only the exchange term Vx survives in Eq. (4.43). Thus, the single‐
particle Schrödinger equation of a free‐electron gas can be written as
Ix = −
ϕ∗k′ ,σ′ (r′ )ϕk′ ,σ′ (r)
|r − r′ |
ϕk,σ (r′ )dr′
k
(4.67)
σ ′ ′
∗
ϕk′ ,σ′ (r′ )ϕk,σ (r′ ) ′
dr .
|r − r′ |
=− ϕk′ ,σ′ (r)
′ σ′ k
192 | Density‐Functional Theory
V3/2 |r − r′ |
′
dr′
k ′
f(k − k) ϕk,σ (r),
′
= −
k′
1 1 4π
f(q) =
, (4.69)
exp(−iqx)
V V q2
dx =
|x|
1 4π
Ix = −
V ′ |k′ − k|2
ϕk,σ (r)
k
1 4π
kF
(4.70)
(2π)3 0 |k′ − k|2
= − dk ϕk,σ (r)
′
kF k
=− F
kF
ϕk,σ (r),
π
1 − x2 1 + x k
F(x) = 1 + , with x = . (4.71)
2x 1 − x kF
ln
k = kF
kF k
− F
∇r
(4.72)
kF
ϕk,σ (r) = ϵk ϕk,σ (r).
2
−
π
k2 kF k
− F . (4.73)
kF
ϵk,σ
2
=
π
By taking the sum of all occupied states, we obtain the total energy
as
k2 1 kF
k
E= F
2 2 σ kF
−
σ |k| < kF
π
|k|<kF
2V k kF V k
kF 2 kF
F
(2π)3 >0 2 π (2π)3 0 kF
= dk − dk
V k kF V k
kF 4 kF
4π dk − 4 4π k F
2
(4.74)
4π3 2 8π k
= dk
0 0 F
V 5 V 4 1 2
k − k x F(x)dx
10π2 F 2π3 F 0
=
V 5 V
=1/2
kF − 3 k4F ,
10π 2 4π
=
194 | Density‐Functional Theory
with the factor of 1/2 in the second term of the right‐hand side in the
first line of Eq. (4.74) is needed to compensate for double counting
of the exchange interaction.
For the free‐electron gas, the number of electrons Ne is equal to
the number of occupied states in the systems:
kF
V kF
k3F
Ne = k2 dk = V
= 2V . (4.75)
dk
(2π)3 3π2
= 2
σ k 0 π 0
Ne k3
n= = F2 . (4.76)
V 3π
By substituting Eqs. (4.75) and (4.76) into Eq. (4.74), the total energy
per electron is expressed by
E
= C1 n2/3 + C2 n1/3 , (4.77)
Ne
3 3 3
1/3
C1 = (3π2 )2/3 = 2.871 and C2 = − = −0.738. (4.78)
10 4 π
The first term (C1 n2/3 ) in Eq. (4.77) represents the kinetic energy, and
the second term (C2 n1/3 ) represents the effective electron‐electron
interaction due to exchange potential. We can see that the exchange
potential reduces the total energy of the system from the kinetic
energy.
For a system with the free electron gas, n is a constant. However,
for a system with the non‐uniform‐electron charge, n becomes
a function of r. Slater [Slater (1951)] introduced the exchange
potential as a function of n(r) as
3 3
1/3
Vx [n(r)] = 2C2 n1/3 (r) = − n1/3 (r), (4.79)
2 π
where α usually taken in the range 2/3 < α < 1. For example, α =
0.772 and 0.978 [Schwarz (1972)] for He and H atoms, respectively.
Eq. (4.80) is called the Xα potential. > >
E Z 1 n(r′ )
= C1 n2/3 + C2 n1/3 −
dr′ , (4.81)
Ne |r| 2 |r − r′ |
+
where Z/|r| is the external potential that each electron feels the
presence of the ions and the last term is the Coulomb repulsion
between an electron at position r and all other electrons at r′ , which
is expressed by the density n(r′ ). The factor of 1/2 of the last term
196 | Density‐Functional Theory
Ne = n(r)dr.
