Вы находитесь на странице: 1из 35

INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY, VOL XXIV, 243-277 (1983)

Density Functionals for CouIomb Systems


ELLIOTT H. LIEB
Departments of Mathematics and Physics, Princeton University, P.O.B. 708, Princeton,
New Jersey 08544, U.S.A.

Abstract
This paper has three aims: (i) To discuss some of the mathematical connections between N-particle
wave functions $ and their single-particle densities p ( x ) . (ii) To establish some of the mathematical
underpinnings of “universal density functional” theory for the ground state energy as begun by
Hohenberg and Kohn. We show that the HK functional is not defined for all p and we present
several ways around this difficulty. Several less obvious problems remain, however. (iii) Since the
functional mentioned above is not computable, we review examples of explicit functionals that have
the virtue of yielding rigorous bounds to the energy.

Introduction
It is a pleasure to dedicate this article to Laszlo Tisza on the occasion of his
seventy-fifth birthday. As a colleague at MIT he was a source of inspiration and
encouragement, especially in drawing our attention to the importance of careful
and precise thought in mathematical physics. The subject, if not the content, of
this article may therefore not be inappropriate in a book dedicated to Professor
Tisza (see the Acknowledgment).
The idea of trying to represent the ground state (and perhaps some of the
excited states as well) of atomic, molecular, and solid state systems in terms of
the diagonal part of the one-body reduced density matrix p ( x ) is an old one. It
goes back at least to the work of Thomas [l]and Fermi [2] in 1927. In 1964
the idea was conceptually extended by Hohenberg and Kohn (HK) [3]. Since
then many variations on the theme have been introduced. As the present article
is not meant to be a review, I shall not attempt to list the papers in the field.
Some recent examples of applications are Refs. 4 and 5 . Some recent examples
of theoretical papers which will play a role here are Refs. 6-12. A bibliography
can be found in the recent review article of Bamzai and Deb [13].
This article has three aims:
(i) To discuss and prove some of the mathematical relations between N-
particle functions J/ and their corresponding single-particle densities p.
(ii) To discuss the mathematical underpinnings of general density functional
theory along the lines initiated by HK. In that theory a universal energy functional
F(p) is introduced. Despite the hopes of HK, F(p ) is not defined for all p because
it is not true (see Theorem 3.4) that every p (even a “nice” p ) comes from the
ground state of some single-particle potential ZI ( x ) . This problem can be remedied
by replacing the HK functional by the Legendre transform of the energy, as is
done here. However, the new theory is also not free of difficulties, and these

@ 1983 John Wiley & Sons, Inc. CCC 0020-7608/83/090243-35$04.50


244 LIEB

can be traced to the fact that the connection between z, and p is extremely
complicated and poorly understood.
(iii) To present briefly another approach to the ground state energy problem
by means of functionals that, while not exact, are explicitly computable and yield
upper and lower bounds to the energy.
The analysis in this paper gives rise to many interesting open problems. It
is my hope that the incompleteness of the results presented here will be partly
compensated if others are encouraged to pursue some of the questions raised
by them.
It is not my intention to present a brief for HK theory. However, it deserves
to be analyzed for at least two reasons: The HK theory is used by many workers
and it gives rise to some deep problems in analysis. While it is my opinion that
density functionals are a useful way to approach Coulomb systems, there are
other approaches besides the HK approach [e.g., see (iii) above]. Apart from the
difficulties mentioned above, the HK approach may be too general because all
potentials have to be considered. Coulomb potentials are special and do lend
themselves to a density functional approach; for example, Thomas-Fermi theory
is asymptotically exact as 2 + 00 (see Sect. 5E and Ref. 14). In addition to this
question of generality there is also the crucial point that the “universal functional”
is very complicated and essentially uncomputable. If one is going to make
uncontrolled approximations for this functional, then the general theory is not
very helpful.
It is a pleasure to thank Barry Simon for some very helpful conversations
and the proofs of Theorems 4.4 and 4.8. I also thank Hai’m Brezis for the proof
of Theorem 1.3.

1. Single-Particle Densities
The first order of business is to describe the single-particle densities of interest.
For simplicity we confine our attention to three dimensions whenever dimension-
ality is important. z = (x, (+) will denote a space-spin variable, that is, x E R3 and
(+ E (1, . . . , q}. q = 2 for electrons, of course, but one might wish to consider

q = 1, which would mean that a ferromagnetic state is under consideration. We


use the notation

Let (L = J / ( z l., . . , z N ) be an N-particle function (which may be complex


valued). To simplify notation we will not indicate N explicitly except where
needed. However, the condition of fixed N is crucial and frequently glossed
over. The density functionals that will be introduced later are explicitly N
dependent in a highly nontrivial way (see Sect. 4A). (I is assumed to be nor-
malized :
COULOMB DENSITY FUNCTIONALS 245

(with = d z 1, . . . , dzN) and to have finite kinetic energy, that is,


c[
N
~(rf/)= Ivirf//2<a. (1.3)
i=l

Notes. f € L P means that f is a function satisfying llfllP=u/flP}'/P


<m. f € H 1
means that f and each component of Vf are in L 2 . Thus, the above $ E H ' . If
f is differentiable everywhere, then Vf and T ( f )= jlVfI2 are well defined. Other-
wise, the correct definition of T ( f )for functions in L2(Iwm)is

T ( f )= ( 2 7 ~ - k21f(k)12
~ dk, (1.4)

where f is the Fourier transform of f. Since EL', f exists and f~ L 2 . H' is a


Hilbert space with inner product (f,g ) = f * g + Vf * .Vg.
In most of the following it will be assumed that rf/ satisfies the Pauli principle,
that is, $ is antisymmetric. However, some of the theorems are easier for
symmetric (i.e., bosonic) $ with q = 1, and occasionally this will be mentioned
explicitly. In either case, the symmetry implies that

We define the single-particie density to be [see Eq. (A.l)]


2 dx2...dx~.
p ( x ) = N C [ l$((x,Vl),...,(x~,c~))l
(1.6)
U

Notice that j p ( x ) dx = N, not 1.


Determinants. If 4 1 ( z ) ,. . . , ~ N ( z are
) orthonormal functions, we can form
the determinantal
$(zI, . . . ,z N ) = ( N ! ) -' l 2det { 4 i ( z j ) } , (1.7)
which is normalized. Then

i=l u=I J

Returning to the general case, the finiteness of T ( $ )implies the following [15].
Theoreml.l. p(x)l/' €L2(rW3) t h a r i s , p ( ~ ) €H'([W3'
and Vp(x)"' €L2([W3), ~/*
Moreover, j ( v p 1/2)2 s T ( Q ) .
Proof. p 1 / 2 ~ L 2 ( 1 w 3 )because f p = N . Now V p ( x )= N f' ( V l $ ) * ~
N I'$*Vl$, where j' means the integral in (1.6). By the Schwarz inequality,
t
[
[ V P ( X ) I 2 ~ 4 N P ( X ) IV1$I2.
Thus ( V p 1/2)2 dx = t j (Vp)'p-' dx G T ( $ ) .
246 LIEB

We know p E H1(R3) '" = { f l f ~ L2, V ~ L2}.


E (Here we use the standard
convention that {AJC}means the set of A such that condition C holds.) To
discuss the converse of Theorem 1.1 some definitions are useful.
Definition. .aN = ( p I p ( x ) a O , p 1 / 2 ~ H 1 ( R 3 ) , j p ( x =
) dNx} .
Definition. B N = ( P I p ( x ) a O , ~ p ( x ) d=x N , p EL'(R~)}.
g Ncontains .aN by the Sobolev inequality (see Ref. 16) because i f f € H1(R3),
then

(1.10)

Equation (1.10) is true only in three dimensions, but analogous inequalities hold
in other dimensions. By Theorem 1.1,
T ( * )3 3(~/2)4'311p113.
g Nis clearly a convex set; that is, if p1 and p2 E BN,then p = Apl+ (1- A)p2 E
BN for all 0 S A s 1. $N is also convex by the same proof as in Theorem 1.1;
that is, by the Schwarz inequality
1/2 2 1/2 2
(Vp)2c4p[A(Vpi +(1-A)(Vp2 ) 1.
In particular, the functional 5 [Vp1/2]2 is convex. The convexity of SN will be
important in Sect. 3.
Definition. A function (or functional) f is convex if
f ( A x + (1- A ) Y 1 < A f ( x ) + (1- A M Y 1
for all 0 < A s 1 and all x and y in the domain of f .
Theorem 1.2. Suppose p E ~ N Then, for either Bose or Fermi statistics there
exists a 4 (which is a determinant in the ferrnion case) such that (1.6) holds and,
moreover,

T ( 4 )s J [Vp'/2(x)]2 dx (bosons), (1.11)

' [
T ( + )s ( ~ T ) ~ N [Vp1/2(x)]2dx (ferrnions). (1.12)

Proof. For bosons the proof is easy; simply take

For fermions the construction is much more complicated. Some ideas from Ref.
',
17 will be used in the following. Write x = (x x2, x 3 ) and define
COULOMB DENSITY FUNCTIONALS 247

Then f is monotone increasing from 0 to 27r. For k = 0, . . . ,N - 1 define


d k ( x )= [p(x)/N]'/' exp [ikf(x')].
It is easy to check that the q5k are orthonormal functions in L2(R3).[First do
the x2 and x3 integrations and then note that the overlap integral is of the
form Jmm (dfldx')exp [i(Ak)f(x')] dx' = {exp [i(Ak)f(m)]-exp [i(Ak)f(-~)]}/
i(Ak) = 0. Furthermore,

(1.13)

with

As in Theorem 1.1, we conclude that

gEH'(R') and j ( 2 ) ' d s GI(Vp'/')'=A.


