An Introduction to DFT
An Introduction to DFT
Contents
References 15
Density Functional Theory (DFT) is one of the most widely used computational meth-
ods in quantum mechanics, providing a powerful framework for studying the electronic
structure of matter. By modeling the behavior of electrons within atoms, molecules,
and solids, DFT has become an essential tool in fields such as chemistry, physics, and
materials science. Unlike traditional wave-function-based methods, DFT simplifies the
complex many-body problem by focusing on electron density as the fundamental quan-
tity, rather than dealing directly with the more complicated wave functions. This shift
1
allows DFT to efficiently describe systems with many particles while maintaining a high
level of accuracy. The development of DFT was spearheaded by Walter Kohn, who was
awarded the Nobel Prize in Chemistry in 1998 for his pioneering work. Kohn’s contribu-
tions revolutionized the way scientists solve the Schrödinger equation for multi-electron
systems. Today, DFT strikes a balance between computational cost and precision, mak-
ing it suitable for large-scale simulations in both research and industrial applications.
It is indispensable for investigating a wide range of properties, from chemical reactivity
and molecular dynamics to the electronic, mechanical, and optical behavior of materials,
helping to unlock new insights into the fundamental nature of matter.
2
The mathematical form of the Schrödinger equation is
where, Ĥ is the Hamiltonian operator, Ψ is the wave function of the system, and E is the
energy eigenvalue corresponding to the quantum state described by Ψ. This equation
is crucial in finding the stationary states of quantum systems. The Hamiltonian, Ĥ,
represents the total energy operator of the system and is typically composed of two parts:
2
the kinetic energy operator, T̂ = − 2m
ℏ
∇2 and the potential energy operator, V̂ = V (r).
The kinetic energy operator describes the motion of particles, while the potential energy
operator accounts for the forces acting on them due to their positions within the system.
Together, these components determine the total energy of the quantum system. Hence,
the Schrödinger equation in three dimensions becomes
ℏ2 2
− ∇ + V (r) Ψ(r) = EΨ(r). (2)
2m
3
with the terms representing:
These interaction terms make solving the Schrödinger equation for large systems of atoms
exceedingly complex. The Coulomb interactions V̂nn , V̂ee and V̂ne describe the forces
between particles, which must be accounted for when determining the total energy of
the system. As the number of atoms increases, the number of interactions grows rapidly,
making direct analytical or numerical solutions impractical. With the explicit forms
of the kinetic energies and interaction terms, the many-body Hamiltonian for a system
consisting of nuclei and electrons becomes:
X ℏ2 X ℏ2 1 X ZI ZJ e2
Ĥ = − ∇2RI − ∇2ri +
I
2MI i
2me 2 I,J |RI − RJ |
(4)
1X e2 X ZI e2
+ − .
2 i,j |ri − rj | I,i
|R I − ri |
Here, the indices I and J run over the nuclei, while i and j run over the electrons. RI and
MI are the position and mass of the nuclei, respectively, and ri and me are the position
and mass of the electrons. The terms |RI − RJ |, |RI − ri |, and |ri − rj | represent the
distances between nuclei-nuclei, nuclei-electrons, and electrons-electrons, respectively. ZI
is the atomic number of the I-th nucleus.
4
interactions, further contribute to the difficulty. While analytical solutions are only fea-
sible for very simple systems, such as the hydrogen atom, most many-body systems with
multiple electrons and nuclei do not permit exact solutions.
Given the complexities associated with the many-body Hamiltonian, particularly in sys-
tems containing both electrons and nuclei, it becomes essential to simplify the problem
for practical analysis. One of the most widely used approaches is the Born-Oppenheimer
approximation, which exploits the significant difference in mass and, consequently, the
motion between electrons and nuclei. Since nuclei are much more massive than electrons,
their motion can be considered relatively slow compared to the fast-moving electrons.
This allows us to treat the nuclei as fixed in space while solving for the electronic wave
functions. By separating the total wave function into electronic and nuclear compo-
nents, the Born-Oppenheimer approximation reduces the complexity of the many-body
Schrödinger equation.