(4.82)
By substituting Eq. (4.82) into Eq. (4.81), the total energy of the
system is expressed as
(4.83)
n(r) 1 n(r)n(r′ ) ′
−Z
2 |r − r′ |
dr + dr dr,
|r|
where the first and second terms of the right‐hand side are the
kinetic and exchange energies of the free‐electron gas, respectively,
the third term is the Coulomb attraction between the ions and the
electron density, and the last term describes Coulomb repulsion
between the electrons. n(r) for the ground‐state and the total
energy of the system can be obtained by minimizing the energy
functional E[n(r)]. Although the Thomas‐Fermi‐Dirac approximation
is not accurate good‐enough compared with the present electronic
structure calculation, the approach shows a prototype expression of
the DFT, in which n(r) is a crucial physical quantity to calculate the
ground‐state properties of a many‐electron system.
where F includes all the terms in the Hamiltonian except for the
external potential. That means that F contains the kinetic energy and
electron‐electron interaction terms. Therefore, F has the same shape
for all Ne ‐electrons systems with any external potentials.
The total energies of the ground states of the Hamiltonians is
given by
(r)] n(r)dr
⟨ψ |Ven −
′
Ven
′
|ψ′ ⟩ = [Ven (r) − Ven
′
(4.89)
and
(r)] n(r)dr.
⟨ψ|Ven − Ven
′
|ψ⟩ = [Ven (r) − Ven
′
(4.90)
potential Ven (r) and hence the Hamiltonian H. Because the ground‐
state wavefunction ψ is obtained by solving the Schrödinger equation
for H, ψ must be a unique functional of n(r). Therefore, we obtain the
following equation:
1 n(r)n(r′ )
drdr′ + Exc [n(r)], (4.94)
|r − r′ |
F[n(r)] = Ke [n(r)] +
2
i
200 | Density‐Functional Theory
∇2
Ke [n(r)] = ϕi − r ϕi , (4.96)
i
2
1 n(r)n(r′ )
E[n(r)] = Ke [n(r)] +
2 |r − r′ |
drdr′
(4.97)
+ Exc [n(r)] + Ven (r)n(r)dr,
∇2r
+ Veff [n(r)] ϕi (r) = ϵi ϕi (r), (4.99)
2
−
n(r′ )
δExc [n(r)]
, (4.100)
|r − r′ |
Veff [n(r)] = Ven (r) + dr′ +
δn(r)
where the first term Ven (r) is the external potential due to the ions,
the second term is the Hartree potential VH (see Eq. (4.33)) and the
Quantum ESPRESSO Course for Solid‐State Physics | 201
last term is the variational functional derivative of the exchange‐
correlation interaction Exc [n(r)]. The last term in Eq. (4.100) is
defined as the exchange‑correlation potential:
δExc [n(r)]
Vxc [n(r)] = . (4.101)
δn(r)
Multiplying the Kohn‐Sham equation in Eq. (4.99) by ϕ∗i (r) from the
left and summing over all occupied states, we obtain the Kohn‑Sham
energy as
(4.102)
ϵi = Ke [n(r)] + Veff (r)n(r)dr.
i
202 | Density‐Functional Theory
By substituting Veff (r) from Eq. (4.100) into Eq. (4.102), then
subtracting the total energy E[n(r)] in Eq. (4.97), we find:
1 n(r)n(r′ )
E[n(r)] =
|r − r′ |
ϵi −
2
drdr′ − ΔExc [n(r)]
i
(4.103)
1
ϵi − VH (r)n(r)dr − ΔExc [n(r)],
2
=
i
Note that the ratio of the exchange potential in Eq. (4.109) to the
Slater exchange potential in Eq. (4.79) is 2/3.