Since

we conclude by the Schwarz inequality that g(s)" G 4[J g2][1 ( d g l d y ) ' ] . Thus, the
last term in (1.13) is less than 4(2.rrk/N)'N2A. Finally, we take $ to be a
determinant as in (1.7) using the functions q5 ( x ) x (spin up). Equation (1.12)
follows by summing on k.
Theorem 1.2 is closely related to the results of Gilbert [8] and Harriman [9].
For fermions, the extra factor N 2 in (1.12) is noticeably different from the
factor N o in Theorem 1.1. Although (1.12) can be improved, it is not easy to
do so. In any case, the conclusion is that the map from $ to p 1 l 2 given by (1.6)
is a map from H ' ( I R ~onto
~ ) H1(R3). But the map is clearly not 1: 1; different
$'s can give the same p.
Question 1. Is this map continuous as a map from H ' ( R 3 N )to H'(R3)?That
is, if r// is fixed and $, is a sequence (with corresponding p and p ) such that
J 14 - $,I* + 0 and J JV$- V$,I2 + 0, does it follow that Ip '/' - p i ' 2 l2 + 0 and
J IVp'/2-Vp~/212+~?
Question 2. Although the map is not invertible (since it is not 1: l ) , we can
ask the following: Given a sequence p ; j 2 that converges to pl/' in the above
H ' ( R 3 )since, and given some r// satisfying (1.6) for p, does there exist a sequence
4, [related to p1 by ( 1 . 6 ) ] that converges to $ in the above H ' ( R 3 )sense? [This
is equivalent to the statement that the map r// * I 2 is "open," that is, the map
takes open sets in H ' ( R 3 N )into open sets in H? (R3).]
248 LlEB

Intuitively, the answer to both questions should be affirmative. The continuity


can indeed be proved, but the proof is not entirely elementary. A proof of
Theorem 1.3, due to H. Brezis, is given in the appendix.
Theorem 1.3. The map 4 - p l f 2 given by (1.6) is continuous as a map from
ro H’(R3).
H1(lW3N)
I cannot offer any proof of the openness of the map, however. The fact that
these questions do not have simple answers should serve as a warning that the
connection between $ and p is not as obvious as one might intuitively think.

2. Single-Particle Density Matrices


If $ is given as before, we can define the single-particle density matrix

This definition is different from the usual one because we sum on a1in (2.1j.
Usually one defines the quantity q ( x , a ;x ‘ , (T‘), so our y ( x , x ’ ) = C, q ( x , a ;x ’ , c).
Clearly, p ( x > = y ( x , x ) .
Theorem 2.1. y satisfies
(i) T r y = f y ( x , x ) d x = N ,
(ii) As an operator, 0 S y s q1, forfermions ;that is, 0 S (f,yf) G q (f,f).For bosom,
0s y G N I .
Proof. (i) is “obvious” but not trivial. The point is that if an operator K is
given, then its kernel K ( x ,y ) is defined only almost everywhere. In particular,
K ( x ,x ) can be anything. Thus, Tr K need not be f K ( x , x j dx. However, (i) can
be proved from (2.1). This is left as an exercise. To prove (ii let M(x,x’)=
b
f ( x ) f ( x ‘ ) *be a one-particle operator with (f,f)= 1. Then A =Ct=lM(x,,x : ) has
as its largest eigenvalue on the antisymmetric space the value q. Moreover, A
is clearly positive semidefinite. Thus, 0 c (f, yf) = Tr y M = ($, A*) s q.
Definition. Let y ( x , y ) be any kernel. y is said to be admissible if T r y = N
and 0 5 y 5 qI (fermions) or 0 Iy IN (bosom). The set of admissible y is clearly
convex; that is, if y and 6 are admissible, then so is a y + (1- a ) S for O-ia I1.
Now we come to a subtle point. If y is an admissible operator, we can ask
two questions:
Question 3. Does an N-particle density marrix l? always exist, where r =
T ( t l , . . . ,t Nt;i , . . . ,zb), so that y is given by (2.1) with (c.c/* replaced by r?
(r is a density matrix if 0 Ir and Tr I‘= 1. r must also satisfy the appropriate
symmetry.)
Question 4. Does a 4 always exist so that (2.1) holds; that is, can r be chosen
to be a pure state, namely, r = $)($?
COULOMB DENSITY FUNCTIONALS 249

The answer to question 4 is No! (for fermions). For bosons, the answer is Yes.
The proof of question 3 (which we now call Theorem 2.2) has been known
for a long time. An explicit construction is given in Ref. 26. An example in
which r fails to be of the form (I,)(+, for N = 2 and q = 1, is the case in which
4,
y has three nonzero eigenvalues 1, i. To see this, let the normalized eigenvec-
tors of y be f ( x ) , g ( x ) , and h ( x ) ,respectively; that is,
y ( x ,x ' ) = f ( x ) f ( x ' ) * + $g(x)g(x')* + ih (x)h(x ')*
Let A = -y(xl, x i ) - y ( x 2 , x;) be an operator on the antisymmetric states. Its
lowest eigenvalue is -1 - 1/2 = -3/2, which is doubly degenerate. If I'=(I,)(@,
then (I, must be a ground state since Tr I'A = -Tr y 2 = -1 - 1 /4 - 1 / 4 = -3/2.
But every ground state is of the form (I, = 2-*12 det ( f , p ) , where p = ag + bh,
l ~ / ' + 1 6 1 ~ = 1But
. then y = f ) ( f + p ) ( p , and this is never of the form f ) ( f +
k > ( g+ kh)(h.
The moral of all this is the following: On the one-particle level we can study
density matrices y ( x , x') or densities, p ( x ) = y ( x , x ) . The former do not always
come from pure states $)((I,. The latter do, as Theorem 1.2 shows. While y is
more complicated than p (it has two variables), it has the distinct advantage that
the map I'-y is linear! The map 4 - p is nonlinear, and this, as will be seen,
is the source of some difficulty.
The relation among (I,, r, y, and p can be summarized by the following
diagram:
(I,-r++y-p, (2.2)
by which we mean (i) the map 4-r = (I,)($, (ii) y by (2.1) with $(I,* replaced
by r, (iii) y Hy ( x , x ) = p ( x ) . (ii) and (iii) are linear while (i) is nonlinear.
Notation. We shall use the symbol $ - p (or any other combination such as
y~ p to indicate
) that (I, and p are related by the above maps.
Technical remarks. Since y is self-adjoint and trace class, it can always be
written in the form

where the f i are orthonormal and 0 S A j z q (fermions), 0 s A j s N (bosons). If


T E H ' ( I W ~ by
~ ) which
, we mean Tr-Ci,l Aircoo, then each f€H1(rW3) (see
Theorem 1.1). Although y ( x , x ) is not a priori well defined, as stated before, it
is well defined almost everywhere (in rW3) by (2.3) and
m
N=Try= 1 Ai. (2.4)
j=l

3. General Density Functional Theory


The problem that will concern us in calculating the ground state energy for
N electrons interacting with each other via a repulsive Coulomb potential
250 LIEB

/ x i-xi]-' and also interacting with a single-particle potential v(x). If v = 0, the


Hamiltonian is

(3.1)

where K is the kinetic energy operator


N
K=- c
i=l
Ai (3.2)

in units in which h2/2m = 1. Also of interest is the case where Ho = K alone


(see Sec. 4C). Recall that N is fixed and will not be mentioned unless necessary.
Also, to simplify matters we shall confine our attention in the following to
fermions. However, many of the following results have obvious analogs for
bosons.
The total Hamiltonian is
H,, =Ho+ V, (3.3)
where
N
v = i c= l U(Xi). (3.4)

The ground state energy E ( v ) is defined to be


E ( v )= inf ((4,HU4)I46 WN}, (3.5)
where
WN = { ~ ~T(II/)<co}.
~ ~ ~ ~ ~ = ~ (3.6)
,
Technical remark. Something should be said about the meaning of (4,H&)
and about the class of v's under consideration. We shall always interpret (4,H,,$)
in the sense of a quadratic form; in particular, this means that (4,K 4 )= T ( 4 ) .
+
It is not assumed that A+ E L 2 .Since E H ' , it is easy to prove that (+, \xi - x,\-'+)
is finite for all i # j . The part containing v is Jp(x)v(x) dx. As p EL' and p E L 3
(since Vp"' E L'), p E L p for all 1 s p 5 3. The integral is then well defined if
v E L3I2+ L". This means that we consider v ' s that can be written as v = v3/2 + vm
with ~ 3 1 E2 L3/' and with Ival a bounded function. This choice precludes v's that
go to 00 as jxJ+co, such as the harmonic oscillator potential. Unbounded
potentials can also be handled by the methods given here, but then we have to
place additional restrictions on p so that vp makes sense. We restrict ourselves
here to L3I2+ L" for simplicity of exposition. The class includes Coulomb
potentials because Jxl-'= @(x)IxI-l+[l-@(x)]IxI-'with 6(x) = 1if 1x15 1, O(x) =
0, Ix I > 1.The two terms on the right are in L3" and in L", respectively.
L3l2+ L" is a Banach space with the norm

IIV II = inf M 1 3 / 2 + Ilk ll"k + h = V I - (3.7)


