Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Density Functional Theory — an introduction

Nathan Argaman1,2 and Guy Makov2


1
Institute for Theoretical Physics, University of California, Santa Barbara, CA 93106, USA
2
Physics Department, NRCN, P.O. Box 9001, Beer Sheva 84190, Israel

Density Functional Theory (DFT) is one of the most widely used methods for “ab initio” calculations
of the structure of atoms, molecules, crystals, surfaces, and their interactions. Unfortunately, the
customary introduction to DFT is often considered too lengthy to be included in various curricula.
arXiv:physics/9806013v2 [physics.ed-ph] 19 Jul 1999

An alternative introduction to DFT is presented here, drawing on ideas which are well–known from
thermodynamics, especially the idea of switching between different independent variables. The central
theme of DFT, i.e. the notion that it is possible and beneficial to replace the dependence on the external
potential v(r) by a dependence on the density distribution n(r), is presented as a straightforward
generalization of the familiar Legendre transform from the chemical potential µ to the number of
particles N . This approach is used here to introduce the Hohenberg–Kohn energy functional and to
obtain the corresponding theorems, using classical nonuniform fluids as simple examples. The energy
functional for electronic systems is considered next, and the Kohn–Sham equations are derived. The
exchange–correlation part of this functional is discussed, including both the local density approximation
to it, and its formally exact expression in terms of the exchange–correlation hole. A very brief survey
of various applications and extensions is included.

I. INTRODUCTION
number of retrieved records per year
The predominant theoretical picture of solid–state 1000
and/or molecular systems involves the inhomogeneous Hartree−Fock
electron gas: a set of interacting point electrons mov-
ing quantum–mechanically in the potential field of a
set of atomic nuclei, which are considered to be static
(the Born–Oppenheimer approximation). Solution of
such models generally requires the use of approximation 100
schemes, of which the most basic — the independent Density Functional Theory
electron approximation, the Hartree theory and Hartree–
Fock theory — are routinely taught to undergraduates
in Physics and Chemistry courses. However, there is an-
other approach — Density Functional Theory (DFT) —
10
which over the last thirty years or so has become in- 1970 1975 1980 1985 1990 1995
creasingly the method of choice for the solution of such year
problems (see Fig. 1.). This method has the double ad-
vantage of being able to treat many problems to a suf- FIG. 1. One indicator of the increasing use of DFT is the
number of records retrieved from the INSPEC databases by
ficiently high accuracy, as well as being computationally
searching for the keywords “density”, “functional” and “the-
simple (simpler even than the Hartree scheme). Despite
ory”. This is compared here with a similar search for keywords
these advantages it is absent from most undergraduate “Hartree” and “Fock”, which parallels the overall growth of
and many graduate curricula with which we are familiar. the INSPEC databases (for any given year, approximately
We believe that this omission stems in part from the 0.3% of the records have the Hartree–Fock keywords).
tendency of the existing books and review papers on
DFT, e.g. Refs. [1–3], to follow the historical path of
development of the theory. Although appropriate for of the density distribution as a free variable arises in a
a thorough treatment, this approach unnecessarily pro- natural manner, as do more advanced concepts which are
longs the introduction and grapples with problems which central to recent developments in the theory [5], e.g. the
are not directly relevant to the practitioner. It is our pur- exchange–correlation hole and generalized compressibili-
pose here to give a brief and self–contained introduction ties. The discussion is sufficiently detailed to provide a
to density functional theory, assuming only a first course useful overview for the beginning practitioner, and the
in quantum mechanics and in thermostatistics. We break relatively novel point of view may also prove illuminat-
with the traditional approach by relying on the analogy ing for those experienced researchers who are not familiar
with thermodynamics [4]. In this formulation, the use with it. We hope that the availability of such an introduc-

1
tion will encourage teachers to include a one or two hour
class on DFT in courses on quantum mechanics, atomic Ω (a) (b)
F
and molecular physics, condensed matter physics, and
materials science. ΔF
The general theoretical framework of DFT, involving µ N
the Hohenberg–Kohn free energy FHK [n(r)], is presented Ωµ
in Sec. II, which for simplicity focuses on classical sys-
tems. The generalization to the quantum–mechanical
electron gas is given in Sec. III, together with the dis-
FIG. 2. (a) The Legendre transform which gives F (N )
cussion of the Kohn–Sham equations and of the local
corresponds to describing the curve Ω(µ) by the properties
density approximation, which is the simplest practical
of its tangents: minus their slopes N = −(∂Ω/∂µ) and their
approximation for the exchange–correlation energy. Var- intercepts with the energy axis, F = Ω + µN . The fact that
ious issues relating to the accuracy of this approach are the derivative ∆F/∆N is equal to µ follows from asking the
discussed in Sec. IV, followed by a summary in Sec. V. question: if two neighboring lines intercept the energy axis at
a distance ∆F from each other, and have slopes which differ
by ∆N , how far from the axis will they cross each other?
II. GENERAL THEORY (b) The Legendre transform back from F (N ) to Ω(µ) has a
similar interpretation. The minimization suggested in Eq. (4)
In this section, a unified treatment of thermodynamics corresponds to studying a family of lines with a fixed slope
and density functional theory is presented. For simplic- µ, which pass through points (N, F ) on the free energy curve.
ity, the case of a classical interacting system of point par- Their intercepts, Ωµ = F − µN , have a minimum (marked by
an asterisk) for that line which is tangent to the curve.
ticles will be discussed first. Although classical DFT has
its own applications, e.g. liquids [6], the reader is advised
to keep in mind electronic systems, which will be the sub- the temperature is in energy units (i.e. kB = 1), and
ject of the next section. Thus, Eqs. (1) through (8), to be the classical trace, Tr, represents the 6M –dimensional
derived here using classical notation, are equally applica- phase–space integral (the division by M ! compensates for
ble to quantum–mechanical systems, where Hilbert space double counting of many–body states of indistinguishable
with its position and momentum operators replaces the particles).
classical phase space and its scalar coordinates. It follows directly from these definitions that the ex-
pectation value of the number of particles in the sys-
tem is given by a derivative of the grand potential,
A. Thermodynamics: a reminder
N = "M # = −(∂Ω/∂µ). The convexity of the ther-
modynamic potential [8] implies that N is a monoton-
We begin by rederiving the equations of thermodynam- ically increasing function of µ. Other partial derivatives
ics from statistical mechanics [7]. Consider a classical of Ω give the values of additional physical quantities,
system of M interacting particles in a container of vol- such as the entropy, S = −(∂Ω/∂T ) and the pressure
ume V . The many–body Hamiltonian is: P = −(∂Ω/∂V ). This may be summarized by writing
dΩ = −N dµ − S dT − P dV .
HMB = T + U , (1)
A basic lesson of thermodynamics is that in different
!M
where T = i=1 p2i /2m is the kinetic energy, and U = contexts it is advantageous to use different ensembles.
! For example, in studying systems where the number of
i<j u(|ri − rj |) is the interaction energy, assuming a
simple pair potential u(r). Here ri and pi are the posi- particles rather than the chemical potential is fixed, it is
tions and momenta of the particles, and m is their mass. preferable to use the Helmholtz free energy [9], which
We consider the grand–canonical ensemble, where the is obtained from the grand potential
% Ω by
& a Legen-
system is in contact with a heat reservoir of tempera- dre transform: F (N, T, V ) = Ω µ(N ), T, V + µ(N )N .
ture T and a particle reservoir with chemical potential µ. Here µ(N ) is no longer an independent variable, but
It is well–known from statistical physics that the grand a function of N obtained by inverting the relationship
potential, which is the free energy in this case, is given N = N (µ, V, T ) = −(∂Ω/∂µ). The derivative of F
by: with respect to the “new” free variable N is equal to
the “old” free variable µ. The derivatives with respect
Ω(µ, T, V ) = −T log Ξ , (2) to the other variables are unchanged (but are taken at
constant N rather than at constant µ). We thus write
where Ξ is the grand partition function, dF = µ dN − S dT − P dV .
∞ # $ For the purpose of comparison with DFT, it is useful to
" 1 HMB −µM
Ξ(µ, T, V ) = Tr exp − , (3) make a variation on the inverse Legendre transform which
M! T
M=0