Since Eq. (4.107) is based on the free‐electron gas (see Sec. 4.8),
the correlation interaction is needed to capture accurately the many‐
body system. Therefore, the energy of correlation interaction ϵc (r) is
added to Eq. (4.107) as
LDA
Exc [n(r)] = [ϵx (r) + ϵc (r)]n(r)dr. (4.110)
4πr3s 1 3
1/3
, or rs = , (4.112)
3 n(r) 4πn(r)
=
ci c0 c1 c2
+ 3/2 + 2 + … (rs ≫ 1), (4.113)
∞
rs rs
ϵc (rs ) =
i=0 rs rs
i/2+1
=
n↑ (r) − n↓ (r)
, (4.114)
n(r)
ζ=
b
0)
X(x0 ) X(x)
3.72744 7.06042
ln (x−x +
VWN
c = 12.9352
=
18.0578
]}
2(b+2x0 )
arctan 2x+b , where
Q
Q x0 = −0.10498 −0.32500
x = rs , Q = 4c − b2 , and
√ √
X(x) = x + bx + c 2
....................................................................................................................................................................
A = 0.0311 0.0311
B = −0.048
A ln(rs ) + B + Crs ln(rs ) + Drs ,
C = 0.0020
−0.048
0.0007
where rs < 1 and
PZ D = −0.0116
γ/(1 + β1 rs + β2 rs ) with
√ −0.0048
rs ≥ 1
γ = −0.1423 −0.0843
β1 = 1.0529 1.3981
0.3334 0.2611
>
β 2 =
....................................................................................................................................................................
A = 0.031091 0.015545
α1 = 0.21370 0.20548
[
−2A(1 + α1 rs ) ln 1 + β1 = 7.5957 14.1189
PW ] β2 = 3.5876 6.1977
1 β3 = 1.6382 3.3662
β4 = 0.49294 0.6251)
p+1
2A β1 rs +β2 rs +β3 rs +β4 rs
1/2 3/2
p = 1.0 1.0
where n↑ (r) and n↓ (r) are the electron densities of up‐spin and
down‐spin states. ζ = 0 and 1 are the spin‐unpolarized and spin‐
polarized systems, respectively.
In Fig. 4.10, we show −ϵc by VWN, PZ, and PW parameters
as a function of rs for the uniform electron system with both
the spin‐unpolarized and spin‐polarized systems by using the
parameters in Table 4.2. All functions fit very well with the
numerical results (the circles and diamonds symbols for the spin‐
unpolarized and spin‐polarized cases, respectively) of Ceperley and
Alder [Ceperley and Alder (1980)]. For rs < 5, which is relevant for
condensed metals, the correlation energy becomes more important.
On the other hand, the correlation energy > becomes a small
206 | Density‐Functional Theory
WVN PZ PW
spin-unpolarized spin-unpolarized spin-unpolarized
spin-polarized spin-polarized spin-polarized
Fx (s) = 1 + κ − , (4.117)
κ
1 + μs2 /κ
where κ = 0.804 and μ = 0.21951 are constants. Note that the range
of s for the real systems is 0 ≤ s ≤ 3. Eqs. (4.115) and (4.117) show
that
1 + At2
H[n(r), ζ] = γϕ3 ln 1 + t2 . (4.120)
β
γ 1 + At2 + A2 t4
9 The repulsive Coulomb energy (exchange plus correlation energy) has a lower
bound of the form C2 n4/3 (r)dr, where n(r) is the single‐particle electron density.
208 | Density‐Functional Theory
LDA-PW
Expt.
GGA-PW91 GGA-PBE
μ (N+1)
Δxc
ϵg Eg
μ (N)
k k
Figure 4.12 Illustration of the contribution of Δxc , the discontinuity in
Vxc [n(r)]. The energies ϵ of the Kohn‐Sham equation are shown in the form
of a band structure for the (N)‐ and (N + 1)‐ electron systems. The two
differ in a uniform increase of the eigenvalues by Δxc . The quasiparticle
band‐gap Eg is the difference between the two eigenvalues as Eg = ϵN+1 −
(N+1)
ϵg = ϵN+1 − ϵN .