COULOMB DENSITY FUNCTIONALS 25 1

Technical remark. BN is a subset of the Banach space X = L 3f l L'. X * , the


dual of X, is Y = L3/' + L". However, the dual of Y is not X because while L"
is the dual of L 1 ,L' is not the dual of L". However, X c Y * . The duality will
be useful.
There may or may not be a minimizing 4 for (3.5), and if there is one it may
not be unique (for bosons it is unique because it is a positive function). Any
minimizing 4 (called a ground state) would satisfy
Hu4 = E ( u ) 4 (3.8)
in the distributional sense. The proof of this assertion is not difficult. For example,
a minimizing 4 will not exist if u is an attractive square well and if N is too
large; the extra, unbound electrons will simply "leak away" to infinity. In such
a case, E ( u ) would still have physical significance. It would be the ground state
energy for fewer than N particles.
There are three simple, but important, properties of E ( v ) :
Theorem 3.1. (i) E ( u )is concaue in u : that is,
E ( u )>aE(u1)+ (1 -a)E(uz), (3.9)
for all u1, u2, 0 5 a 5 1 and u = aul + (1- a ) u z .
(ii) E ( u ) is monotone decreasing: that is, if u l ( x ) s u z ( x ) for all x , then
E(u1) s E ( u 2 ) .
(iii) E ( u ) is continuous in the L3/'+L" norm and is, moreouer, locally Lip-
schitz. In particular, E ( u ) is finite.
Proof. (i) If 4~ WN, then
(4,HU4)= a(4,HUI4)+ (1- a h % H"&) z a E ( u 1 )+ (1- a ) E ( u z ) .
2 (4,Hu14)
(ii) (4,Hu24) ~E(u1).
(iii) Fix uo and let S = u -uo. We want to show that when llSll s L / 3 ,
IE(u)-E(uo)l<C1181( for some C, independent of u. [Here, L is the constant
in (l.lO).] Since E ( u ) is concave, it is sufficient to show that for some fixed D,
E ( u ) - E ( u o ) r Dwhenever IlSll=L/3; because if 0 1 y s 1,
Y [ E( 0 0+ - E (uo)] 5 E (uo + Y S -E (00)5 Y [ E(00) - E (00- 11.
Let E(u, i)denote (3.5) with K replaced by K / 2 . Then
~ ( u ) r ~ ( u o , i ) + i(4,
n f[ i ~ + z ~ ( x i ) I 4 ) .
The last term is bounded by -LN/2 because S = g + h with Ilg))3/2<L/2and
I(h < L / 2 . Thus,

by (1.10) andTheorem 1.1. Finally,notethatE(uo, 4)-


But (4,K 4 ) / 2 2(L/2)11p1)3
E(vo)is a constant, D', independent of u.
252 LlER

Now we begin the study of density functional theory in the manner of


Hohenberg and Kohn. Their work is based on the following theorem [3]:
Theorem 3.2. Suppose $l (respectively,(CI2) is a ground state for u 1 (respectiuely
u2) and v 1 # u2+constant. Then p l f p Z .
Proof. Suppose p 1 = p2 = p. $l # (CI2 because they satisfy different Schrodinger
equations, (3.8). [Note. To prove this we must know that u 1 4 = u2$ implies that
v 1 = u2. This, in turn, requires that $(x) does not vanish on a set of positive
measure. This technical point is discussed in remark (ii) preceding Theorem 3.5.1
Moreover, $2 (respectively, +bl) does not satisfy (3.8) for v1 (respectively, v 2 ) .
Therefore,

) E ( u l ) +J ( u 2 - u l ) p . This is a contradiction.
Likewise, E ( u Z <
Hohenberg and Kohn assume that every p comes from some 4 that is a
ground state for some u. For such p they define the functional

(3.10)

and we shall retain this definition for p E d N , where


dN= (pip comes from a ground state}. (3.11)

dN# YN,as remarked earlier, and it is not convex (see Theorem 3.4)! The
definition given by (3.10) requires Theorem 3.2, according to which there is a
unique u (up to a constant) associated with p. We can also define
'VN= {ulH, has a ground state}. (3.12)

It then follows easily that for u E 2',

~ ( u=)min [ FHK(P) + I vp IP E dN}. (3.13)

This is the H K uariationai principle, but it is important to note that it holds only
for u E 'VN,which is unknown, and that the variation is restricted to the unknown
set d N .
We also do not know what F H K is, and that is a very serious problem. But
there are also conceptual problems, which will be addressed here.
If F is to be used in a variational principle, it is clearly desirable that F be
a convex functional. In particular, it should be defined everywhere on YN, or at
least on some known convex subset of YN.
The domain of F H K (i.e., d N is
) not all of 9Nand it is not convex. This last
fact is closely connected with the following difficulty: One can define a functional
COULOMB DENSITY FUNCTIONALS 253

for all p in 9, by*

It then follows trivially that

fib)= F ~ ( P ) , if p EdN. (3.16)


So far, so good. The difficulty is that fi is not convex either. However, fi has
one important property that is proved in the appendix.
Theorem 3.3. For each p in 9, there is Q E WN such that fib)= (4, H04).
In other words, the infimum in (3.14) is a minimum.
The following functional F is one choice for “the density functional” that
remedies the difficulties mentioned so far:

We shall explore the properties of F, but it, too will be seen to have subtle
difficulties of its own.
Remarks. (i) (3.17) defines F ( p ) for all p € g Nnot
, just $N, provided F is
interpreted in the extended sense as a function that can have the value f m . In
fact, (2.17)defines F on the much larger set X = L3f l without the restrictions
L’,
p ( x ) 2 0 and 5 p = N. As Theorem 3.5 shows, however, it is only necessary to
consider F on the convex subset 9 , of X.
(ii) Recall that F depends explicitly on N through E.
(iii) Since F is the supremum of a family of linear functionals, it is convex.
(iv) Theorem 3.8 shows that F(p)= +a if p&gN.There is an alternative
definition of F,namely, F’, by which F’ is finite on the set
9 L = ( p I p ( x ) r O and V p 1 / 2 ~ L 2 } ,
without requiring J p = N. This is

F’(0)= 0. (3.18)
It is easy to check that the convexity and lower semicontinuity (a concept to be
defined later) of F carry over to F’.This definition has the virtue that F‘is finite
* Levy [lo] also defined P ( p ) which he called Q, and derived (3.15). He did not prove Theorem
3,.3, but assumed the existence of a minimizing $. Also, he did not establish the connection between
F and the Legendre transform, F (Theorem 3.7). In Ref. 11, Levy proved Theorem 3.4(ii),
independently and virtually at the same time as myself, using essentially the same construction. See
Ref. 12 for additional remarks about Q.
254 LlEB

on a dense subset of the set of nonnegative functions in X.However, this does


not change the theory in any important way, so we shall continue to use the
definition given by (3.17).
(v) Other characterizations of F, directly in terms of F,are given in Theorem
3.7, and in Eqs. (4.5)-(4.7).
There is an obvious relation between F and F,namely,

~ ( p ) SF@) for all p E 9 N , (3.19)

since E ( v )s F ( p )+ J vp for all p E YN.Furthermore, since F is convex and F is


not convex (by Theorem 3.4), there are p’s in 9~ for which F ( p )<E(p).
First we prove that not all p’s come from ground states. The essential
ingredient is the existence of v with a degenerate ground state. (Such v’s,
incidentally, preclude the existence of a map v ~ p . )
Theorem 3.4. Let N > q = number of spin states. Then
(i) F ( p ) )is not convex
(ii) There exists a p E 9 N that does not come from a ground state J/.Moreover
this p is a convex combination of p’s that do come from a ground state.
Proof. Let v be a spherically symmetric potential having a ground state and
with the property that its ground state has orbital angular momentum L z 1. We
assume the degeneracy is no greater than necessary, namely M = 2L + 1. The
orthonormal ground states are J/’, . . . ,1,4~ and 4, ++pi. Under simultaneous
rotation of all N coordinates, they transform as a basis for the M-dimensional
irreducible representation of O(3).The following fact is easy to prove: (a) If
p = M-’Zpi, then p ( x ) is spherically symmetric: that is, p depends only on r = ( x 1.
A second fact that will be needed is (b): if (b is any ground state (and hence
a linear combination of the &) and d, ~ p then , p is not spherically symmetric.
This fact must follow from some group-theoretic agreement, but I have not
found one. However, it is not hard to see that (b) is equivalent to (c): There
exists a perturbation of v, v ( x ) Hv (x) + A w (x), with w bounded and of compact
support, so that to first order in A the M-fold degeneracy is broken. Such pairs
u and w certainly exist, so we can henceforth assume that (b) holds. [A proof
that a v satisfying (b) exists is the following. First, take the case that H o = K ,
that is, independent particles. The ground states are determinants. Choose ZI so
that the ground state has L 2 1 , in which case (b) obviously holds. Next, consider
H = K + A Z(xt- x, I-’ + V. Angular momentum is still conserved and for
sufficiently small A the ground state will have the same L and, by continuity of
the ground states, (b) will continue to hold for small A. We are interested in
A = 1 but, under the scaling

x +X / A , v (x) -, ) v~x),
A - ~ u ( x / A=

the u, A problem is converted into the v ’ ,A = 1problem. Thus, u’ has the desired
properties. I thank B. Simon for this remark.]
COULOMB DENSITY FUNCTIONALS 255

5
Clearly E ( p i )= constant = D = E ( u )- upi. We claim FG)>D,thereby prov-
ing lack of convexity. Obviously, E @ )2 D,for otherwise we could use p instead
5
of pi in (3.15). (Note: up = upi =constant = C.) Suppose PG)=D.Then p
comes from some q5 that must be a ground state for u. But (a) and (b) show this
to be impossible. Thus $ ( j ) > D . Moreover, p cannot come from any ground
state 4 for any other u’. If it did, then

This implies that for some 1 4 i S M , E(pi)+I


u’pi < E ( v ‘ ) ,which is a contra-
diction. =
Remarks. (i) The foregoing proof holds just as well if (4,Ho4) is replaced
by T ( 4 )in the definition of I? [This functional will be denoted by f ( p ) and the
analog of (3.17) by T ( p ) . ]In other words, the interelectron Coulomb repulsion
plays no role in Theorem 3.4 (see Sec. 4C).
(ii) There are other p’s that do not come from some u, namely, those p E J”
that vanish on a nonempty open set. If u E L312+ L“ and 4 is its ground state,
then 1,4 cannot vanish in an open set by the unique continuation theorem [18].
(Strictly speaking, this theorem is only known to hold for u E L?,,, but it is believed to
hold for L312+ L“.) Presumably, suchp’s can, in many cases, be obtained as limits in
which u + on the open set. Therefore, if the set of allowed u’s can be extended
properly to include infinite u ’s,the existence of suchp’s may not have any particular
importance. The question is very delicate, however, as Englisch and Englisch [7]
showed recently. Even for one particle there are densities which never vanish but
which do not come from any u , even if one allows density matrices (see Sect. 4B)
instead of pure states. These densities have regions in which they are “small” so that
the obvious u (defined by u = p-*12Ap ‘ I 2 )has the property that -A + u cannot be
defined as a semibounded operator.

Theorem 3.5.