Under this approximation, the total Hamiltonian can be expressed as a sum of two parts:
the electronic Hamiltonian Ĥel , which describes the motion of electrons in the field of
fixed nuclei, and the nuclear Hamiltonian Ĥnuc , which accounts for the motion of the
nuclei interacting with each other. The total Hamiltonian Ĥ is then given by:
where
X ℏ2 1X e2 X ZI e2
Ĥel = − ∇2ri + − (6)
i
2me 2 i,j |ri − rj | I,i
|RI − ri |
and
X ℏ2 1 X ZI ZJ e2
Ĥnuc = ∇2RI + . (7)
I
2M I 2 I,J
|R I − RJ |
This simplification allows for easier calculations since the focus is primarily on the elec-
tronic structure. In many cases, especially in density functional theory, the nuclei are
often treated as fixed due to their relatively large mass compared to electrons. This
approximation streamlines the calculations by reducing the complexity of the system to
that of the electrons interacting in a static potential created by the fixed nuclei. How-
ever, if nuclear motion needs to be included, the Schrödinger equation for the nuclei
can be solved, treating them as quantum mechanical entities. The total energy Etotal of
the system is then obtained as the sum of the electronic energy and the nuclear energy.
5
This dual approach captures the intricate interplay between electronic and nuclear dy-
namics, providing a comprehensive view of the system’s energy by accounting for both
contributions.
Once the electronic Hamiltonian Ĥel is defined, the electronic Schrödinger equation can
be expressed as:
Ĥel Ψ(r1 , r2 , . . . , rN ) = EΨ(r1 , r2 , . . . , rN ). (8)
Despite its advantages, the Hartree-Fock approximation has significant limitations. One
of the primary drawbacks is that it assumes a mean-field approximation where each
electron moves in an average field created by all other electrons. This simplification
neglects the electron correlation effects, which are crucial in accurately describing the
behavior of many-electron systems, particularly in cases where strong correlations are
present. As a result, the Hartree-Fock method often leads to inaccuracies in calculated
properties such as bond lengths, reaction energies, and excitation energies.
Furthermore, while the Slater determinant effectively accounts for the antisymmetry of
the wave function, it does not capture the dynamic correlation between electrons, which
arises from their instantaneous interactions. Additionally, the computational cost of
6
the Hartree-Fock method scales as O(N 4 ), where N is the number of electrons in the
system. This scaling arises because the method involves calculating integrals over all
pairs of electron orbitals. As the number of electrons increases, the number of required
calculations grows rapidly, making it computationally expensive for larger systems.
In Density Functional Theory (DFT), the electron density n(r) serves as the central
variable for describing a quantum system. The electron density represents the probability
of finding an electron at a particular point in space and can be derived from the many-
body wave function Ψ(r1 , r2 , . . . , rN ) by integrating over all electron coordinates. This
relationship is expressed mathematically as:
Z
n(r) = N |Ψ(r1 , r2 , . . . , rN )|2 dr2 dr3 . . . drN . (9)
Additionally, we must remember that all electrons are identical; thus, we cannot label
them as electron 1 or electron N . Instead, we can determine the probability of any order
or set of N electrons being located at the coordinates r1 to rN .
While the wave function contains comprehensive information about the quantum state of
a system, it is the electron density that ultimately determines all measurable properties.
The total number of electrons N in the system can also be calculated from the electron
7
density using the equation: Z
N= n(r) dr. (10)
This integration highlights that the electron density encodes vital information about
the total number of electrons, making it a fundamental aspect of DFT. By focusing on
n(r) instead of the complex multi-dimensional wave function, DFT simplifies calculations,
making it a practical and efficient approach for studying the electronic structure of various
materials.
Density functional theory as we know it today was born in 1964 when a landmark paper
by Hohenberg and Kohn appeared in the Physical Review. The theorems they intro-
duced represent the major theoretical pillars on which all modern-day density functional
theories are built. These theorems laid the groundwork for the fundamental relationship
between electron density and the properties of quantum systems, making DFT a powerful
alternative to wave-function-based methods.