2
expresses Ω in terms of F , and to define the following energy [8,12]). The partial and functional derivatives of
“grand potential function”, which depends explicitly on FHK [n(r)] are given by the( usual rules for Legendre trans-
both µ and N : forms: dFHK = −SdT − dr v(r)δn(r).
The direct generalization of the free energy function of
Ωµ (N, T, V ) ≡ F (N, T, V ) − µN . (4)
Eq. (4) is the free energy functional:
This function gives the original grand potential of Eq. (2) '
when minimized with respect to N , i.e. when the deriva- Ωv [n(r)] ≡ FHK [n(r)] + dr n(r) v(r) , (7)
tive (∂F/∂N ) − µ vanishes, which is equivalent to the
condition N = N (µ, T, V ) conventionally used in the in-
with v(r) and n(r) treated as independent functional
verse Legendre transform. For other values of N , the variables. If this free energy functional is minimized with
function Ωµ (N, T, V ) describes a “cost” in free energy of respect to n(r) at constant v(r) (and given T , etc.), the
having a configuration with the “wrong” number of elec-
relation
trons. For a geometric interpretation of Legendre trans-
forms, including the minimization procedure of Eq. (4), δFHK
= −v(r) (8)
see Fig. 2. δn(r)

is obtained. For n(r) and v(r) obeying this physical re-


B. Nonuniform systems and the Hohenberg–Kohn lation, the free energy functional is equal to the grand
theorem potential by inspection. The existence of a functional of
n(r) with this property is one of the basic tenets of DFT,
The discussion above can be generalized in a quite and is the (second) Hohenberg–Kohn theorem [13].
straightforward manner to the treatment of particles in Note that below we will use Eq. (8), which also fol-
an external potential v(r). The many–body Hamiltonian lows directly from the properties of Legendre transforms.
is now Discussion of the Hohenberg–Kohn theorem, Eq. (7), is
nevertheless important even in a Legendre–transform–
HMB = T + V + U , (5)
based introduction to DFT, because the free–energy–
where the potential energy, V =
!M minimization procedure embodied in it is central both
i=1 v(ri ), has been
added. The grand potential and the partition function to forming a physical intuitive picture of DFT, and to
are defined as before, Eqs. (2) and (3), but they now devising efficient numerical schemes for solving the DFT
depend on the potential function v(r) rather equations in practice [14].
% than on &the
scalar volume V . In this sense, Ω = Ω µ, T, [v(r)] is
now a functional [10] of v(r) as well as a function of µ
C. Examples
and T — the square brackets denote functional variables
(it is also implicitly a functional of the pair potential
u(r)). As is well known, the potential v(r), is an energy We next apply the expressions above to two (simplis-
which is measured from an arbitrary origin, i.e. shifting tic) physics problems: finding the density distribution of
the potential by a constant does not affect the physics air and of water in the ecosphere.
of the system. It is convenient here to set this origin at 1) Air: First we consider air, using the approximate
the chemical potential, i.e. to take µ = 0. Equivalently, model of an ideal gas. For such a gas the particles do
one may define the new functional variable as v(r) − µ, not interact, u(r) = 0, and the partition function can be
as Ω depends only on this difference, and not on v and µ evaluated directly. The Hamiltonian, Eq. (5), reduces to,
separately [11]. HMB = T +V. The grand partition
!∞ function, Ξ of Eq. (3),
M
The functional derivative of Ω with respect to the new may be expressed as Ξ = M=0 1 Ξ /M ! = exp(Ξ1 ),
variable gives the density distribution of! the particles, in terms (of the partition sum for a single particle, Ξ1 =
(2πh̄)−3 drdp exp −(p2 /2m +
% &
M
n(r) = "ρ(r)# = δΩ/δv(r), where ρ(r) = i=1 δ(r − ri ) v)/T . The momentum
integral is trivial, giving Ξ1 = (dr/λ3T ) exp −v(r)/T ,
( % &
is the unaveraged density. Using a (functional) Legendre
transform as above, we can define a new free energy which where) volume is normalized by the “thermal wavelength”
depends on n(r) rather than on v(r), and is called the λT = h̄2 /mT . One finds from Eq. (2) that Ω = −T Ξ1 .
Hohenberg–Kohn free energy: In this case the density distribution% is easily &found di-
rectly: n(r) = δΩ/δv(r) = λ−3 T exp −v(r)/T , but for
'
FHK [n(r)] = Ω[v(r)] − dr n(r) v(r) , (6) pedagogical purposes we proceed to evaluate the Legen-
dre transforms explicitly.
where the explicit temperature variable has been omit-
Inverting the ideal gas n[v] relationship just derived
ted, and v(r) on the right hand side is chosen to cor-
gives v(r) = −T log(nλ3T ), and proceeding with the func-
respond to the given n(r) (that such a choice is possi-
tional Legendre transform of Eq. (6) gives
ble follows from the “generalized convexity” of the free

3
'
FHK [n(r)] =
% &
dr n(r) f n(r), T , (9) term. The Helmholtz free–energy per particle in the uni-
form fluid is thus taken to be
where f (n) = T 3 log(λT ) − log(n−1 − b) − 1 − an , (12)
% &

f (n, T ) = T log(nλ3T ) − 1 ,
% &
(10) which replaces Eq. (10). Correspondingly, the pressure
is
is the free energy per particle. The latter is equal to # $
the Helmholtz free energy per particle, F (N, T, V )/N , ∂F ∂f T
P =− = − −1 = −1 − an2 , (13)
evaluated for a uniform ideal gas with n = N/V . ∂V N ∂n n −b
The DFT free energy functional, Eq. (7), for an ideal
gas is which is a more familiar expression.
' ' The free–energy functional to be minimized with re-
% & spect to n(r) is again that of Eq. (11), but the extremum
Ωv [n(r)] = dr n(r)f n(r), T + dr n(r)v(r) . (11)
condition,

λ3T
# $
Minimizing this energy functional with respect to n(r) ∂(nf ) b
v=− = −T + log + 2an
gives v = −∂(f n)/∂n = −T log(nλ3T ), which is equiva- ∂n n−1 − b n−1 − b
lent to n ∝ exp(−v/T ), as discussed above.
(14)
The density distribution of the atmosphere in the grav-
itational field of the Earth may now be determined. The may now have several solutions, and the value of n which
external potential, measured from the chemical potential, gives the lowest free energy must be selected. In the
is v(r) = mgh − µ, where h is the height and g & 10m/s2 present model, for temperatures which are not too high
is the acceleration due to Earth’s gravitational field (as and for a limited range of values of v, there are two com-
we take h = 0 at Earth’s surface, we have subtracted µ peting local minima, corresponding to the liquid and the
explicitly here). Substituting this in the expression for gas phases of the fluid [16].
the density, we find that the density decreases exponen- We now return to water, representing a typical location
tially with height, n(r) ∝ exp(−h/l), with a length–scale on Earth as a unit of area (one square meter), covered
l = T /mg. Taking m as the mass of a nitrogen molecule, by 108 moles of water (one mole & 6 × 1023 molecules).
m & 5 × 10−26 kg, and the temperature as T = 4 × 10−21 We use the same potential v = mgh − µ and the same
Joules (& 17◦ C), gives l = 8 × 103 m or 8 kilometers, less values of T and g as above. The mass of a water molecule
than the height of the Everest (in practice, the atmo- is m & 18 × 10−3 kg/mole, and we choose the values of
sphere is not in equilibrium, T (= const., and this descrip- a & 0.48Pa×m6/mole2 and b & 16×10−6m3 /mole to give
tion becomes increasingly inaccurate at higher altitudes). the boiling point of water as 100 degrees Celsius at a pres-
2) Water: Finding the density distribution of water re- sure of one atmosphere & 105 Pa, and the particle den-
quires a different analysis, because the ideal gas is a good sity of the liquid water at that point as n & 56moles/m3,
description of real fluids only for low densities such as corresponding to a mass density of 103 kg/m3 . Finding
those found in the atmosphere. At higher densities, the the density distribution, n(h), requires the inversion of
interactions, u(r), must be taken into account. An ex- Eq. (14) which is not readily available analytically; there-
act, explicit evaluation of the partition function and its fore we have performed this exercise numerically. The
derivatives is no longer possible, and we are compelled to chemical potential µ plays the( role of a Lagrange multi-
use models and approximations. The inhomogeneity of plier, imposing the constraint dr n(r) = N , and finding
the distribution will be given an approximate treatment it requires a few trials or iterations (using the Newton–
here, by assuming that the Hohenberg–Kohn energy may Raphson algorithm).
still be expressed as in Eq. (9), in terms of the Helmholtz The resulting value of µ is −2.5385 × 105 Joules/mole
free energy per particle for a uniform system. This is — several significant digits must be kept because v (or
a “local density approximation”, a concept which is in v − µ) varies by only 180 Joules/mole for each kilome-
common use in DFT. For simplicity, we will use the van ter of altitude. The density distribution n(r) contains
der Waals model for f (n, T ). In this model [15], both both liquid and gaseous regions. The liquid region or
attractive and repulsive interactions of real atoms and “ocean” occupies the lower altitudes, and a gaseous re-
molecules (typically included in u(r) for large and small gion occupies higher altitudes. In the lower region, the
r, respectively) are taken into account. Correspondingly, fluid is more or less incompressible, with the density at
two simple modifications in the ideal–gas expressions for h = 0 higher by only a couple of percent than that at
f , Eq. (10) are made: first, a “higher order in n” attrac- “sea level”, h & 1750m. Above this point one finds wa-
tive term is added, and second, an “excluded volume” ter vapor with a density of & 4moles/m3 , corresponding
b, representing the “hard core” of real particles, is sub- to a pressure of 104 Pa or one–tenth of an atmosphere
tracted from the volume per particle n−1 in the thermal (real water deviates significantly from our van der Waals