Remark, The right sides of (3.20) and (3.21) are automatically concave
functionals, which is a property we already proved for E.
Proof. Let M - ( u ) [respectively, M + ( u ) ] be the infimum in (3.20)
[respectively, (3.21)]. Obviously, M - ( u ) ~ M ’ j u ) .First, pick uo. Clearly
F ( p )s E ( u o )- J uop = F l ( p ) .Therefore,

p u l p E ~ 3 n ~ 1 ] .
256 LIE9

and hence M-(vo)?E(uo). Second, by (3.19), F ( p ) l E ( p ) , so that M ' ( v ) s


inf { F ( p ) + j p v I p € Y N ) = E ( v ) .
Let us pause briefly to review the situation. Three functionals have, been
defined:
FHK definedon ~ N C Y N ;
P definedon Y N c L 3 n L 1 ;
F defined on X = L3nL1.

Of these, only F is convex and only P and F satisfy the variational principle
for all v.
The next step is to find out something of the nature of F. It is at this point
that the analysis becomes complicated and where difficulties and incompleteness
arise. The basic reason is that the connection between v and p is anything but
simple. We have X = L3 flL' and its dual X * = L3I2+ L". Although X is not
the dual of X " , it is a subset of X**, the dual of X * .
Definitions. (i) A sequence pn E X is said to converge to p E X ( p , + p ) if and
only if llpn -pI13 -+ 0 and lip, -plll + 0. This is also called norm convergence. pn
converges weakly to p(p, - p ) if and only if 4 ZI (p,, - p ) -+ 0 for all v E Y = X * .
Clearly, strong convergence implies weak convergence.
(ii) A functional f on X is continuous (or norm continuous) if and only if
,on+ p implies f(Pn) + f ( p ) . Weak continuity requires the concept of nets to define
but, iff is weakly continuous, then whenever pn - p , f ( p n )+ f ( p ) . Weak continuity
implies norm continuity.
(iii) A real functional f on X is lower semicontinuous (1.s.c.) if and only if
p n + p implies f ( p ) 5 lim inf f ( p n ) . Weak lower semicontinuity requires nets to
define, but if f is weakly 1.s.c. then p n - p implies f(p)sliminff(p,). (Weak)
lower semicontinuity is equivalent to the following: Ifb)IA ) is (weakly) closed
for all real A.
Remarks. (i) Weak lower semicontinuity always implies lower semicon-
tinuity, but not conversely. It is a theorem of Mazur [19], however, that i f f is
convex and norm l.s.c., then it is automatically weakly 1.s.c.
(ii) The function p ( x ) = 0 is not in the L3fL1
l weak closure of $N.
The reader may be puzzled by all these definitions, especially lower semi-
continuity, because finite convex functions on R" are always continuous.
Unfortunately, this is not true in infinite-dimensional spaces such as the space
X we are considering. Even 1.s.c. cannot be taken for granted.
Theorem 3.6. F ( p ) is weakly (and hence also norm) lower semicontiwous.
Proof.

K A= b l F ( p ) i A ) = vp I A
I
for all v E Y .
COULOMB DENSITY FUNCTIONALS 251

Now if p, +p in norm and p, E K Athen, for each v E Y.

E(v)-
I (
up=lim E ( u ) -
I >up, < A .

Therefore K Ais norm closed, so that F is norm 1.s.c. Weak 1.s.c. is a consequence
of Mazur’s theorem.
Next we define the convex envelope (cE).
Definition. Let f be a real functional defined on a subset A of X. f ( p ) is
allowed to be +a, but not for all p E A. CE f is defined on all of X as follows:
CE f ( p ) =sup { g (p )l g is weakly I.s.c., g is convex on X, and g ( p ’ ) s f ( p ’ ) for all
p ’ A}.~
It is easy to check that CE f is convex and weakly 1.s.c. and CE f ( p ) S f ( p )
for a11 p E A. However, CE f(p) may be -too for some p.
The function of interest is CE @ with A = 4;N. Note that A is convex and that
E (and hence CE E ) is finite on A by Theorem 1.2. Since CE F < E on A, it is
obvious from (3.19) and Theorem 3.6 that F ~ C EE on X. On the other hand,
suppose we use CE @ instead of @ in (3.15). This gives a new function, which
we call E’. Clearly E ’ s E . Then, if E’ is used in (3.17), we get a new function
F’, and F’IF. However, an infinite-dimensional generalization of Fenchel’s
theorem [29] (which uses the Hahn-Banach theorem) states that if the original
function (in our case, CE @) is convex and weakly 1.s.c. on X , then its double
Legendre transform (in our case F’) is equal to the original function. Thus,
F = CE 3 and we have
Theorem 3.7. F ( p ) = CE E ( p ) for all p E L3nL’
The reader may wonder what Theorem 3.7 is good for; the following is an
example of the usefulness of the foregoing functional analysis (see Theorem 4.3).
Theorem 3.8. For all p E L3fl L1let

= +a otherwise.

Then F ( p ) 2 G(p), for all p E L3 fl L1.


Proof. G is obviously convex on X [see the remark after (1.10)]. We claim
that G is norm I.s.c. (Note: The norm in question is L3flL1, not the H’ norm
on ~ l ’ ~ If. ) so, we are done because G is then weakly 1.s.c. and, by Theorem
1.1, G s @ o n 4 ; N :b u t t h e n G s c E @ = F .
To prove norm l.s.c., let p, be any sequence in X with p, + p ; that is,
1/p, -plil + 0 and I(p, 0. We can assume that G = lim G ( p , ) exists and is
finite, and we have to show that G 2 G(p). We can also assume p ( x ) 2 0 a.e.
because if p < 0 on a set S of positive measure, then, for sufficiently large n,
258 LIEB

pn < 0 on some set of positive measure; hence p, i? and G(pn)= a.For a similar
reason we can assume 5 p = N . Since G(p,) < 00, pn E $N. Thus, if we define
g, =p:/' and g = p ' / ' , we have: (a) g, is bounded in H ' ; (b) g', + g 2 in L 3 and
L'. By the Banach-Alaoglu theorem there is an f E H ' such that gn-f and
Vgn-Vf weakly in L2. Clearly f ( x )2 0 . It is not hard to prove that if gn-f in
L2 and g z +q2
in L 1 ,then g = f . Hence V g = V f , and thus Vg,-Vg. But since
5 (Vg)2 is H -norm continuous, it is H' weakly I.s.c., so that lim G(p,) I)
W).
Theorem 3.8 is certainly not obvious. Among other things it says that if
p@ (and such p's can be quite smooth and innocent looking), then there exists
a sequence of potentials ZI, E L3l2+ L" such that E ( z I ,-l
) v,,p + 00. The reader
is asked to reflect on this fact. Another interesting fact is that F is convex and
finite on .9N,but infinite off 9N. However, the complement of 9N (in X ) is
dense (in the X norm) in 9N and 9N is dense in the cone of nonnegative
functions in X .
The following upper bound complements Theorem 3.8.
Theorem 3.9. If p E 9 N ,then

JI
F ( p ) ~ E ( p ()4s~ r ) ~ N ~ G ( p ) + p; ( x ) p ( y ) l x- y I - l dx dy. (3.22)

Proof. Use the definition (3.14). By Theorem 1.2 there is a determinantal


I,$, with I,$++p, such that (1.12) holds. With this I,$ we can calculate the Coulomb
repulsion I = ( I , $ , X ~ x i - x j ~ - ' I , $I) . has a direct term, given in (3.22), plus an
exchange term. The latter is negative, as is well known, since Ix - y I-' is a positive
definite kernel. Thus g(p)(right side of (3.22). Then use (3.19).
Remark. By one of Sobolev's inequalities,

D= Jj p ( x ) p ( y > l x - y l - ' d x d y ~(const.)llp1126/5.

Again, by Sobolev's inequality, (1.lo),


By Holder's inequality l\p\\&5 5 //p\\?\\p113.
llp1I35 (const.)G(p). Thus,

D 5 ( c o n ~ t . ) N ~ ' ~ G ( p5) (const.)"


''~ +N2G(p)].
To continue the study of F the following concept is needed.
Definition. Let f be a real functional on a subset A of a Banach space B,
and let p o € A .A linear functional I on B is said to be a tangent functional (TF)
at po if and only if for all p E A
f (P 1> f(Po) - - Po). (3.23)
1 may not be unique. If 1 is continuous, then 1 is a continuous tangent functional
at PO.
COULOMB DENSITY FUNCTIONALS 259

1 is a continuous linear functional on X if and only if it has the form j u p


with u EX*= Y. If f is convex, then at every point po at which f is finite, f has
at least one TF. This is guaranteed by the Hahn-Banach theorem. However, f
may have no continuous TF at po.
The functional of interest is obviously F. In general, F s g , but the following
says something about those p for which F(p)= E ( p ) .
Theorem 3.10. Let po E 4N.The following are equivalent:
(1) F(po)=g(po) and F has a continuous TF at po.
(2) p_OE d N *
(3) F has a continuous TF at P O ; that is, 8 ( p ) rE(po)-l v ( p - P O ) for p E 2%.
(4) (3) and ( 5 ) hold with the same u.
( 5 ) ~ ( u= )E ( p o ) + J upo for some u.
(6) ( 5 ) holds and, in addition, u E 'TNand u - P O .
( 7 ) E has a continuous TF at po and u is unique up to a constant. Moreouer, F
has the same continuous TFS at po, and no others.
Proof. (1)+(3): For p E ~ N ,

P(p)a%)2 F ( p o ) - I (p - P o ) = &o) - I
( 3 ) 3 ( 4 ) :Let F l ( p ) = f i ( ~ o ) - J u- (ppo ) s E ( p ) . Then