The Hohenberg-Kohn theorems are central to the formulation of DFT, and they can be
summarized as follows:
First theorem: The ground-state electron density n(r) uniquely determines the external
potential Vext (r) acting on the electrons. This means that if the electron density of
a system is known, the external potential can be uniquely inferred, allowing for the
derivation of all ground-state properties, including the total energy, from the electron
density.
According to the first theorem, the ground-state density and the external potential corre-
spond in a one-to-one manner. Since the external potential is fixed, the Hamiltonian, and
hence the wave function Ψ, is determined by the ground-state density n0 (r). The proof
of this theorem is straightforward: Consider the ground states of two N -electron sys-
′
tems, characterized by two different external potentials Vext (r) and Vext (r), which differ
by more than an additive constant. The corresponding Hamiltonians, Ĥ and Ĥ ′ , would
both have the same ground-state density n(r), but different ground-state wave functions,
′ ′ ′ ′ ′
Ψ and Ψ , with ĤΨ = E0 Ψ and Ĥ ′ Ψ = E0 Ψ . Since Ψ is not the ground state of Ĥ, it
8
follows that
E0 < ⟨Ψ′ |Ĥ|Ψ′ ⟩
< ⟨Ψ′ |Ĥ ′ |Ψ′ ⟩ + ⟨Ψ′ |Ĥ − Ĥ ′ |Ψ′ ⟩ (11)
Z
′ ′
< E0 + n0 (r)[Vext (r) − Vext (r)]dr
Similarly,
E0′ < ⟨Ψ|Ĥ ′ |Ψ⟩
< ⟨Ψ|Ĥ|Ψ⟩ + ⟨Ψ|Ĥ ′ − Ĥ|Ψ⟩ (12)
Z
′
< E0 + n0 (r)[Vext (r) − Vext (r)]dr.
Hence, no two different external potentials Vext (r) can give rise to the same ground state
density n0 (r) which determines the external potential Vext (r), except for a constant. That
is to say, there is a one-to-one mapping between the ground state density n0 (r) and the
external potential Vext (r), although the exact formula is unknown.
Second theorem: For any trial electron density n(r), the energy functional E[n] will
yield a value that is greater than or equal to the ground-state energy E0 . The equality
holds when the trial density corresponds to the true ground-state density. This variational
principle implies that one can minimize the energy functional E[n] with respect to the
electron density to find the ground state of a system.
There exists a universal functional F [n(r)] of the density, independent of the external
potential Vext (r), such that the minimum value of the energy functional
Z
E[n(r)] ≡ n(r)Vext (r)dr + F [n(r)] (14)
yields the exact ground-state energy of the system. The exact ground-state density n0 (r)
minimizes this functional. Thus, the exact ground-state energy and density are fully
determined by the functional E[n(r)]. The universal functional F [n(r)] can be written
as:
F [n(r)] ≡ T [n(r)] + Eint [n(r)] (15)
where T [n(r)] is the kinetic energy and Eint [n(r)] is the interaction energy of the particles.
According to the variational principle, for any wave function Ψ′ , the energy functional
reaches its global minimum only when Ψ′ is the ground-state wave function Ψ0 , with
the constraint that the total number of particles is conserved. According to the first
9
Hohenberg-Kohn theorem, Ψ′ must correspond to a ground state with particle density
n′ (r) and external potential Vext
′
(r), making E[Ψ′ ] a functional of n′ (r). Applying the
variational principle:
Z Z
′ ′ ′ ′
E[Ψ ] = n (r)Vext (r)dr + F [n (r)] > E[Ψ0 ] = n0 (r)Vext (r)dr + F [n0 (r)] = E[n0 (r)]
(17)
R
Thus, the energy functional E[Ψ] ≡ n(r)Vext (r)dr + F [n(r)] evaluated for the correct
ground-state density n0 (r) is lower than the value of this functional for any other density
n(r). Therefore, by minimizing the total energy functional of the system with respect to
variations in the density n(r), one can find the exact ground-state density and energy.