4
model — a more accurate description of f (n) would give with the Hamiltonian of Eq. (5) replaced by the corre-
∼ 2.5 × 103Pa). At this low density, the equation of state sponding operator, ĤMB = T̂ + V̂ + Û. The statistical–
of the water vapor does not differ significantly from that mechanics derivation above is affected by this, e.g. the
of an ideal gas, and indeed one finds an exponential decay trace (Tr) in the definition of the partition function,
with further increase in altitude, as discussed above for Eq. (3), is taken over the Hilbert space rather than
air (the lengthscale in this case is & 14km, as the water over the classical phase space. However, the thermody-
molecules are lighter). namic considerations and Legendre transforms, includ-
Obviously, water and air coexist on the surface of ing Eqs. (6), (7) and (8) for nonuniform systems, are
Earth, and taking this into account would modify the unchanged.
results of our examples (most importantly, the atmo- As electrons are Fermions, they form a degenerate gas
sphere begins only at sea level). Such situations can be at the prevailing high densities, and their energy in the
addressed by generalizing the free–energy density, nf , cases of interest here does not deviate considerably from
allowing it to depend on both densities nwater and nair . its ground–state value. Indeed, electronic DFT was de-
With a sufficiently accurate parameterization of f , a good veloped in Refs. [13] and [17] as a ground–state theory.
description of many nonuniform thermodynamic systems Correspondingly, we will from here on take the zero tem-
can thus be obtained. However, for such a local descrip- perature limit, T → 0, in all our equations and expres-
tion to hold, the spatial variations in n(r) (or equiva- sions (cf. Ref. [18]). In this limit the partition function
lently, in v(r)) must be very slow on the scale of the is dominated by a single quantum–mechanical state with
range of the potential u(r). In contrast, one may be inter- an integer number of electrons, M = N , and the grand
ested in studying how the density distribution n(r) near potential is equal to the ground–state energy (using our
a liquid–vapor interface changes gradually, on a scale of µ = 0 convention). The Hohenberg–Kohn free–energy,
Angstroms, from that of the liquid to that of the gas. FHK [n], is then, by Eq. (6), equal to the total energy
Such problems are typical applications of the DFT of minus the potential energy, i.e. to the internal energy.
classical systems, and must employ nonlocal model func- For noninteracting quantum mechanical particles,
tionals u(r) = 0, the internal energy is just the kinetic energy,
* for
A discussion of such nonlocal classical functionals is which we introduce the notation Fni [n] = FHK [n] *u(r)=0 .
beyond the scope of the present paper, but we would like Whereas in the classical case, the noninteracting or ideal–
nevertheless to mention the following two points: (a) As gas system had simple local expressions for the energy
opposed to gases, classical liquids and solids are charac- functionals, here the relationships between the energy,
terized by strong and complicated correlations between the density distribution, and the confinement potential
particles, even when the interaction potential u(r) is sim- are nontrivial and nonlocal.
ple. This makes the task of finding good FHK [n] func- Practical implementations of DFT require an explicit
tionals nontrivial. (b) The general framework of DFT construction of the Hohenberg–Kohn free–energy func-
can nevertheless give accurate descriptions of real sys- tional, FHK [n]. It is customary to write FHK [n] for in-
tems, provided that one uses good “anchoring points”. teracting electrons as a sum of the noninteracting kinetic
For example, the hard–sphere liquid has been studied energy, Fni [n], and two interaction terms — the electro-
extensively numerically, and accurate information for it static energy and the exchange–correlation energy [19]:
is available. A good description of the liquid–solid tran-
sition in several more general systems may be found by FHK [n(r)] = Fni [n(r)] + Ees [n(r)] + Exc [n(r)] , (15)
perturbing in the difference between the actual u(r) and
that of the hard where the last term, Exc [n], is defined as the remainder
( sphere system,
% uhs (r), &e.g. by taking
and thus contains everything that is not included in the
FHK [n] = Fhs + dr (u − uhs) δFHK /δu(r) , where in the
last term, the functional derivative is equal to δΩ/δu(r) first two terms. Each of the three terms on the right
which is the pair correlation function. With this as back- hand side is in principle a functional of the independent
ground, we now proceed to treat electronic systems in variable n(r). Only the second term — the electrostatic
the next section. energy — is easily expressed explicitly:
e2 dr dr$
'
Ees [n(r)] = n(r) n(r$ ) , (16)
III. APPLICATION TO ELECTRONS
2 |r−r$ |
The first and last terms are much more complicated:
The electron is much lighter than an atom or molecule, knowledge of the former implies a full understanding
and thus√has a relatively large thermal wavelength, e.g. of the quantum–mechanical noninteracting problem; the
λT & h̄/ mT & 17Å at room temperature, more than an latter contains all of the many–body physics, and is in
order of magnitude larger than the typical inter–electron principle even more complex.
distance. Its treatment must be quantum–mechanical, In the following, we briefly introduce two alternative
methods for confronting this situation: (i) construction

5
of explicit approximate expressions for both Fni [n] and B. The Kohn–Sham equations
Exc [n], and (ii) the orbital method developed by Kohn
and Sham, which uses the noninteracting Schrodinger In 1965, Kohn and Sham [17] made a major step to-
equation to evaluate Fni [n], with only the smaller term, wards quantitative modeling of electronic structure, by
Exc [n], replaced by explicit approximations to the desired introducing an orbital method by which Fni [n] can be
complicated functional. evaluated exactly. In other words, in order to evaluate
the kinetic energy of N noninteracting particles given
only their density distribution n(r), they simply found
A. Explicit functionals
the corresponding potential, called veff (r), and used the
Schrodinger equation,
One of the simplest models of the electronic structure
of atoms was developed by Thomas [20] and Fermi [21] h̄2 2
# $
already in the late 1920’s. For the noninteracting part of − ∇ + veff (r) ψi (r) = 'i ψi (r) , (18)
2m
the calculation, they used a local density approximation:
!N
such that n(r) = i=1 |ψi (r)|2 . The states ψi here are
'
Fni [n(r)] & C n5/3 (r) dr , (17) ordered so that the energies 'i are non–decreasing, and
the spin index is included in i. If 'N is degenerate with
3
with C = 10 (3π 2 )2/3 & 2.87 in atomic units; here 'N +1 (and also at finite temperatures! [25]), fractional oc-
5/3
Cn (r) is the kinetic energy density of a uniform cupations fi are to be used, n(r) = ∞ 2
i=1 fi |ψi (r)| , but
electron–gas of density n(r). They also approximated if only spin–degeneracy is involved, the result for the den-
the Coulomb interaction energy of the electrons by the sity is not affected. The kinetic energy is( then given by
Fni [n(r)] = N
!N
electrostatic term only, which corresponds in the present
!
i=1 "ψi |t̂ i |ψi # = i=1 'i − dr n(r)veff (r),
language to taking Exc [n] = 0. Using these simplifica- where! t̂i is the kinetic energy operator for the ith electron
tions and the spherical symmetry of an atom, analytical (T̂ = i t̂i ).
progress could be made. In practice, it is the external potential of a given sys-
Both approximations made in the Thomas–Fermi tem which is known, not the density distribution or the
model are expected to be accurate at very high densities. effective potential. One may find the effective potential
The approximation made for the noninteracting part is by taking a functional derivative of the three–term ex-
accurate when the density of electrons changes slowly in pression for FHK [n], Eq. (15), and rearranging the terms:
space relative to the Fermi wavelength, or equivalently,
if the density is sufficiently high that n(r)1/3 is (much) veff (r) = v(r) − eϕ(r) + vxc (r) , (19)
larger than the spatial rate of change of the density.
The approximate treatment of the interactions, which where we have used Eq. (8), δF [n]/δn = −v, for both
ignores exchange and correlation effects, is the leading the interacting and the noninteracting system. The elec-
order of a high–density expansion, in powers of rs /a0 trostatic potential is here
where rs = (3/4πn)−1/3 is the Wigner–Seitz radius, and
n(r$ )
'
a0 = h̄2 /me2 = 0.53Å is the Bohr radius. The Thomas– ϕ(r) = −e dr$ , (20)
Fermi model indeed has applications for very dense mat- |r−r$ |
ter, and is useful in the description of certain stars [22]. and the exchange–correlation potential is defined as
Unfortunately, the results for more down–to–earth sys-
tems are rather poor. For example, it is known that in δExc
vxc (r) = . (21)
the Thomas–Fermi model no molecules can form — the δn(r)
dissociated atoms always have a lower energy [23].
The Thomas–Fermi method was extended over the Given a practical approximation for Exc [n], one obtains
years in two main directions. At first, approximate ex- vxc (r), and can thus find veff (r) from n(r) for a given
pressions for the exchange–correlation energy were in- v(r).
cluded, but this did not lead to significant improvements The set of equations described above is called the
in the results. Such improvements were obtained only Kohn–Sham equations of DFT, and must be solved self–
when an element of nonlocality was taken into account consistently: veff (r) determines n(r) in Eq. (18), and is
in the noninteracting kinetic energy term, Fni [n]. This determined by it in Eq. (19). They provide a mecha-
was achieved by including gradient terms (e.g. terms pro- nism for minimizing the functional Ev [n] (or Ωv [n]) of
portional to |∇n|2 in the integrand of Eq. (17), with Eq. (7), despite the fact that Fni [n] is known only im-
n–dependent prefactors), and lead in particular to the plicitly. This method assumes that the external poten-
possibility of modeling chemical bonds. However, the tial v(r) and the total number of electrons N are given,
description of electronic structure by the Thomas–Fermi and does not use the chemical potential µ (if a nonzero
model and its extensions remains qualitative to date [24]. value of µ were restored, it would simply shift both sides