+
(4)3 ( 5 ) , (7) (3), (6)j (5): All trivial.
(5)w): R p o ) + J pov = E ( u )s ~ ( +pJ p~ O u)w ( p o ) =
$;(Po). Then,for a l l p E X , F ( p ) + l p u r E ( u ) = F ( p o ) + J p o u .
(2)+(5): By (3.16).
(5)+(2), (6): By Theorem 3.3, E ( p o ) = ( & H o Jfor / ) some CC, with J / - p o .
Then E ( u ) = ( * , H o ~ ) + J u p o j p oU~E d' T~N ,, and u - p O . Thus (1)-(6) are
+
equivalent and (7) (3). Now we show that (1)-(6) 3 (7). If u is a continuous TF
for F, then u is a continuous TF for E [by the proof of ( l ) j ( 3 ) ] . If u is a
continuous TF for p, then F ( p ) z E ( u ) - j u p , so u is a continuous TF for F.
Suppose kT has two continuous TFS u and w with u - w # constant. Then E ( u )=
P ( p o ) + j up0 and E ( w )= E(po)+I wpo. Since po E dN,this is impossible by
Theorem 3.2. H
It should be noted that the only place that the HK Theorem 3.2 entered in
the analysis of F was in establishing the uniqueness (modulo constants) in (7).
Now we turn to two important questions whose answers we cannot give but
that are obviously importanf for the theory. We replaced FHK by F because F H K
was not defined on all of .aN.Theorem 3.10 states that on dN, where FHKis
defined, F = fi = FHK and F has an essentially unique continuous TF.
Question 5. For which points of 4Ndoes F have a continuous TF? Where
there is one, is it unique (modulo adding a constant to u ) ?
260 LIEB

Question 6. If F has a continuous TF at po E Ca, given by some v E L3I2+ L",


is this v E VN?
Questions 5 and 6 have alternative formulations, given below.
Theorem 3.11. Let po E 9~and v E L312+ L". v is not necessarily in VN.Then,
for all p,

F ( ~ ) ? F ( ~ ~v)( p- -[p o ) (continuous TF) (3.24)

if and only if

E ( v )= F(po)+ v [ po [minimum in (3.21)]. (3.25)

Proof. Assume (3.24) and let M , ( p ) be its right side. Then

For the converse,

Question 5 is equivalent to the following: For which po E $N is there a z' such


that (3.25) holds? Is this v unique (up to constants)? Question 6 is the following:
If (3.25) holds, is u E YN?
Some insight into the continuous TFS of F are provided by the Bishop-Phelps
theorem. We refer the reader to Ref. 20 for this as well as other interesting
facts about convexity. A definition is needed.
Definition. Let F be a real functional on a real Banach space B with dual
B * (the set of continuous linear functionals on B ). b * E B * is said to be F-bounded
if there is a constant C (depending on b* but not on b ) such that F ( b )2 b*(b)+ C
for all b E B.
In our case B = X and F is our density functional.
Theorem 3.12. Every v E X *= L312+L" is F bounded.
Proof. By Theorem 3.8, F ( p ) = m if p & j N , so we only have to consider
p E9 N and prove that G ( p )2 up + C for some C. The proof of this is identical
to the last part of the proof of Theorem 3.1.
The Bishop-Phelps theorem is the following.
Theorem 3.13. Let F be a 1.s.c. convex functional on a real Banach space B.
(Note: Norm and weak I.S.C. are identical.) F can take the value +a,but not
everywhere. Then
(i) The continuous tangent functionals to F (over all of B ) are B *-norm dense
in the set of F-bounded functionals in B*
COULOMB DENSITY FUNCTIONALS 261

(ii) Suppose boE B and'b; E B* with F(bo)<co. For every E > O there exists
b, E B and b: E B* such that lib: -bglle*sE and
EllbE-b& sF(bo)+b;(bo)-inf {F(x)+b;(x)lx EB}.
Moreover, bz is tangent to F a t b,, namely F ( b ) 2 F(b,) - 6 %(b -be) for all b.
The significance of Theorem 3.13(i) is the following. There are certainly
many 0's in Y that are not in VN.(Example: Suppose 21 € L 3 l 2and llv113/2<L,
where L is the constant in (1.10). Then (t,b,H,t,b)>O for all t,b, but E ( v ) = O
because we can always take a sequence 4, that "leaks away to infinity.") Let
VE' VN,whence P(p)+ jup does not have a minimum. What Theorem 3.13 says
is that there always exists a sequence v, (not necessarily in V N such
) that
(a) F(p) +I pv, has a minimum at some pn E 9 N and this minimum is E(u,).
(b) F ( p ) r F ( p , ) - ~ u , ( p - p , ) forallp.
(c) u, + u in the L3'2+Lw norm.
Point (c) means the following: 0, = ZI + g, + h, with Ilg,113/2 + 0 and llhnIlrn+ 0.
In particular, if z1 E L3I2 with llv113/2<L,then IIv +g,(l<L for large n. Hence
o + g, & 7".If ZI, E VN,then it can only be because of the (vanishingly small) L"
piece h,.
One consequence of Theorem 3.13(ii) is the following.
Theorem 3.14. Let po E 9N.
Then there exists a sequence p, E 9~such that
(i) pn + p o in L~nL1norm.
(ii) F has a continuous TF a t p,.
Proof. Given n >0, by (3.17) there exists v , such that E ( u , ) - ~ p o v , >
F(po) - l/n. Hence
~ ( p2)~ ( v , - j'
) pv, 2~ I) (p - P o ) - 1/n.
( p-~ u,

Take E = 1 in Theorem 3.13. There exists w, E Y such that w, is a continuous


TFat some pn and
Ib" -poll'F(po)+I v"P0-Z
with

By the above,
Z = inf
{ I
F(p)+ v,plp E X
I.
z >~(p,)+[ pov, -npl.

4. Additional Remarks about Density Functionals


A. The N-Dependence of F
As was stressed earlier, any functional F that satisfies (3.20) or (3.21) must
depend explicitly on the particle number N. This fact is unavoidable and
262 LIEB

frequently overlooked. Let us denote the N dependence by F ( N , p ) . It might


be hoped that F is jointly convex in N and p in the sense that for N 2 2
F ( N + 1, p 1) + F ( N - 1, p 2 ) 2 2F(N,&J 1 + ipp2). (4.1)
This convexity definitely does not hold as a general feature, as will be demon-
strated.
The importance of convexity is shown by the following.
Theorem 4.1. Consider the following two statements about any two functionals,
Fund E : (i) F(N,p ) is jointly convex in N a n d p in the sense of (4.1).(ii) E ( N , u )
is convex in N for all fixed v ; that is, for N 2 2
E ( N + 1, v ) + E ( N- 1, v ) 2 2 E ( N ,u). (4.2)
(a) If (3.17) holds, then (ii) implies (i).
(b) If either (3.20)or (3.21) holds, then (i) implies (ii).
Proof. (a) For each v , E ( N , v ) - j u p is jointly ( N , p ) convex. By (3.17),
F ( N ,p ) is the supremum of such convex functions and hence is convex.
(b) Pick E > 0. For N + 1 there is a p+ such that

Likewise,
A = F ( N + 1, p + ) +
I p+v I E ( N + 1, v ) + ~ .

B = F ( N - 1,p - ) +
I p-v s E ( N - 1, a ) + & .

For the N problem, define 2p = p+ + p - . Then

I I1
2 F ( N , p ) + pv S A + B .

Since this holds for all E > 0, (ii) is proved. W


Equation (4.2) has a simple physical meaning. The ionization potential
increases as the number of electrons is decreased. This is intuitively expected
to be true, but if it is true, it must be because of some special property of the
Coulomb repulsion. A non-Coulombic counterexample is given below.
The kinetic energy functional F(N,p ) is not even convex in p (Theorem 3.4),
but the Legendre transform T ( N ,p ) is jointly convex. This is so because E ( N , v )
is indeed convex in N for independent particles as the explicit expression for
E,as the sum of the first N eigenvalues (counted with an extra multiplicity q )
shows.
What about the convexity of F when the Coulomb repulsion is included?
While it has been conjectured that E ( N , v ) is convex in N (for all v ) in the case
of Coulomb repulsion, this has never been proved. It has not even been proved
that E(3,v ) + E ( l , v ) z 2 E ( 2 , u ) .
COULOMB DENSITY FUNCTIONALS 263

Lest the reader think that convexity in N is a general feature, we present a


counterexample. Replace Ix 1-l by the hard-core repulsion B ( x ) = 00 if [xI < 1 and
6(x) = 0 otherwise. Pickfourdistinct pointsxo, y l , y 2 , y 3 inR3such that Iyi - y,I > 1
for all i # j but Ixo-yil < 1 for all i. Let u ( x ) I-2h < 0 in small balls about the
yi, u ( x ) = -3A in a small ball about xo, and u ( x ) = 0, otherwise. If the kinetic
energy be neglected, then E(1,u ) = -3A, E(2, u ) = -4h, and E ( 3 , u ) = -6h.
Convexity does not hold. This can be turned into a proper example by letting
A be sufficiently large so that the kinetic energy can effectively be neglected; it
is also possible to replace the hard core by a soft core.
Remark. The foregoing example is not applicable if 0 is replaced by ] X I- ' ,
thereby keeping alive the hope that convexity holds in the Coulomb case. The
reason is the following: Given any four points xo, y l , y2, y3, let
lxo-y1l =maxi { I x o - ~ ~ I l .
Then
Ixo-yil-' 5 Iyi -Yzl-' i- Iyi - y3l-l.
The proof of this is left as an exercise, as well as the implication that if the
kinetic energy is neglected, then convexity holds in the Coulomb case.
Question 7. For the case of Coulomb repulsion, is F(N, p ) jointly convex in
N and p ?