This functional, however, only determines ground-state properties and does not provide
any insight into excited states.
The Hohenberg-Kohn theorems are fundamental to density functional theory and offer
significant advantages in computational chemistry and materials science. One of the
primary benefits is their ability to relate the ground-state properties of many-electron
systems directly to electron density, simplifying calculations compared to wave-function-
based methods. The first theorem establishes a unique mapping between the ground-state
electron density and the external potential, ensuring that all ground-state properties can
be derived from the electron density alone. Additionally, the second theorem introduces
a variational principle that allows for the efficient optimization of electron density to find
the ground state. By shifting the focus from finding a function of 3N variables (the wave
function) to a function of three variables (the electron density), the Hohenberg-Kohn
theorems significantly simplify the process of resolving the Schrödinger equation. These
theorems make DFT a versatile tool applicable to a wide range of systems, from small
molecules to large biological complexes.
However, the Hohenberg-Kohn theorems also come with notable disadvantages. They
are limited to ground-state properties, providing no direct insight into excited states,
which poses challenges in studying electronic excitations and charge transfer processes.
The effectiveness of DFT is highly dependent on the choice of exchange-correlation func-
tional, which may not accurately capture all correlation effects, especially in systems with
strong electron-electron interactions. Furthermore, while DFT is generally computation-
ally less intensive than wave-function methods, the calculations of exchange-correlation
energies can still be demanding. The reliance on external potentials and the interpreta-
tion challenges associated with electron density further highlight the limitations of the
10
Hohenberg-Kohn theorems, necessitating careful consideration when applying DFT to
complex systems.
The Kohn-Sham equation is a central component of density functional theory that allows
for the practical calculation of electronic structures in many-body systems. It stems from
the foundational Hohenberg-Kohn theorems, which established the uniqueness of the
ground-state electron density and its relationship with the external potential. The Kohn-
Sham framework reformulates the complex many-body problem into a set of simpler,
non-interacting single-particle equations, making it more tractable for computational
purposes.
In the Kohn-Sham approach, the true interacting electron system is mapped onto an
auxiliary system of non-interacting electrons that yield the same electron density as
the original system. This is achieved through the Kohn-Sham equations, which can be
expressed as:
ℏ2 2
− ∇ + Veff (r) ψi (r) = ϵi ψi (r), (18)
2me
where Veff (r) is the effective potential that includes the external potential and the exchange-
correlation potential. The Kohn-Sham orbitals ψi (r) are used to construct the electron
density n(r) as:
X
n(r) = |ψi (r)|2 . (19)
i
This approach significantly reduces the complexity of solving the many-body Schrödinger
equation by allowing for the treatment of a system of independent particles, while still
capturing the essential effects of electron correlation through the exchange-correlation
functional.
Here, Vext represents the external potential acting on the electrons in the system. This
potential typically arises from the interaction between the electrons and fixed nuclei or
any other external fields applied to the system. In many cases, Vext is described by
the Coulomb potential due to the nuclei, reflecting how electrons experience attraction
towards positively charged atomic cores. This term plays a crucial role in determining
11
the overall potential landscape in which the electrons move, significantly influencing the
electronic structure of the system.
The term VHartree [n(r)] is the Hartree potential, which describes the classical electrostatic
interaction between electrons in a many-body system. It accounts for the repulsion
between charged particles, reflecting that the potential energy experienced by an electron
is due to the distribution of other electrons around it. The Hartree potential is calculated
as: Z ′
2 n(r ) ′
VHartree [n(r)] = e ′ dr . (21)
|r − r |
′
This formulation integrates the electron density n(r ) over all space, considering the effect
of all other electrons on a given electron located at r. This approach provides a mean-field
treatment of electron-electron repulsion, avoiding the complexity of considering every pair
of interactions explicitly.