6
of Eq. (19) by a constant, and thus drop out of consid-
eration). Together with any explicit approximation for
the exchange–correlation term (see below), it provides an
efficient scheme for finding n(r) and the ground state en- 0 ε
ergy [26] for a system of N interacting particles. Note ε tot
that for the interacting system, only the energy and its −0.1
derivatives (including n(r)) are accessible — the compli-
cated many–body wavefunctions do not take part in this −0.2
scheme. ε
Historically, additional properties [27] of the Kohn– xc
−0.3
Sham noninteracting system, e.g. the band structure for
crystals, have also provided surprisingly accurate predic- 0 1 2 3 4 5 6 7 8
tions when compared with experiments. In fact, the rs
agreement between the calculated Kohn–Sham Fermi
surface and the measured one was so remarkable for some FIG. 3. The exchange–correlation energy per electron,
systems [28], that it motivated analyses of soluble (per- "xc (n) (in Hartrees) of a uniform electron gas, as a func-
turbative) models for which the difference between the in- tion of the density, parameterized by the Wigner–Seitz ra-
teracting and noninteracting Fermi surfaces could be cal- dius (in Bohr radii), according to the interpolation formula
culated explicitly, and shown not to vanish [29]. Clearly, of Ref. [31]. The dashed line is the corresponding total
the accuracy of DFT predictions for ground–state ener- energy per electron, i.e. includes the kinetic energy. The
gies and density distributions can be improved by find- Monte–Carlo data, if plotted, would be indistinguishable on
ing better practical approximations for Exc [n], whereas the scale used here from the curves shown. For reference,
improving the accuracy of such band–structure calcula- we also show the total energy per electron (dotted line) ac-
cording to an alternative parameterisation, the well–known
tions may require “going back to the drawing board” and
Wigner interpolation formula, Ref. [32].
devising other, more appropriate, calculational schemes
[30].
to the electrostatic potential at the appropriate step in
the self–consistency loop [33]. The corresponding expres-
C. The Local Density Approximation sion for the ground–state energy is:

As a practical approximate expression for Exc [n], Kohn N


" '
% &
and Sham [17] suggested what is known in the context of E0 = 'i − Ees [n(r)] + dr n(r) 'xc (n(r)) − vxc (n(r)) ,
DFT as the local density approximation, or LDA: i=1

' (24)
Exc [n(r)] & dr n(r) 'xc (n(r)) , (22)
where the first term is the noninteracting energy, the sec-
ond term subtracts half of the double counting of the elec-
where 'xc (n) is the exchange–correlation energy per elec- trostatic energy as in the Hartree scheme, and the last
tron in a uniform electron gas of density n. This quan- term is a similar subtraction for the exchange–correlation
tity is known exactly in the limit of high density, and energy.
can be computed accurately at densities of interest, us- The LDA has been shown to give very good results for
ing Monte Carlo techniques (i.e. there are no free pa- many atomic, molecular and crystalline interacting elec-
rameters). In practice one usually employs paramet- tron systems, even though in these systems the density
ric formulas, which are fitted to the data and are ac- of electrons is not slowly varying. As an example, we
curate to within 1–2%. As an example, we quote that show in Fig. 4 the solution for an atom of Argon. One
given by Gunnarson and Lundqvist (Ref. [31]): 'xc (n) = can see that the shell structure, which is absent in the
−0.458/rs − 0.0666G(rs /11.4) Hartrees, where rs is in Thomas–Fermi model, is described here in detail. The
units of the Bohr radius, and G(x) = 21 {(1+x3 ) log(1+ calculated ground–state energy is −525.9 Hartrees, com-
x−1 ) − x2 + 21 x − 31 }, see Fig. 3. pared to −652.7 Hartrees in the Thomas–Fermi model
Note that the only difference between the resulting (one Hartree = me4 /h̄2 = 27.2eV). The experimental
computational scheme and a naive mean–field approach value is −527.6 Hartrees — the LDA result is accurate
is the addition of the potential to within less than half a percent, compared to the 20%
% &* inaccuracy of the Thomas–Fermi model. Both the de-
d n 'xc (n) **
vxc (r) = (23) scription of the shell structure and much of the improve-
dn
*
*
n=n(r) ment in the energy estimate are due to the introduction
by Kohn and Sham of an exact method for evaluating

7
1
∂FHK [n]
'
Ar Z=18 FHK [n] = Fni [n] + dΛ . (25)
2 0 ∂Λ
1s
DFT The general rules for Legendre transforms give the deriva-
2s+p
T−F tive ∂FHK+[n]/∂Λ
, as equal to ∂Ω/∂Λ, which in turn is
equal to Û , the expectation value of the interaction
2
n×r

1
energy Û . Note that the derivative of Ω is taken at a
constant potential, but in order to reproduce the given
3s+p
density distribution n(r), this potential must depend on
Λ (it is usually denoted by vΛ (r), but we need it here
0 only at the two extremes, for which we already have ap-
0 1 2 3 4 propriate notations: v1 (r) = v(r) and v0 (r) = veff (r)).
r Comparing Eq. +(25)
(1 , with Eq. (15), we find that
FIG. 4. The density n(r) for an Argon atom, multiplied Exc [n] = 0 dΛ Û − Ees [n], i.e. the exchange–
by r 2 for convenience of presentation (inverse Bohr radii), as
a function of the distance from the origin (in Bohr radii). The
correlation energy is the difference between the Λ–
full line is the DFT result using the LDA, and shows the shell averaged expectation value of the interaction energy and
structure; the dashed line is the result of the Thomas–Fermi the electrostatic approximation to it. Both of these are
model for this atom. integrals over the product of the -Coulomb% interaction and
the density of pairs, which is 21 ρ̂(r) ρ̂(r$ ) − δ(r − r$ )
&.

and 21 n(r)n(r$ ) for the exact and approximate expres-


the highly nonlocal kinetic energy functional. However, sions, respectively (the factor of 12 corrects for the dou-
the accuracy of the exchange–correlation term is also of ble counting of each pair, and the δ function removes
central importance, and will be discussed next. the interaction of each electron with itself). We are thus
motivated to define a pair correlation function, g(r, r$ ),
equal to the difference between the two pair densities
IV. ACCURACY OF THE
EXCHANGE–CORRELATION ENERGY
g(r, r$ ) = ρ̂(r) ρ̂(r$ ) − δ(r − r$ ) n(r),Λ − n(r) n(r$ ) .
- % &.