B . Density Matrices
Another possible modification of the theory of Sect. 3 is to replace densities
p (x) by single-particle admissible density matrices y ( x , x'). (See Questions 3 and
4 in Sec. 2. We do not restrict ourselves to y's that come from pure states I))($.)
This set of y ' s is convex, and E ( y ) ,defined analogously to (3.14), is convex [see
the proof of Theorem 4,1(b)].
Despite the attractive feature just mentioned, there are three drawbacks to
the approach:
(i) The problems about continuous tangent functionals remain and may even
be more complex than before.
(ii) The original aim of the theory was to express the energy in terms of p ( x )
and not y ( x , x').
(iii) While the set of admissible y ' s is well defined, it is not easy to identify.
Given some y, it is easy to verify that Tr y = N, but it is difficult to verify that
05ysqL
Still another possible modification is to retain p but to consider all N-particle
density matrices r instead of merely pure states $)(I,+. In other words, consider
r - p instead of 4 - p and define
FD&) = inf {Tr H J l r - p } (4.3)
on Y N and FDM(p)= +a otherwise. Because T - p is linear, F D M is convex on
$N. (Note: The example in Theorem 3.4 does not yield nonconvexity of F D ~ . )
264 LIEB

Obviously, the analog of (3.15)holds, namely,

E(fl)=inf{FDM(p)+jp v ) p E $ N ] . (4.4)

Since F Dis ~
convex, (4.4) can be used directly instead of (3.20) or (3.21).
Both F and FDM are convex. The amusing fact is that
F b )= F D M b ) r p E3N. (4.5)
Equation (4.5)is not at all obvious, but it does say that the modification does
not change the theory in any way. Equation (4.5) also yields another
characterization of F. Equation (4.5)is proved in Theorem 4.3.
First, I‘ is admissible if and only if
m

r(z, z‘)= c A ~ M Z ) M Z ’ ) *
1=1

with 0 IA,, ZA, = 1, and the (I/, are orthonormal. If ~ p ,then


,
Tr H 0 r = Z A l H d ,1.
Thus we conclude that for all p E .9N

FDMO))
= inf { C Af(pi)(
m

i=l
hipi = p, pi E 3N,
Ai I0,C hi = 1
I . (4.6)

A simpler expression (which has to be proved) is

= inf { c A f O ) t ) l c AQ, = P, P I E YN?AI 2 0,


I
C AI = 1 9 (4.7)

where the sums in (4.7)are restricted to finite sums. In view of (4.5),(4.7)is an


alternative characterization of F ( p ) for p E 9 ~ .
Theorem 4.2. Equation (4.7)is true.
Proof. Pick E >O. Using (4.6),let {Az, p l } be an infinite sequence satisfying
ZAQ, = p,pI E ,aN,andFDMLp)2 Z A & ( ~ ,-) E . Since x A 1 = 1 andZAf;(p,) < 03, there
exists K such that A = X I = K A r IE and B =CEKAF(pI) 5 E . Assume A > 0 for
otherwise we are done. By Theorem 1.1 and the convexity of G ( p ) = J (Vp1’2)2,

with p K = xz Aipi/A E 9 ~By.Theorem 3.9 and the remark following it,


F(pK)sC[N2G(pK)+N].
Therefore the finite sequence{&, p i } f ~with{AK,
~ p ~ =}{A, pK}satisfiesZ A f ( p i ) I
F D M b ) +&CN(N+ 1)f &.

Theorem 4.3. Equation (4.5)is true.


COULOMB D E N S I T Y F U N C T I O N A L S 265

Proof. The easy part is that for p E ~ N ,FDM(P)r F ( p ) . By (4.4). E ( v )5


FDM(p)+ J p v for all u. Hence, by(3.17),FDM(p)rF(p).Thehardpartiscontained
in Corollary 4.5, which will be assumed for now. Then: (i) F D M ( P ) s E ( p )by
(4.6); (ii) FDM is convex and I.s.c. Hence FDM(p)5 CE F ( p ) = F ( p ) by Theorem
3.7.
Theorem 4.4. Suppose bn}and p E$N and p n A p weakly in L'. Then there
exists a density matrix r, with T - p , such that Tr H0r 5 lim inf FDM(Pn).
The proof of Theorem 4.4, due to Barry Simon, is given in the appendix.
Corollary 4.5. (i) FDM is (norm and weakly) 1.s.c.
(ii) If p E 9N,there exists a density matrix r with r e p such that Tr HOT =
F D M ( ~(see
) Theorem 3.3).
Proof. (i) If pn + p , FDM(p)s T r H o r 5 1 i minf F D M ( p n )Norm
. 1.s.c. implies
weak I.s.c.
(ii) Take pn = p in Theorem 4.4.

C. The Kinetic Energy Functional


Kohn and Sham (KS) [30] define a kinetic energy functional TKs(p).There
are several other possible kinetic energy functionals and we shall explore their
interrelations, as well as the fact that T K S does not have a property assumed by
KS. KS define the exchange and correlation functional Ex&) by

~ , K ( p ) - t l J p(.\:)p(Y)Ix-YI-ldxdy+TKs(p)+E,c(p). (4.8)

FHK and T K S are defined on different subsets of 9 ~so,E x , is defined only on a


third unknown subset of .YN.This difficulty can be remedied by using E and
in (4.8), but there is another point that should be stressed: There is no reason
to believe that Ex, is convex on 9 ~ .
First, let us give some definitions. These use K instead of Ho but otherwise
are self-explanatory (with the aid of the equation numbers on the left):
(3.5): E'(v) +L ~ ;
on L~~~
(3.10): TKs(p) on d L (3.11);
(4.9)
(3.14): f(p) on 9N;

(3.17): T(p) on L ~ ~ L ' .


( T ( $ )= (9,K 9 ) was defined in (1.3) but it is quite different from T ( p )above.
It is hoped that this notational lapse will not be confusing.) All the previous
theorems [except for 3.9, wherein the last term in (3.22) should be omitted]
carry over to these quantities. The primes on E ' ( v )and d b indicate that these
are different from before. Since Theorem 3.4 still holds, d Q
rN
I is not 9i~.It is left
as an exercise to show that d N # dh.
Question 8. What is dNfl d h ?
266 LIEB

There is one more kinetic energy functional that can be defined on 9 ~ ,


namely,
Tdet(p)= inf {($, K$)l$ ~ p 4 E, WN,4 is a determinant}. (4.10)
Clearly, Tdet(P) 2 f ( p ) . The question to be addressed is whether T d e t = F. The
answer is No!, not even on all of dh. KS assumed implicitly that T K s ( p ) = Tdet(p)
for p E d h ; any such p minimizes K + V, but it is not true that such a p (x ) can
N
always be written as C,=l I$l(x)12 with the 4, being orthonormal functions on R3.
(Spin is a complication that is ignored at this point for simplicity.) In other words,
not every ground state of K + V is a determinant when degeneracy is present.
I thank B. Simon for drawing my attention to this subtlety and for the construction
in Theorem 4.8, which is reminiscent of the construction in Theorem 3.4.
Of course, T K S = ? on J& by definition. Also = T on d h by Theorem
3.10. The following shows that there are cases in which ? = T d e t .
Theorem 4.6. Suppose p E d h , so that K 4 V has a ground state. If this ground
state is nondegenerate, then T d e t ( p )= ? ( p ) .
Proof. The CC, that minimizes (4,[ K + V]*)is, of course, a determinant. H
The following analog of Theorem 3.3 will be needed for Theorem 4.8.
Theorem 4.1. Let p E $ ~ . Then there exists a determinant that minimizes
(4,KG) under the condition that t+h - p, 4 E WN, and 4 is a determinant. Thus,
(4.10) is actually a minimum.
Proof. Let 0,be a sequence of determinants with
0,- p and lim (D,,KD,) = Tdet(p).
The proof of Theorem 3.3 shows that 4 exists such that (i) 4 - p ; (ii) (4,K $ ) =
T d e t ( p ) ; (iii) D, - 4 strongly in L*. It suffices to show that 4 is a determinant.
Let f : , I = 1, . . . ,N , be the orthonormal single-particle functions of 0,.By the
Banach-Alaoglu theorem, N functions f ' , . . . ,f" exist so that (after passing to
a subsequence) f ; - f ' weakly. The f ' are not necessarily orthonormal. The
function
pJ(z19. .. 7 z,)=nf;(Zl)

then converges weakly to P = IIf ' . This so because any 4 E L2(lR3N)can be


approximated in norm by sums of product functions. Therefore,

-
But 0, 4, so D
0,-(N!)"* det [ f ' ( z , ) ] = D

= 4.

Theorem 4.8. Let N


H
=7
weakly.

and q = 1. Then there is a p E d h such that Td&) >


&4.
Proof. Take v (x)= Ix I-1, the hydrogen potential. The eigenvalues of -h + v
are -1/4 (onefold), -1/16 (fourfold), -1/36 (ninefold). All other eigenvalues
are greater than -1/36. The ground state for N = 7 and q = 1 is = 36-fold (z)
COULOMB DENSITY FUNCTIONALS 267

degenerate, and a basis for this eigenspace consists of the determinants


(7!)-’” det (lS, 2S, 2P1, 2P2, 2P3,f , g ) where f and g are any orthonormal func-
tions in the nine-dimensional space M spanned by S, PI,Pz,P3,D1, . . . ,D s (an
orthonormal set for the 3S, 3P, and 3 0 waves). Let d ( f ,g ) denote the above
normalized determinant and let
Ds).
31’z$ = d ( S , D 1 ) + d ( D , D3)+d(D4,
,
Then$-p Withp=p,+pb and
3
pa(x)=lls(x)12+12s(x)12+ c 12pi(x)12,
r=l

S
3Pb(X)=IS(X)12+C IDi(x)12.
i=l

Clearly p E d h since t,b is a ground state.


If Tdet(p)
= F ( p ) , then there exists a determinant q5 with q5 ~p and such that
q5 must be a ground state. Therefore q5 = d ( f ,g ) for some orthonormal f , g E M .
Thus,
I f(x )I2 + lg (x 1l2 = Pb (x 1. (4.11)
I claim that this is impossible. Write f = A + D and g = B + d , with A and B
being linear combinations of S and the Pi while D and d are linear combinations
of the D,. Now the S, P, and D waves behave as lxlo, lxll, lxI2, respectively, near
the origin. By examining the behavior of (4.1 1) near the origin we conclude that
5
=3C
ID(X)I~+I~(X ) I ~IDi(x)12.
1

Since all the D, waves have the same radial wave functions, this is really an
equality about spherical harmonics Y2,,,.The right side of the last equality is
spherically symmetric, so the problem is to find two linear combinations F and
G of the Yzmsuch that
IF(R)~’+ IG(R)~~ = constant > 0.