(N) (N)
GGA always underestimate the band gap. First, let us explain the
reason why both the LDA and GGA underestimate the band gaps
for semiconductors and insulators even though they are the basis of
good approximations for the ground‐state properties.
The origin of the band‐gap problem is the discontinuity
of chemical potential in the exchange‐correlation potential
Vxc [Sham and Schlüter (1985)]. This is because the density
is given by the summation up to the chemical potential μ as
n(r) = i Θ(μ − ϵi )|ϕi |2 , where Θ is the Heaviside step function.10
<
10 The Heaviside step function Θ(x) is a discontinuous function, whose value is zero
for negative arguments x < 0 and one for positive arguments x > 0.
>
210 | Density‐Functional Theory
Eg = ϵg + Δxc , (4.122)
where ϵg is the gap obtained from the Kohn‐Sham equation for the
ground state. In Fig. 4.12, we show the contribution of Δxc to Eg =
ϵN+1 − ϵN , where ϵJ denotes J‐th energy for (N)‐electron system.
(N+1) (N) (N)
As shown in Secs. 4.11.1 and 4.11.2, both the LDA and GGA do not
consider the discontinuity in Vxc [n(r)] since they are the smooth and
local functions of n(r). Therefore, the LDA and GGA would give zero
Δxc , which is the reason why they always underestimate the value of
Eg even they give a good approximation to ϵg .
Several hybrid functionals have been proposed to obtain non‐
zero value of Δxc . The idea of the hybrid functional is based on a
non‑local functional to the exchange‐correlation functional Exc [n(r)].
As discussed in Sec. 3.6, the exchange potential of the Hartree‐
Fock equation is a non‑local potential (see Eq. (4.47)). The exchange
energy can be obtained from the Hartree‐Fock exchange potential as
1
ϕ∗i (r1 )ϕ∗j (r2 )ϕj (r1 )ϕi (r2 )
ExHF dr1 dr2 , (4.123)
2 i,j |r1 − r2 |
=−
1 1 − erf(ηr) erf(ηr)
, (4.124)
r r r
= +
SR LR
1
EH [n(r)] = VH (r)n(r)dr, (4.126)
2
where a factor of 1/2 takes into account the double counting and
VH (r) is the Hartree potential, which is given in Eq. (4.33).
For a periodic solid, the total electron density and the potentials
can be expressed in the terms of the plane waves, eiGr , with G is the
reciprocal lattice vectors (see Sec. 5.3) as
G G
n(G) = n(r)e
−iGr
dr, and VH (G) = VH (r)e−iGr dr. (4.128)
For the neutral case, the total negative charge of the electrons is
canceled by the total positive charge of the ions. Therefore, the
average potential is zero. Since G = 0 in Eq. (4.127) corresponds
to the average potential over all the space, we can omit G = 0 in
the summation due to charge neutrality.12 Then, by substituting
12 There might be a contribution from the G = 0 term if the systems have an intrinsic
1
ei(G+G )r VH (G)n(G′ )dr
EH [n(r)] =
2 ′
′
GG
Ω
δG+G′ ,0 VH (G)n(G′ )
2
=
Ω (4.129)
VH (G)n(−G)
2
=
G̸=0
Ω
VH (G)n(G),
2
=
G̸=0
where Ω is the volume of the unit cell. Here we assume that n(+G) =
n(−G). On the other hand, by substituting Eq. (4.127) into the Poisson
equation (see Sec. 4.5), we have
n(G)
∇2 VH (r) = −4πn(r) ⇐⇒ VH (G) = 4π . (4.130)
|G|2
n2 (G)
EH = 2πΩ . (4.131)
|G|2
G̸=0
where ϵxc (r) = ϵx (r) + ϵc (r) can be expressed in the terms of the
plane waves as
(4.134)
Exc [n(r)] = Ω ϵxc (G)n(G).