In this section, we discuss a few aspects of Exc [n] which (26)


are somewhat more advanced (and may be skipped in a
first reading). In the first subsection, a formally exact It is of relevance here that g(r, r$ ) at typical electron den-
expression for the exchange–correlation energy is derived, sities is relatively featureless — it is only at very low den-
and a sum–rule which applies to it is obtained. The next sities (rs ∼ 100a0 ) that electrons tend to develop strong
subsection provides references to some more recent and correlations, and may form a Wigner crystal.
accurate approximations, which were devised with this ( Using the fact that the number of particles, M =
sum–rule in mind. Finally, a general discussion of the dr$ ρ̂(r$ ), does not fluctuate at very small( temperatures,
level of accuracy achieved by these approximations for and is equal to its expectation
( $ value, N = dr$ n(r$ ), it is
$
different applications is given. easy to show that dr g(r, r ) = −n(r). It is convenient
to define a normalized version of this correlation func-
tion ρxc (r, r$ ; [n], Λ) = g(r, r$ )/n(r), which is the density
A. The exchange–correlation hole of the so–called exchange–correlation hole. ρxc describes
the region in r$ –space from which an electron is “miss-
A deeper understanding of the exchange–correlation ing” if it is known to be at the point r, and the fact that
energy can be achieved by considering a continuous tran- its integral over r$ is equal to −1 corresponds to the fact
sition between the interacting and noninteracting sys- that there is exactly one “missing” electron. In terms of
tems which appear in the definition of Exc [n], rather the Λ–integrated value of this quantity,
than using a simple subtraction as in Eq. (15). To do ' 1 ' 1
this, we reduce the strength of the Coulomb interaction, w(r, r$ ) = dΛ ρxc (r, r$ ; [n], Λ) = dΛ g(r, r$ )/n(r) ,
e2 , and contemplate a more general interaction potential, 0 0
u(r) = Λe2 /r with 0 ≤ Λ ≤ 1. In other words, we de- (27)
fine the Hamiltonian as ĤMB = T̂ + V̂ + ΛÛ, with the
noninteracting system corresponding to Λ = 0, and the the exchange–correlation energy is given formally as
interacting–electron system to Λ = 1. For a given density
e2 dr dr$
'
distribution n(r) one can consider FHK [n] for intermedi- Exc [n(r)] = n(r) w(r, r$ ) . (28)
ate values of Λ, which leads to the exact expression [34]: 2 |r−r$ |

8
In other words, the exact exchange–correlation energy (GGAs) [37,38], in which the spatial variations of n(r) en-
may be written as in the LDA, Eq. (22), provided ter in a manner which conforms with the sum rule, and
that the corresponding energy per particle, 'xc is in- which have succeeded in reducing the errors of the LDA
terpreted not as a local quantity but is evaluated ac- by a factor which is typically about 4.
cording to the
( density of the exchange–correlation hole, Further improvements in practical expressions for
'xc = (e2 /2) dr$ w(r, r$ )/|r − r$ |. Exc [n] are actively being pursued [39]. One direction
The normalization prop- which may perhaps achieve the accuracy needed for ap-
erty of the exchange–correlation hole, ρxc , carries over plications in chemistry [40], is to use the fact that the
to its Λ–averaged counterpart: exact form of the exchange–correlation hole can be cal-
' culated for Λ = 0 relatively easily, directly from the non-
dr$ w(r, r$ ) = −1 . (29) interacting Kohn–Sham system. There is thus no need
to use an approximation such as the LDA or the GGA
This “sum–rule”, Eq. (29), has been used [35] as the ba- for the low–Λ portion of the integral in Eq. (28). Ulti-
sis for an “explanation” of the relatively high accuracy mately, one hopes that a systematic method of improving
achieved by the LDA: the reference system here (the ho- the approximation would be found, although so far this
mogeneous electron gas) has properties which are also has been an elusive goal.
exact for the inhomogeneous system. More importantly,
it restricts and guides the search for more accurate prac-
tical approximations: expressions which break the sum– C. Successes and failures
rule can not be expected to work well (see below).
Over the years, many different types of applications
of DFT have been developed. This variety evolved be-
B. Refinements of Exc cause knowledge of the electronic ground–state energy as
a function of the position of the atomic nuclei determines
The fact that the LDA achieves a high relative accu- molecular and crystal structure, and gives the forces act-
racy, i.e. can predict the ground–state energy of various ing on the atomic nuclei when they are not at their equi-
systems to within less than a percent, does not mean librium positions. At present, DFT is being used rou-
that the absolute accuracy is sufficient. In particular, tinely to solve problems in atomic and molecular physics,
typical applications in chemistry require that the en- such as the calculation of ionization potentials [41] and
ergy of a molecule be known to within a small fraction vibration spectra, the study of chemical reactions, the
of an electron–volt. Indeed, improving upon the accu- structure of bio–molecules [42], and the nature of active
racy of the LDA is a goal which has been persistently sites in catalysts [43], as well as problems in condensed
pursued. One improvement which is very often imple- matter physics, such as lattice structures [44], phase tran-
mented is the local spin–density (LSD) approximation sitions in solids [45], and liquid metals [46]. Furthermore
[36], which is motivated in part by the fact that the these methods have made possible the development of ac-
exchange–correlation hole is very different for electrons curate molecular dynamics schemes in which the forces
with parallel and with antiparallel spins. In this scheme, are evaluated quantum mechanically “on the fly” [47].
separate densities of spin–up and spin–down electrons are It is important to stress that all practical applications
used as a pair of functional variables: n↑ (r) and n↓ (r), of DFT rest on essentially uncontrolled approximations,
and the Hamiltonian contains separate potentials for spin such as the LDA discussed above. Thus the validity of
up and spin down electrons — a Zeeman–energy mag- the method is in practice established by its ability to
netic field term is introduced. The exchange–correlation reproduce experimental results. A discussion of the ac-
energy per particle is then taken from the results for curacy achieved by DFT, compared to other alternative
a homogeneous spin–polarized electron gas, 'xc (n↑ , n↓ ). approaches, necessarily depends very much on the spe-
The spin dependence allows Hund’s rule to be discussed cific applications one has in mind, as detailed below.
within DFT. For atoms and small molecules, the simplest version
The next degree of sophistication is to allow 'xc to of the LDA already provides a very useful qualitative
depend not only on the local densities but also on the and semi–quantitative picture. It is of course a dra-
rate–of–change of the densities, i.e. to add gradient cor- matic improvement over the Thomas–Fermi model. It
rections. Unfortunately, it was found that such correc- even improves on the more labor–intensive Hartree–Fock
tions do not necessarily improve the accuracy obtained. method in many cases, especially when one is calculat-
In fact, introducing gradient corrections in a straightfor- ing the strength of molecular bonds, which are substan-
ward and systematic manner, by expanding around the tially overestimated in Hartree–Fock calculations. This
uniform electron gas, breaks the sum rule of Eq. (29) can only be considered as a surprising success, keeping in
and is less accurate [35]. This situation led to the de- mind that an isolated atom or molecule is as inhomoge-
velopment of various generalized gradient approximations neous an electronic system as possible, and therefore the