This is impossible, and the proof is left as an exercise. (It is easily carried out
if the following five basis functions are used: xyr-2, xzr-2, yzr-’, 3x2rP2-1,
3yZF2-1, with r 2 = x 2 + y z + z 2 . ) W

Remarks. (i) N = 7 is not special; it was chosen for convenience in the proof.
(ii) An alternative way of viewing Theorem 4.8 is following. Suppose K + V
has a degenerate ground state, so that the ground eigenspace G is more than
one-dimensional. 4 E G is a linear combination of determinants. Consider a
perturbation w of o, namely, o -+ v +Aw. In first-order perturbation theory,
V+A W picks out a subspace g of G as the new ground eigenspace. If g is one
dimensional, then g consists of one determinant since the ground eigenspace of
V + A W always contains determinants (see Theorem 4.6). Now we ask, if IL0 E G
268 LIEB

and $ o ~ p o can
, w be chosen so that g is one dimensional and g ={40}?
Alternatively, can w be chosen so that min fl wp11J-p and ((I E G} occurs
uniquely for p = po? If so, t,bo is a determinant. Theorem 4.8 says that there can
be a po such that no w can pick it out uniquely.
Even though Tdet(p)>?(p) for some p, Tdetstill satisfies the variational
principle for E ( v ) .
Theorem 4.9. For all v E L3I2+ L“

Proof. Equation (4.12) is equivalent to the following:


E’(u)= inf ((4,[ K + VM)M E W N }
= inf ((4,[K + v]I,b)I~,b
E WN, is a determinant}

=&).
Clearly E ’ ( v )s,??(v). Consider the operator -A+v(x). We define its “eigen-
values” e l 5 e2, . . . (here, spin degeneracy is included) by the min-max principle:
en+l=sup { E n ( 4 1 r . . . , &)I,
where
En(41, ,4n)=inf{(~,[-~+vI4)I4~~’,II4II2=1
* *

and 4 is orthogonal to . . . ,&}.


From this definition, it follows by a standard argument that
N
E N ( u ) =1 ei =inf
i=l
[ 1 (4,,[-A+v]4i)lq51,.
N

i=l I
. . , 4Nare orthonormal . (4.13)

But this least infimum equals

[ c hi(&, [ - A + ~ ] 4 ~ ) ( qb,.
00

inf 4 ~ , . . , are orthonormal,


i=l

m
OIAirland
i=l
1Ai=N. I
This is easy to verify.
Let II, E ‘WNand let y = ZAifi>(fibe its one-particle density matrix (including
spin and with the fi-orthonormal. 0 Ihi 5 1, ZAi = N ) . Then
($3 [ K + Vl4) = ZAi(fi, [-A + vlfi).
Thus E ’ ( v )? E N ( v) . But E N ( v )=g(v)by inspection.

Remark. This proof gives a formula for E ’ ( v ) ,namely, E N ( v ) .


C O U L O M B DENSITY FUNCTIONALS 269

The situation is complicated, so let us summarize it. TKS is defined only on


d k , the set of p’s that come from ground states for some 0. d h has a smaller
subset, d k , in which p comes from a determinantal ground state, d k includes.
but is larger than, d;,the set of p ’ s that come from nondegenerate ground
states. (Note: By Theorem 3.2 any p comes from a unique v (up to constants).
Thus, if p comes from a determinant in a degenerate ground eigenspace, then
pi! dK.)On d&we have
= Tdet(p)= fb).
TKS(P)
Elsewhere on dh,
Tdetb)>TKS(p)=fb).
Thus, there are two choices for (4.8): either TKsor Tdet.On BN,the complement
of d h , TKS is not defined (but f,Tdet,and T are defined). The preferred functional
here is T ( p ) because it is convex and hence most manageable. On BN,T ( p )is
probably strictly less than f ( p ) 5 Tdet(p)-at least this is so when T has a
continuous tangent functional (Theorem 3.10). Such points are dense (Theorem
3.14). In any case, since T, f,and Tdetcan be interchangeably used in (4.12);
it makes no difference which is used as far as E’(u)is concerned.

5. Some Density Functionals That Are Bounds


In this section we forego the abstract functional theory of the previous sections
and instead expound a different philosophy. Rather than pursuing “the correct
density functional,” which seems to be uncomputable, we shall content ourselves
here with finding upper and lower bounds to the various quantities of interest
in terms of p ( x ) . This latter program can provide rigorous bounds on ground
state energies that, while they may not always be extremely accurate, do have
a proper place in our conceptual scheme. Some of these bounds will be briefly
displayed here; the interested reader is referred to the original papers for proofs.
It should be remembered that if one has bounds on two quantities (e.g., T
and I ; see below) and even if these bounds are optimal, then, in general, the
sum of the bounds is not optimal for the sum of the two quantities (e.g., T + I ) .

A . Kinetic Energy Lower Bound


Lieb and Thirring (LT) [21] (also see Ref. 16) proved (for fermions in three
dimensions) that if CC, -p, then (for all N)

I
T ( ~ ) Z K ‘ ( ~ T ) - p~ ( /~~) ’~dx,
/ -~ ~ / ~ (5.1)

where K “ is the “classical” value (3/5)(6w2)”l’. LT conjectured that (5.1) holds


in three dimensions with the (47r)-*l3deleted. [Note: Although an analog of
(5.1) holds in all dimensions, the corresponding constant is definitely less than
K“ in one and two dimensions.] In Ref. 22 (also see Ref. 16) ( 4 ~ ) - ’ / ~was
replaced by 1 . 4 9 6 ( 4 ~ ) - * / ~ .
270 LlEB

Incidentally, the statement r ( $ ) ~ K q -( ~ p 5/l 3~for all 4, all N , and some K


is equivalent to the following [21]. Let ZI be any nonpositive potential in L5"(R3)
and let e l s e 2 s . . . , be the negative eigenvalues (if any) of - A + u ( x ) counting
degeneracy, but not counting the q- fold degeneracy. Then

Leiz - L 5 \ u ( x ) ( ~dx
/~

with K = (3/5)(2/5L)'I3.

B. -Finetic Energy Upper Bound


There is, of course, no upper bound for T ( 4 )in terms of p. March and Young
(MY) [17] proposed that for all p E -aN there is a determinantal 4, with 4 * p ,
such that

where n is the dimension and K " = m 2 / 3 for n = 1. (Compare (5.2)with Theorem


1.2.)They proved (5.2)for n = 1,but their proof for n > 1has an error. Equation
(5.2) for n > 1 is still an open problem. The MY construction for n = 1 motivated
the construction in the proof of Theorem 1.2.

C, Lower Bound for the Indirect Part of the Coulomb Repulsion


Let f' be a density matrix (which may be a pure state, r = $)($) with JI-p.
Let

(5.3)

be the Coulomb repulsive energy. The indirect part of this energy, E ( r ) , is


defined by

m = D ( P ) +Em, (5.4)
with

being the direct part.


In Ref. 23 it was shown that

E ( T )2 - C j p ( ~ ) ~ dx
/' (5.6)

with C = 8.5. In Ref. 24 this was improved to C = 1.68. The sharp (i.e., best)
C in (5.6) is not known, but it is larger than 1.23.
COULOMB DENSITY FUNCTIONALS 271

It is well known that in any pure, determinantal state, E ( T )< 0. For other
states, E ( r ) can be positive. Indeed, for any fixed p there is no upper bound
for E ( T )(see Ref. 24).
There is no q-dependence in (5.6) and, indeed, (5.6) holds for all statistics
(i.e., C does not depend on statistics). This is explained in Ref. 24. The Dirac
approximation has CDq-1’3in (5.6) with CD= 3 ( 6 / ~ ) ” ’ / 4= 0.93, but this q
dependence is an artifact of the particular q-dependent determinantal (c, used
to evaluate E from (5.4).
It should be noted that the bound

I(r)2 D ( p ) - c J’ p4/3 (5.7)

is nof coneex in p. It is not even positive. These two faults lead to absurd
conclusions when the right side of (5.7) is used in Thomas-Fermi-Dirac theory
(see Ref. 25).
Since r - p is linear,

l ( p ) = inf {I(r)lrHp] (5.8)


is convex in p. In other words, an optimal positive, convex lower bound must
exist. Any reader who is devoted to abstract density functional theory, in the
spirit of Sec. 3 or (5.8), should try to guess a plausible form for I@).(Proving
it is another matter.) It will quickly be seen that f ( p ) must be extremely
complicated, and to say that it is “nonlocal” is an understatement. To see this,
consider N = 2 and p consisting of two “bumps,” p 1 and p2, very far apart. As
long as s p 1 = j p 2 = 1, [ ( p ) - D ( p ) = O , independently of p 1 and p2. But when
J p l > l , j p 2 < 1 , then I ( p ) - D ( p ) depends heavily on p 1 but not on p2. The
reason is that in the former case the two electrons can be far apart in the two
bumps; in the latter case the two electrons must partly be close together in the
first bump.
A problem that is physically more relevant and that illustrates the hidden
complexity of density functional theory is the following problem about induced
dipolar (or Van der Waals) forces raised in Ref. 25. When two atoms are a
distance R apart, and R is large, there is an attraction -RP6(neglecting retarda-
tion effects). This attraction comes from the Coulomb repulsion, but it is not a
static effect. The atomic dipole moment is almost zero. (There are, in fact, tiny
dipole moments, but these are opposite in sign by symmetry, and hence repulsive.
They must exist by the Feynmann-Hellman theorem: d E / d R = electric potential
at the nucleus. I thank C. Herring for this remark.) There is almost no static
dipole moment because to create one would cost a polarization energy ad’. The
attractive energy is - d 2 R P 3 and, if R - 3 < a , d = 0 for minimum energy. The
cause of the -RP6 energy is more subtle, but it has a semiclassical basis: The
electrons in each atom move in phase while maintaining the spherical symmetry
about each atom. The energy cost is then a d 4 and the minimum energy occurs
when 2ad’ = RP3.Thus, the -R-6 attraction comes from the fact that the electron
272 LIEB

cloud cannot be thought of as a simple “fluid.” This effect is somehow built into
I@),but an explicit form of f@) that will produce this effect has yet to be
displayed.