G̸=0
Since the external potential term Ven (r) in Eq. (4.137) is costly
computation with the plane waves, Ven (r) is replaced by a pseudopo‑
tential Vps (r), which is related to replacing the effects of the core
electrons with an effective potential.
A local form of the pseudopotential of a single ion, vps (r),
is proposed by Ashcroft‐Heine‐Abarenkov [Ashcroft (1966),
Heine and Abarenkov (1964)] as
<
v0 r < rcut
vps (r) = , (4.138)
−r Z
r > rcut
where Z is the charge of the ionic core and rcut is the effective radius,
>
v
Core electrons Shell electrons
rcut
r
v0
vps
Z
−
r
1
vIps (|r − T − RI |)e−iGr dr
V
Vps (G) =
I T
Ncell
e vIps (|r|)e−iGr dr
V
−iGRI
=
I
(4.140)
1
e vIps (|r|)e−iGr dr
−iGRI
=
I
Ω
SI (G)FI (G),
=
I
Quantum ESPRESSO Course for Solid‐State Physics | 217
in which SI (G) and FI (G) are defined by
1
SI (G) = e−iGRI , and FI (G) = vIps (|r|)e−iGr dr,
(4.141)
Ω
factor13 and the atomic form factor14 of the I‐th atom, respectively.
For a given pseudopotential such as Eq. (4.138), the external
energy can be expressed as
(4.142)
Eext [n(r)] = Vps (r)n(r)dr = Ω Vps (G)n(G).
G
Finally, we will discuss the last term EEwald in Eq. (4.125). In Quantum
ESPRESSO, EEwald denotes the ion‐ion Coulomb energy, which is given
by
ZI ZJ
N
1n
. (4.144)
2 |RI − RJ |
EEwald =
I̸=J
13 The structure factor contains all the information about the lattice structure. The
structure factor gives the extinction rule for X‐ray diffraction (see Sec. 5.2).
14 The atomic form factor contains the information of each atom, which can be
calculated or obtained by fitting experimental data. The form factor is important for
obtaining the intensity of X‐ray scattering.
218 | Density‐Functional Theory
ZJ 1 − erf(η|R − RJ |)
Nn n N
ZI
|R − RJ | |R − RJ |
=
J J
SR
(4.145)
erf(η|R − RJ |)
Nn
ZJ ,
|R − RJ |
+
J
LR
ZJ ZJ . (4.146)
|R − RJ | V J |G|2
=
J G̸=0
Then the both SR and LR parts can converge quickly with increasing
|G| and |R| for a given value of η. Therefore, the ion‐ion Coulomb
energy can be obtained by a few terms in the summations over |G|
and J.
When we insert Eqs. (4.145) and (4.146) into Eq. (4.144), we
must subtract separately the term for R = RJ in Eq. (4.146), to avoid
the divergence. If we adopt the following equation:
erf(η|R − RJ |)
= 2√ , (4.147)
η
|R − RJ |
lim
R→RJ π
Quantum ESPRESSO Course for Solid‐State Physics | 219
Eq. (4.144) can be rewritten as
1 − erf(η|RI − RJ |)
N
1n
ZI ZJ
2 |RI − RJ |
EEwald =
I̸=J
ZI ZJ (4.148)
V I,J |G|2
+
G̸=0
Nn
Z2J √ .
η
−
J
π
The ionic forces are used to study the dynamics of ions such as
optimizing ionic positions (see Sec. 3.1.4). Using the total energy in
Sec. 4.12, we can calculate the forces that act to the ion I, which is
given by taking the derivative of the total energy with respect to
individual ionic position RI as
There are two contributions to the ionic forces, one from the
ion‐ion interaction energy EEwald in Eq. (4.148), and one from the
ion‐electron interaction energy Eext in Eq. (4.142). For the ion‐ion
interaction energy, the force of ion I is given by Eq. (4.148) as follows:
∂EEwald ZI ZJ
n N
Fion (RI − RJ ), (4.150)
I
∂RI |RI − RJ |3
=− =
J̸=I
∂Eext ∂Vps
Fion‐e ϕ . (4.151)
I
∂RI ∂RI i
=− =− ϕi
i
∂Vps (r)
Fion‐e n(r)dr.