9
last place where one might expect a local approximation side or very near the surface of the metal; this results in
to work. In other words, electronic correlations in such image forces, i.e. a 1/r behavior of vxc (where r is the
systems are in a sense weak, and are on average similar distance from the surface) which is nonlocal. However,
to those of a uniform electron gas (see the discussion of this deficiency can be corrected for “by hand”, yielding
the sum–rule, Eq. (29)). However, the many–body quan- satisfactory results [52].
tum states of such relatively small systems can be solved In general it is useful to note that, in contrast to ap-
for extremely accurately using well–known techniques of proximations using free parameters which are empirically
quantum chemistry, specifically the configuration inter- optimized to fit a certain set of data and may thus be used
action (CI) method [48]. Furthermore, these techniques reliably for interpolation, the LDA and the GGA have
use controlled approximations, so that the accuracy can proved to exhibit a consistent degree of accuracy or inac-
be improved indefinitely, given a powerful enough com- curacy for a wide variety of problems — when applied to
puter, and indeed impressive agreement with experiment a new problem, the results can thus be interpreted with
is routinely achieved. For this reason, most quantum some confidence. One should of course also be aware of
chemists did not embrace DFT at an early stage. the cases for which these approximations are known to
It is in studies of larger molecules that DFT be- fail, such as the image forces mentioned above, and van
comes an indispensable tool [5]. The computational ef- der Waals forces [53], which are important e.g. for bi-
fort required in the conventional quantum chemistry ap- ological molecules. Both of these are manifestations of
proaches grows exponentially with the number of elec- the significance of nonlocal correlations — a nonlocality
trons involved, whereas in DFT it grows roughly as the which is by definition absent from the LDA and its im-
third power of this number. In practice, this means that mediate extensions. These examples of practical failure,
DFT can be applied to molecules with hundreds of atoms, together with the unattractiveness of uncontrolled ap-
whereas using CI, one is limited to systems with only a proximations, spur research towards new and more exact
few atoms. Simply solving the noninteracting problem exchange–correlation energy functionals.
for a complicated molecule may also be prohibitive, and Our discussion would not be complete without men-
various methods are used in order to reduce the problem tioning the existence of many other uses of density–
to a computationally manageable task. Of these, we men- functional methods, for electronic systems and for other
tion the well–known pseudopotential method [14], which physical systems. The former include time–dependent
allows one to avoid recalculating the wavefunctions of DFT, which relates interacting and noninteracting elec-
the inert core electrons over and over again, and the re- tronic systems moving in time–dependent potentials, and
cent attempts to develop “order N” methods [49], which relativistic DFT, which uses the Dirac equation rather
make use of the fact that the behavior of the densities than the Schrodinger equation to calculate the Kohn–
at each point is determined primarily by the atoms in its Sham states (these are reviewed in Ref. [2]). The latter
immediate vicinity, rather than by the whole molecule. include applications in nuclear physics, in which the den-
It is for this problem that more and more accurate den- sities of protons and neutrons and the resulting energies
sity functionals are most obviously needed. To illustrate are studied [54], and in the theory of liquids, as already
this, we quote one sentence from Ref. [38]: “Accurate at- discussed in Sec. II.
omization energies are found [using the GGA] for seven
hydrocarbon molecules, with a rms error per bond of 0.1
eV, compared with 0.7 eV for the LSD approximation V. SUMMARY
and 2.4 eV for the Hartree–Fock approximation.”
The remarkable usefulness of DFT for solid–state In describing density functional theory (DFT) and the
physics was apparent from the outset. For example, the approximations typically implied by its use, it is nec-
lattice constants of simple crystals are obtained with an essary to follow two steps, as was done in considerable
accuracy of about 1% already in the LDA [50]. In such detail in Secs. II and III above. The first step is to in-
applications, the electronic structure of a single unit cell troduce the Hohenberg–Kohn energy, FHK [n(r)]. It is
with periodic boundary conditions is studied; more ambi- equal to the internal energy, i.e. the difference between
tious applications are also common, e.g. a supercell con- the ground–state energy E0 and the potential energy, or
taining many unit cells with a single impurity or defect the sum of the kinetic energy and the interaction energy
[51]. Admittedly, this method is inappropriate for treat- of the electrons. FHK [n] is a universal functional of the
ing some more complicated situations, such as antiferro- density distribution — it applies to atoms, molecules,
magnets or systems with strong electronic correlations. crystals, and all other electronic systems. The existence
In other cases, such as for the work–function of metals, of FHK [n] arises from the fact that each system has not
local approximations such as the LDA obviously miss an only a unique external potential, v(r), as in traditional
important part of the physics: for a point r a short dis- many–body theory, but also a unique density distribu-
tance away from the surface of a metal, the exchange– tion, n(r). Within DFT, the different systems are la-
correlation hole ρxc (r, r$ ) is concentrated at points r$ in- beled by their different electronic densities, n(r), and the

10
potential v(r) is considered as secondary to, and depen- direction has been achieved over the years, most notably
dent on, the primary n(r). In fact, the potential is given with the introduction of the generalized gradient approx-
by the functional derivative v(r) = −δFHK /δn(r). This imation (GGA).
property of FHK [n] can be deduced in two equivalent Whereas the outline just given could apply (with mi-
ways: (a) it is the Euler equation for the Hohenberg– nor modifications) to other introductions to DFT, the
Kohn theorem, which is a variational ( principle stating present discussion was based on an analogy with ther-
that the total energy, FHK [n] + dr n(r)v(r), is min- modynamics. It is well known that in treating situations
imized for a given v(r) by the corresponding ground– where the the number of particles N is constrained, it
state density n(r), and has the ground–state energy E0 is preferable to use it as a free variable, rather than the
as its minimum value; (b) it arises as the conjugate of chemical potential µ. Similarly, one may think of the
the well–known relationship(n(r) = δE0 /δv(r), when the Coulomb interaction as imposing strong constraints on
definition FHK [n] = E0 − dr n(r)v(r) is viewed as a the density distribution n(r) required to achieve low–
(functional) Legendre transform. energy structures in inhomogeneous electronic systems;
The functional FHK [n(r)] discussed above is not known it is thus preferable to use it instead of the potential v(r)
explicitly in terms of its variable, n(r). The second cru- as a free variable.
cial step is to introduce a practical approximation for Three of the advantages of the present approach,
this energy functional. It may be written as a sum of as compared, e.g., with introducing DFT using Levy’s
three terms: (i) the kinetic energy term, which is by constrained–search method [12], are: (a) the density dis-
definition equal to the value FHK [n(r)] would have for tribution n(r) appears here as a natural variable — it
noninteracting electrons; (ii) the Hartree term, which is is conjugate to v(r) through a Legendre transform —
simply a double integral over n(r)n(r$ ); and (iii) the re- whereas in the conventional description of DFT the very
mainder or exchange–correlation term, which in princi- existence of the functional FHK [n] appears to be surpris-
ple contains all of the complicated interaction physics ing and requires some digestion; (b) some of the mathe-
ignored by the first two terms, and must be approxi- matical difficulties encountered in the ground–state the-
mated in practice. By taking a functional derivative, ory are not present in the theory of finite temperature
one finds that the external potential v(r) is given by a ensembles; and (c) using the standard properties of Leg-
corresponding sum of three terms: (i) the so–called ef- endre transforms, one immediately obtains the physical
fective potential, veff (r), which is the potential which a expression for the exchange–correlation energy in terms
system of noninteracting electrons must have in order of the density of the exchange–correlation hole, Eq. (28),
to reproduce the n(r) of the interacting system, (ii) the an expression which serves as the basis for a discussion
electrostatic potential, and (iii) an exchange–correlation of the weaknesses and strengths of the approximations
potential. The simplest widely–used expression for the employed in practice. We hope that the availability of
exchange–correlation term is the local density approx- this type of introduction will help increase the awareness
imation (LDA), which takes the exchange–correlation and understanding of DFT amongst potential users, and
energy–density at each point in the system to be equal especially amongst the general audience of physicists and
to its known value for a uniform interacting electron gas scientists.
of the same density, and results in a parameter–free ap-
proximate description of all electronic systems.
These ideas are implemented by the Kohn–Sham set of Acknowledgments
equations, which consists of a noninteracting Schrodinger
equation involving the effective potential veff (r), the The authors wish to express their gratitude to N.W.
abovementioned relationship between veff (r) and the Ashcroft, W. Kohn, H. Metiu, J.K. Percus, Y. Rosenfeld,
given external potential v(r), and expressions for the den- and G. Vignale for helpful discussions. N.A. acknowl-
sity distribution n(r) and the interacting ground–state edges support under grants No. NSF PHY94-07194, and
energy E0 in terms of the properties of the noninteracting No. NSF DMR96-30452, and by QUEST, a National Sci-
single–particle solutions and of the approximate expres- ence Foundation Science and Technology Center, (grant
sion for the exchange–correlation energy which is in use. No. NSF DMR91–20007).
Using modern computers, the Kohn–Sham equations can
be solved even for systems containing dozens of atoms,
and the results for the ground–state energy are typi-
cally accurate to within a small fraction of a percent. In
contrast to other (much more computationally demand-
ing) methods of calculating the ground–state energy, the
[1] R.G. Parr and W. Yang, Density Functional Theory of
LDA is an uncontrolled approximation, and so there is no
Atoms and Molecules (Oxford, Oxford, 1989).
straightforward path to desired further improvements in
[2] R.M. Dreizler and E.K.U. Gross, Density Functional
the accuracy. Nevertheless, remarkable progress in this