The exchange-correlation potential Vxc [n(r)] represents the quantum mechanical effects
of exchange and correlation among electrons. It is defined as:
δExc [n]
Vxc [n(r)] = . (22)
δn
The exchange term arises from the antisymmetry requirement of the total wave function
for fermions, accounting for the reduction in energy when two electrons are spatially
separated. The correlation term reflects the correlated motion of electrons that cannot
be captured by a mean-field approach, accounting for the ways in which the presence
of one electron affects the probability distribution of another electron’s position and
momentum.
ℏ2 2
ĤKS = − ∇ + Vext + VHartree [n(r)] + Vxc [n(r)]. (23)
2me
The major distinction between the Kohn-Sham formulation and the Hartree formulation
lies in the inclusion of both exchange and correlation effects in the effective potential,
providing a more accurate description of many-body systems.
Solving the Kohn-Sham equation is a crucial step in density functional theory for ob-
taining the ground-state electron density of a many-body system. In a condensed matter
system, the Kohn-Sham equation provides a method to derive the exact density and
energy of the ground state. The process begins with an initial electron density n(r), typ-
ically a superposition of atomic electron densities. The effective Kohn-Sham potential
Veff is then calculated, and the Kohn-Sham equation is solved to obtain single-particle
12
eigenvalues and wave functions. A new electron density is subsequently calculated from
these wave functions.
Figure 1: Flowchart illustrating the iterative process for solving the Kohn-Sham equation in
density functional theory to obtain the ground-state electron density and associated properties
of a many-body system.
13
3.5 The Role and Calculation of Exchange-Correlation
To improve upon the LDA, the Generalized Gradient Approximation (GGA) introduces
a dependence on the spatial gradients of the electron density. By including informa-
tion about how the density changes in space, GGA provides more accurate results for
systems where the density varies significantly, such as in molecules or surfaces. These
improvements make GGA one of the most widely used approximations in modern DFT
calculations. More advanced methods, such as hybrid functionals, combine the exchange
from exact Hartree-Fock theory with the approximate exchange-correlation from LDA
or GGA. Hybrid functionals, like B3LYP, tend to improve accuracy further, particularly
for molecular systems, by including a portion of exact exchange, which LDA and GGA
inherently miss.
In practice, the calculation of Vxc is an iterative process within the self-consistent field
(SCF) method. The DFT algorithm begins with an initial guess for the electron density,
usually based on atomic configurations. Using this initial density, the Kohn-Sham equa-
tions are solved to update the potential and electron density. The exchange-correlation
potential, Vxc , is recalculated at each step based on the updated density. This process
continues until the electron density converges to a self-consistent solution, meaning that
the input and output densities agree within a set tolerance. Recent developments also in-
clude meta-GGA functionals, which incorporate even higher-order density-related terms,
such as the kinetic energy density, to account for more complex interactions. While com-
putationally more expensive, these functionals can offer improved accuracy for systems
with intricate electronic structures. Thus, although Vxc is not known exactly, various ap-
proximations—from LDA and GGA to hybrid and meta-GGA functionals—allow DFT to
achieve a good balance between accuracy and computational efficiency for a wide range
of materials and molecular systems.
14
References
1. Introduction to Quantum Mechanics – David J. Griffiths
2. Condensed Matter Physics – Michael P. Marder
3. Density-Functional Theory of Atoms and Molecules – Robert G. Parr and Weitao
Yang
4. Density Functional Theory and the Family of (L)APW-methods: a step-by-step in-
troduction – S. Cottenier
5. A chemist’s guide to density functional theory – Wolfram Koch and Max C. Holthausen
6. A bird’s-eye view of density-functional theory – Klaus Capelle, Brazilian Journal of
Physics, 36, 4A (2006)
7. Inhomogeneous Electron Gas – P. Hohenberg and W. Kohn, Physical Review, 136,
B864 (1964)
8. Self-Consistent Equations Including Exchange and Correlation Effects – W. Kohn and
L. J. Sham, Physical Review, 140, A1133 (1965)
15