(4.152)
I
∂RI
=−
15 The Hellmann‐Feynman theorem states that the forces are given by the
1 d ℓ(ℓ + 1)
r2 R(r) = ϵR(r), (4.154)
d
2r2 dr 2r2
− + + V(r)
dr
1 d2
u(r) = ϵu(r). (4.156)
2 dr2
− + V(r)
d2 u
(4.157)
dr2
= −k(r)u(r),
c0 = 1 − 12 h kn (r)
5 2 2
c1 = 1 + 12 h kn−1 (r)
1 2 2 , (4.160)
c = 1 +
12 h kn+1 (r)
1 2 2
2
u2 (r)dr = 1.
(4.161)
Then, the electron density n(r) of a single electron for the 1s state
(Y(θ, φ) = 1/ 4π) is given by
√
R2 (r) u2 (r)
n(r) = . (4.162)
4π 4πr2
=
Note that, for the helium atom, the total electron density is 2n(r) since
we have two electrons.
How to run: To run this tutorial, the readers will do the following
command lines:
equations of second order in which the first‐order term does not appear.
Quantum ESPRESSO Course for Solid‐State Physics | 223
Input file: The input files include a subroutine file
(schrodinger.py) and a main file (check-schrodinger.ipynb).
The subroutine contains the Numerov algorithm in Eq. (4.159) and
the radial Schrödinger equation in Eq. (4.157), and the main file
calculates the ground‐states energy and wavefunction for the helium
atom.
SSP‑QE/dft‑he/schrodinger.py
38 else:
39 assert num_nodes > 0, 'expect #nodes >0
since u[0]<0 while u[infty]>0'
40 eps_max = eps
41 raise Exception('Not converged after %d
iterations.' %( maxiter))
SSP‑QE/dft‑he/check‑schrodinger.ipynb
Figure 4.15 Electron density of helium. Solid and dashed lines denote the
electron densities, which are obtained by the Numerov algorithm and
analytical methods, respectively.
that the electron density from the numerical method (the Numerov
algorithm) reproduces the exact electron density of the helium atom
given by Eq. (4.21).
u2 (r)
∇2 UH (r) = − . (4.164)
r
17 The Verlet algorithm is a simple method for integrating second order differential
SSP‑QE/dft‑he/poisson.py
SSP‑QE/dft‑he/check‑poisson.ipynb
Figure 4.16 The Hartree potential of helium. Solid and dashed lines denote
the Hartree potential, which are obtained by the Verlet algorithm and
analytical methods, respectively.
4
ϵx (r) = C2 n1/3 (r) and Vx [n(r)] = C2 n1/3 (r), (4.166)
3
>
Quantum ESPRESSO Course for Solid‐State Physics | 229
where the parameters A, B, C, D, γ, β1 , β2 are given in Table 4.2, too.
Then, the correlation potential is given by
∂[ϵc (r)n(r)]
. (4.168)
∂n(r)
Vc [n(r)] =
rs d
Vc (rs ) = 1− ϵc (rs ). (4.169)
3 drs
(4.170)
>
1 d2
en H xc u(r) = ϵu(r), (4.171)
2 dr2
− + V (r) + V (r) + V [n(r)]
where Vxc [n(r)] = Vx [n(r)] + Vc (rs ). Then, the total energy of the
helium atom in Eq. (4.103) can be rewritten by
E = 2ϵ −
VH (r)u2 (r)dr − ΔExc [n(r)], (4.172)
where ϵxc [n(r)] = ϵx [n(r)] + ϵc (rs ). Note that we have to use the
electron density n(r) and VH (r) for the helium atom.
How to run: To run this tutorial, the readers will do the following
command lines:
SSP‑QE/dft‑he/dft‑lda‑he.ipynb