11
Theory: An Approach to the Quantum Many–Body Prob- each point representing a small volume of space Vk , with
lem (Springer, Berlin, 1990). k = 1, 2, . . . , K and with the limit K → ∞ in mind. In
[3] W. Kohn and P. Vashishta “General density functional this picture the functionals such as FHK [n] are simply
theory,” in Theory of the Inhomogeneous Electron Gas, functions of the K variables {nk }, where nk = n(rk ).
S. Lundqvist and N.H. March, Eds., (Plenum, New York, Integration over space is
!now expressed as a sum, e.g.
K
(
1983), pp. 79–147; see also additional chapters in this dr n(r) v(r) becomes V n v , and correspond-
k=1 k k k
book. ingly a δ function near the point rj is replaced by a
[4] The relationship between DFT and Legendre trans- Kronecker δk,j divided by the volume Vj . The func-
forms was discussed mathematically by E.H. Lieb, “Den- tional derivative,
% conventionally defined
& as δF/δn(r) =
sity functionals for Coulomb systems,” Int. J. Quantum limα→0 F [n(r! ) + αδ(r! −r)] − F [n] /α, is then equal to
Chem. 24, 243–277 (1983), and has occasionally been the partial derivative ∂F/∂nk divided by the volume
used in the literature — see, e.g., R. Fukuda, T. Kotani, element Vk . The usual chain rules for derivatives fol-
Y. Suzuki, and S. Yokojima, “Density functional theory low, and can be used in deriving, e.g., the relationship
through Legendre transformation,” Prog. Theor. Phys. n(r) = δΩ/δv(r). A more detailed introduction can be
92, 833–862 (1994). However, we are unaware of any pre- found in C.F. Stevens, “The six core theories of modern
vious references which focus on the pedagogical value of physics” (MIT Press, Cambridge, 1995), pp. 32–38, or in
this relationship. Ref. [1].
[5] W. Kohn, A.D. Becke, and R.G. Parr, “Density func- [11] For alternatives, see R.F. Nalewajski, and R.G. Parr,
tional theory of electronic structure,” J. Phys. Chem. “Legendre transforms and Maxwell relations in density
100, 12 974–12 980 (1996). functional theory,” J. Chem. Phys. 77, 399–407 (1982);
[6] J.S. Rowlinson and F.L. Swinton, Liquids and liquid mix- B.G. Baekelandt, A Cedillo, and R.G. Parr, “Reactiv-
tures, 3rd ed. (Butterworth Scientific, London, 1982); ity indices and fluctuation formulas in density functional
H.T. Davis, Statistical Mechanics of Phases, Interfaces, theory: isomorphic ensembles and a new measure of local
and Thin Films (VCH, New York, 1996). hardness,” J. Chem. Phys. 103, 8548–8556 (1995).
[7] H.B. Callen, Thermodynamics, 2nd ed. (Wiley, New [12] For electronic ground states, the character of the relation-
York, 1985). ship between the potentials v(r) and the density distribu-
[8] The convexity of Ω as a function of µ is easily demon- tions n(r) is more complex. The first Hohenberg–Kohn
strated. For example, consider the free energy at two dif- theorem [13] states that the relationship is one–to–one
ferent values of its argument, µa and µb , and at the for nondegenerate ground states. However, for potentials
midpoint
% µc = &12 (µa + µb ). The fact that Ω(µc ) > v(r) which have (for a given value of N or µ) degener-
1
2
Ω(µa ) + Ω(µb ) , i.e. that for the partition func- ate ground states Ψi , with different density distributions
tion one has Ξ(µc )2 < Ξ(µa )Ξ(µb ), follows directly ni (r), the relationship between v(r) and them is clearly
from Eq. (3) and the inequality of the harmonic and not one–to–one. In fact, there also exist density distribu-
√ tions n(r) which do not correspond to any v(r), (i.e. are
algebraic averages, 21 &(x + y) > % xy, where x =
%
exp (M µa + M µb )/T , y = exp (M ! µa + M µb )/T ,
!
& not ground–state v–representable), e.g. those obtained
and x #= y. This proof generalizes to the functional situ- by interpolating between degenerate ni (r) distributions;
ation, where v(r) replaces µ, see the appendix of M. Va- see, e.g., M. Levy, “Electron densities in search of Hamil-
liev and G.W. Fernando, “Generalized Kohn–Sham den- tonians,” Phys. Rev. A 26, 1200–1208 (1982). Our use of
sity functional theory via the effective action formalism”, thermodynamic ensembles avoids this problem: n(r) is
preprint cond–mat/9702247. Note that the convexity or then equal to the average of the ni (r) in the T → 0 limit,
second derivative of Ω(µ) may vanish in extreme cases when a ground–state degeneracy occurs.
such as at zero temperature limit (because there only In the DFT literature, this problem was solved in a differ-
terms with x = y remain) or in the thermodynamic limit, ent manner (see Ref. [2] for details): an alternative def-
at the point of a first–order phase transition. Mathemat- inition of FHK [n] (denoted by FL [n] below), which does
ical consistency may be restored by restricting attention not explicitly use the potential v(r), but coincides with
to small but finite temperatures and large but finite vol- the definition of Ref. [13] for v–representable distribu-
umes (cf. Ref. [18]). tions, was suggested in M. Levy, “Universal variational
[9] F (N ) thus defined is not identical with that obtained functionals of electron densities, first–order density ma-
by fixing the particle number — the equivalence of the trices, and natural spin–orbitals, and solution of the v–
canonical and grand–canonical ensembles is guaranteed representability problem,” Proc. Natl. Acad. Sci. (USA)
only in the thermodynamic limit. However, we will later 76, 6062–6065 (1979). As Levy’s formulation may also be
focus on the T → 0 limit in which the equivalence is very useful for a pedagogical introduction to DFT (see,
regained because the fluctuations of the particle number e.g., Fig. 3.1 in Ref. [1]), we reproduce it briefly in this
vanish. footnote. The ground state energy is known to be the
[10] Functionals and functional derivatives are not always fa- minimal expectation value of ĤMB , with respect to all
miliar concepts to students, especially undergraduates, wavefunctions Ψ in the Hilbert space of N particles. If
and a discussion of DFT is an excellent opportunity we constrain the range of this search only to wavefunc-
for them to be introduced. For our purposes one may tions which produce a certain n(r) (denoted by Ψ → n),
simply regard the functional variables, n(r) and v(r) we will necessarily find a higher energy, with the ground–
above, as defined on a dense lattice of K points rk , state energy being obtained only if n(r) is the ground–
state density. Levy thus defined, in lieu of our Eq. (6),

12
(
a ground–state energy functional FL [n] + dr n(r) v(r), Fermi theory,” Rev. Mod. Phys. 34, 627–630 (1962).
with FL [n] = minΨ→n &Ψ|T̂ + Û |Ψ', a definition which [24] N.H. March, “Origins: the Thomas–Fermi theory,” in
is valid for any reasonable n(r). Whereas the thermo- Ref. [3], pp. 1–78.
dynamic approach used here clarifies the special role of [25] At finite temperatures, the Fermi–Dirac distribu-
the density distribution n(r), in Levy’s formulation one tion is used for the occupations fi of the Kohn–
could equally well imagine other ways of constraining the Sham
!∞ orbitals,
% and an entropic term,& −T S =
search to other subspaces of the Hilbert space. T i=1 fi log(fi ) + (1 − fi ) log(1 − fi ) must be in-
[13] P. Hohenberg and W. Kohn, “Inhomogeneous electron cluded in Fni ; the approximation used for Exc [n] is also
gas,” Phys. Rev. 136, B864–867 (1964); The general- affected; see, e.g., N. Marzari, D. Vanderbilt, and M.C.
ization to a finite–temperature grand–canonical ensem- Payne, “Ensemble density functional theory for ab ini-
ble was provided almost immediately by N.D. Mermin, tio molecular dynamics of metals and finite–temperature
“Thermal properties of the inhomogeneous electron gas,” insulators,” Phys. Rev. Lett. 79, 1337–1340 (1997).
Phys. Rev. 137, A1441–1443 (1965). [26] For completeness, we write the interacting ground–state
[14] M.C. Payne et al., “Iterative minimization techniques for energy E0 explicitly in terms of the Kohn–Sham eigen-
ab initio total–energy calculations: Molecular dynamics values and the density distribution:
and conjugate gradients,” Rev. Mod. Phys. 64, 1045– N '
1097 (1992). " % &
E0 = "i − dr n(r) veff (r) − v(r) +Ees [n(r)]+Exc [n(r)] .
[15] J.D. van der Waals, Ph.D. Thesis, University of Lieden
i=1
(1873).
[16] Clearly, multiple local minima occur only if −v = [27] Interestingly, the work function of a metal surface is equal
∂(nf )/∂n = f + nf ! varies nonmonotonically with n, i.e. to that of the Kohn–Sham system. Using the fact that
possesses regions of negative slope. This condition is iden- both ϕ(r) and vxc (r) decay as 1/r at large distances away
tical with the conventional one of negative compressibil- from the system, and using the usual convention of tak-
ity, κ < 0, because κ = ∂P/∂n = (∂/∂n)(n2 ∂f /∂n) = ing v(r) and veff (r) to vanish at infinity, one finds that
2nf ! + n2 f !! , which is just n times the abovementioned Eq. (19) reduces to a statement of the equality of the
slope (the density n is always positive). Note that these chemical potentials for the interacting and the Kohn–
negative slopes exist only in the phenomenological model, Sham system.
Eq. (12); a proper statistical mechanical evaluation of the [28] See, e.g., S.B. Nickerson, and S.H. Vosko, “Prediction of
partition function would exhibit regions of phase separa- the Fermi surface as a test of density–functional approxi-
tion instead. mations to the self–energy,” Phys. Rev. B 14, 4399–4406
[17] W. Kohn and L.J. Sham, “Self–consistent equations in- (1976).
cluding exchange and correlation effects,” Phys. Rev. [29] D. Mearns, “Inequivalence of the physical and Kohn–
140, A1133–1138 (1965). Sham Fermi surfaces,” Phys. Rev. B 38, 5906–5912
[18] We consider small but positive temperatures, T → 0, (1988).
rather than the situation at T = 0, where the function [30] G.D. Mahan, “GW approximations,” Comm. Cond. Mat.
N (µ) can take only integer values and becomes discon- Phys. 16, 333–354 (1994).
tinuous (if one insists on working at T = 0, the func- [31] O. Gunnarsson, and B.I. Lundqvist, “Exchange and cor-
tional(derivative, e.g. in Eq. (21), must be redefined so relation in atoms, molecules, and solids by the spin–
that dr δn(r) = 0). For a discussion, see e.g. J.P. density formalism,” Phys. Rev. B 13, 4274–4298 (1976).
Perdew, R.G. Parr, M. Levy, and J.L. Balduz, “Density– [32] E.P. Wigner, “Effects of electron interaction on the en-
functional theory for fractional particle number: deriva- ergy levels of electrons in metals,” Trans. Faraday Soc.
tive discontinuities of the energy,” Phys. Rev. Lett. 49, 34, 678–685 (1938).
1691–1694 (1982). [33] It is interesting to compare the Kohn–Sham equations
[19] Note that this definition differs from the conventional with the well–known Hartree approximation. In the lat-
one for exchange–correlation energy, because Fni [n] is the ter scheme each electronic orbital is calculated with a
kinetic energy of a system of noninteracting particles with different electrostatic potential, which is due only to
the density n(r), rather than the kinetic energy of the the other electrons, and not to the total density dis-
interacting electron system. tribution. There exist also hybrid schemes, called “self–
[20] L.H. Thomas, “Calculation of atomic fields,” Proc. interaction corrected” schemes, where such different po-
Camb. Phil. Soc. 33, 542–548 (1927). tentials are used and an extra potential vxc is included.
[21] E. Fermi, “Application of statistical gas methods to elec- Although such schemes have given good results for some
tronic systems,” Accad. Lincei, Atti 6, 602–607 (1927); systems (e.g. crystals with localized electrons, i.e. insu-
“Statistical deduction of atomic properties,” ibid, 7, 342– lators), they do not follow the logic of either the Hartree
346 (1928); “Statistical methods of investigating elec- or the DFT method. See J.P. Perdew and A. Zunger,
trons in atoms,” Z. Phys. 48, 73–79 (1928). “Self-interaction correction to density-functional approx-
[22] L. Spruch, “Pedagogic notes on the Thomas–Fermi the- imations for many-electron systems,” Phys. Rev. B 23,
ory (and some improvements): atoms, stars, and the 5048–5079 (1981).
stability of bulk matter,” Rev. Mod. Phys. 63, 151–209 [34] J. Harris, “Adiabatic–connection approach to Kohn–
(1991). Sham theory,” Phys. Rev. A 29, 1648–1659 (1984).
[23] E. Teller, “On the stability of molecules in the Thomas– [35] O. Gunnarsson, M. Jonson, and B.I. Lundqvist, “De-

13
scription of exchange and correlation effects in inhomoge- 5684 (1990).
neous electron systems,” Phys. Rev. B 20, 3136 (1979). [53] W. Kohn, Y. Meir, and D.E. Makarov, “Van der Waals
[36] See, e.g., S.H. Vosko, L. Wilk, and M. Nusair, “Accu- energies in density functional theory,” Phys. Rev. Lett.
rate spin–dependent electron liquid correlation energies 80, 4153-4156 (1998).
for local spin density calculations: a critical analysis,” [54] See, e.g., C. Speicher, R. M. Dreizler, and E. Engel,
Can. J. Phys. 58, 1200–1211 (1980), which provides a “Density functional approach to quantumhadrodynam-
widely–used version of "xc (n↑ , n↓ ). ics: Theoretical foundations and construction of ex-
[37] D.C. Langreth and J.P. Perdew, “Theory of nonuniform tended Thomas-Fermi models,” Ann. Phys. (San Diego)
electronic systems I: Analysis of the gradient approxima- 213, 312–354 (1992).
tion and a generalization that works,” Phys. Rev. B 21,
5469–5493 (1980); J.P. Perdew and Y. Wang, “Accurate
and simple density functional for the electronic exchange
energy: generalized gradient approximation,” Phys. Rev.
B 33, 8800–8802 (1986).
[38] J.P. Perdew et al., “Atoms, molecules, solids, and sur-
faces: applications of the generalized gradient approxi-
mation for exchange and correlation,” Phys. Rev. B 46,
6671–6687 (1992).
[39] For example, see the proceedings of the VIth Interna-
tional Conference on the Applications of Density Func-
tional Theory. (Paris, France, 29 Aug.—1 Sept. 1995),
Int. J. Quantum Chem. 61(2) (1997).
[40] A.D. Becke, “Density–functional thermochemistry III:
The role of exact exchange,” J. Chem. Phys. 98, 5648–
5652 (1993); “Density–functional thermochemistry IV: A
new dynamical correlation functional and implications
for exact–exchange mixing,” ibid, 104, 1040–1046 (1996).
[41] R.O. Jones and O. Gunnarsson, “The density functional
formalism, its applications and prospects,” Rev. Mod.
Phys. 61, 689–746 (1989).
[42] M.D. Segall et al., “First principles calculation of the
activity of cytochrome P450,” Phys. Rev. E 57, 4618–
4621 (1998).
[43] R. Shah, M.C. Payne, M.-H. Lee and J.D. Gale, “Un-
derstanding the catalytic behavior of zeolites — a first–
principles study of the adsorption of methanol,” Science
271, 1395–1397 (1996).
[44] M.J. Rutter and V. Heine, “Phonon free energy and
devil’s staircases in the origin of polytypes,” J. Phys.:
Cond. Matter 9, 2009–2024 (1997).
[45] W.-S. Zeng, V. Heine and O. Jepsen, “The structure of
barium in the hexagonal close-packed phase under high
pressure,” J. Phys.: Cond. Matter 9, 3489–3502 (1997).
[46] F. Kirchhoff et al., “Structure and bonding of liquid Se,”
J. Phys.: Cond. Matter 8, 9353–9357 (1996).
[47] R. Car and M. Parrinello, “Unified approach for molecu-
lar dynamics and density–functional theory,” Phys. Rev.
Lett. 55, 2471–2474 (1985).
[48] R. McWeeny and B.T. Sutcliffe, Methods of Molecular
Quantum Mechanics (Acad. Press, London, 1969).
[49] W. Kohn, “Density functional theory for systems of very
many atoms,” Int. J. Quant. Chem. 56, 229–232 (1994).
[50] V.L. Moruzzi, J.F. Janak, and A.R. Williams, Calculated
Electronic Properties of Metals (Pergamon, New York,
1978).
[51] G. Makov and M.C. Payne, “Periodic boundary condi-
tions in ab initio calculations,” Phys. Rev. B 51, 4014
(1995).
[52] See, e.g., Z.Y. Zhang, D.C. Langreth, and J.P. Perdew,
“Planar–surface charge densities and energies beyond the
local–density approximation,” Phys. Rev. B 41, 5674–

14

You might also like