D. A Variational Principle
E ( u ) , given by (3 .3 , satisfies (by definition) the well-known variational
principle

E ( v )5 (4,Hv4). (5.9)
Can an upper bound for E(u) be given in terms of p alone? If ( 5 . 2 ) were true,
then, for any p E .9~,

I
E(v )srig h tsid eo f ( 5 . 2 ) + D ( p ) + up. (5.10)

[See the remark about E ( r )for determinants in Sec. 5C.I


An upper bound for E ( u )can, indeed, be given in terms of the one-particle
density matrices y ( z , 2 ’ ) as follows [26]:
Let y be any admissible one-particle density matrix (0sy 51, Tr y = N ) .
(Note: y includes spin. It was called 7 in Sec. 2 ) . Then

E(u)STr y(-A+v(x))+$
I K&, z’)Ix - x ’ \ - ’ d z d z ’ , (5.11)

where j d z = C,j dx and

K*(z,2 ’ ) = Y ( Z , z ) y ( z ’ , z’)--Iy(z, z’)l2. (5.12)

The form (5.11) is well known if y came from a pure state $)(4with 4 being
a determinant. The point about (5.11) is that it holds for all admissible y.
Incidentally, the minimum of (5.11) over all admissible y occurs when y comes
from a determinantal4. In other words, the best Hartree-Fock function minimizes
( 5 . l l ) , but (5.11) is interesting precisely because this HF function is unknown.

E. Thomas-Fermi Theory
This theory (see Ref. 25 for an exposition) does not yield bounds and therefore
does not properly belong here. However, it illustrates the usefulness of the
bounds in Secs. 5A-C.
The TF functional is

I
gTF(p)=Kcq-2’3p 5 / 3 + D ( p ) + j up, (5.13)

while the TF Weizsaecker functional gTFW(p) is the right-hand side of (5.10). If


-C j p 4 / 3 is added to the right-hand side of (5.13), the result is TF Dirac theory.
COULOMB DENSITY FUNCTIONALS 273

The TF energy for N particles is defined by

[
ETF= inf gTF(p)I p = N] , (5.14)

and ETm.
and similarly for ETFW
Now suppose that v is an atomic or molecular potential, that is,
k
v(x) = - c
j=1
Zj(X (5.15)

with the z j > 0. It is a fact [14] that under the scaling z j + Azj and N +AN, as A + 00

E ~ ~ / E ( v1,
)+ (5.16)
where E ( v ) is the true ground state energy. Note that (5.16) also holds if ETF
is replaced by Em or ETFD (see Ref. 25).
Thus we see that if the conjecture in Sec. 5A holds, then, combining (5.1)
with (5.7), TFD theory is a lower bound that is asymptotically exact. Similarly,
if (5.2) holds, then, as remarked in Sec. 5C TFW theory is an upper bound that
is asymptotically exact.

F. Two-Body Density Matrices


If one is willing to go beyond the one-body density p or one-body density
matrix y and consider the two-body reduced density matrix Y ( ~ )then , E ( v ) is
directly and exactly expressible in terms of since H, has only one- and
two-body terms. The problem is that it is very difficult to decide when a given
y"' is, in fact, the reduction of an admissible N-body density matrix r. This is
called the N-representability problem and it has not been solved. (This is to be
compared with the fact that there is a simple necessary and sufficient condition
for a one-body y to be N-representable; see Sec. 2).
It is possible, however, to find some necessary conditions and some sufficient
conditions for to be N-representable. Using these, bounds on E ( v ) can be
derived. Since this approach is outside the scope of this article, we refer the
reader to the excellent review of Percus [27].

Appendix: Proofs of Theorems 1.3,3.3, and 4.4


The following proof of Theorem 1.3 is due to H. Brezis (pivate communi-
cation.)
Proof. For simplicity of presentation we take N = 2 and q =1 (no spin).
Therefore we have $(x, y ) and

Suppose that $,, +(I, in H 1 (R3x R3); that is, $, + $ and V+,, + V $ in L2. We
want to show that F,, +F in L 2 ( R 3 )and VF,,+ VF in L2(R3).The former is trivial:
214 LIEB

By the Schwarz inequality,

and the right-hand side converges to zero.


The proof that VF, + VF is the difficult one. It is sufficient to prove conver-
gence for some subsequence n,, j = 1 , 2 , . . . . (If VF,,+VF, then there is some
subsequence and some E > 0 such that IlVF,, - VFll> E . But then this subsequence
clearly does not have a subsequence which converges to VF.)Now since 4, + (I,
in H' there is some subsequence and some function G E H' such that
I(I,n,(X, Y ) I ~ GY () andIV(I,nl(x,
~, Y)I'G(~,Y),
a.e. in R6. (The proof of this fact is the same as the first half of the proof of the
Riesz-Fischer lemma that L' is complete.) Henceforth, we shall replace ni by
n. We shall also assume, for simplicity, that F, ( x ) and F ( x )> 0 for all x (otherwise,
an approximation argument can be used).
Now
ZVF, ( x ) =A,, (x)/F,( x ) + C.C.
with

A n ( x ) = [ ~ n ( x , Y ) d y and Bn(x,y)=cCIn(x,y)*vllrn(x,y).

As we saw, F, + F in L 2 , so, by passing to a subsequence, we can assume


F,(x)+ F ( x )a.e. Furthermore, a.e.

JBn(x,y)I'G(x, Y ) ~ E L ~ ( R ~ ) .

By passing to a subsequence we can assume V(I,, + V + and (I,, + (I, a.e. Thus, by
dominated convergence, B , + B in L'(R6). For this subsequence (B,(x,y ) -
B ( x , y ) J + Oa.e.
, (R6). Then, for a.e. x , IB,(x,y)-B(x,y)l+O a.e. y . Thus, by
dominated convergence dylB,(x, y ) - B ( x , y)I +O, a.e. x . In other words, for
some subsequence, VF,(x).+ VF(x),a.e.
Finally, we note that, by the Schwarz inequality,

IVFn(x)12'[ IV(I,n(x,y ) I 2 d y 5 [ G ( x , ~ ) ~ d y = C ( x ) ~ *

Since C is a fixed L 2 function, VF, +VF in L2 by dominated convergence. U


COULOMB DENSITY FUNCTIONALS 27.5

Proof of Theorem 3.3. Let (I,, (with ( I , , j ~ p be


) a minimizing sequence for
6@).The (I,, are obviously bounded in so, by the Banach-Alaoglu
~ ~ ) that (I,, -(I, weakly in H 1 ( R 3 N )Obviously,
theorem, there is a (I, E H ' ( [ W such .
(I, has the same symmetry as the (I,,. It is well known that under weak limits
positive quadratic forms decrease. Thus
&) = lim ((I,,, Ho(I,,)2 ((I,, Ho(I,).
If we can show that (I,-p, we are done. To do so it is sufficient to prove that
(I,, -+ (I, strongly because if (I, H; we have, by the easy part of Theorem 1.3 that
p '/' = p:" + $ ' I 2 in L2,so that 6 = p.
Strong convergence will be proved by showing that 5 [(I,['= 1. Let S be the
characteristic function of some bounded set in R3N. By the Rellich-Kondrachov
theorem [28] there is a subsequence (which can be chosen independent of S ) of
the (I,f such that S(I,, converges strongly (in L 2 )to S(I,.Pick E > 0 and let x be the
characteristic function of a bounded set in R3 such that

E 1.jdl-*)==[l(I,,12c[1-x(x1)l.
But
c 11-x (xr)l 2 1- s,
where S = IIx(x,).Thus, J]t,hf1s
' 2 1- E . Since 1(I,112S-+jJ(I,I2S,
we have that J 2
J1(I,12S21-s foralIE>O. H
Remark. The symmetry of (I, was not needed in this proof provided one
generalizes definition (1.6) to
N
p(X)=c I(I,(Zl,. . . , (X, . . . ,2
gr), ...
~ )d X)1 ~ &,+I . . . dXN. (A.1)
u r=l

The following proof of Theorem 4.4is due to B. Simon (private communica-


tion). It is closely related to the proof of Theorem 3.3 just given.
Proof. Without loss, replace Ho by h 2 = H o + 1 in the definitions. h - l is a
bounded operator. We can assume that g, = F D M ( p n ) < a,g = lim g, exists, and
Tr h T,h = Tr T,h2 5 g, + 1/ n with T, -pn. Thus, y n = h T,h is uniformly bounded
in the trace norm. The dual of the compact operators, com, is the trace class
operators t, and y E t takes A E com into T r y A . A sequence yn E t converges to
y E t, in the weak* topology, if and only if Tr y A + Tr y A for all A E corn.
The Banach-Alaoglu theorem states that a norm-closed ball of finite radius
in t is compact in the weak* topology. For us this means that there exists y with
T r y <a,so that, for a subsequence, Tr y,A + T r yA for every compact A .
Clearly, y 2 0 and therefore lim Tr y n 2 Tr y. Also, y obviously has the correct
(Pauli) symmetry. If we can show that T=h-' yh-' (which is in trace class)
satisfies r e p , we are done, To d o this we shall show that if T - p ' , then
5 (P, -p')f .+ 0 for any f E L". This would mean that p n - p ' weakly in L'. But
sincepk-p i n L ' , p ' = p .
276 LIEB

As in the proof of Theorem 3.3, for any E > O there is a ,y (=characteristic


function of a bounded set in R3) such that

Since p n -1p, p n ( I -x)< E for n sufficiently large. If


+,,(XI,. . . ,x N ) = wl,.
.., X N ;XI,. . . ,x N )
(after summing on spins), and similarly for 4, we have (as in Theorem 3.3)

where S = II,y(xi). In view of this, it is sufficient to show that

1,p-J +P with P = S C f(xi).


i

Let P = P ( x l , . . . ,x N ) be any bounded functions of compact support and let


Mpbe the operator (in L 2 )of multiplication by P. It is a fact that A P= h-’Mph-’
is compact. (This is essentially the same as the Rellich-Kondrachov theorem
used in Theorem 3.3.) Therefore
Tr TnMp= Tr ynAp+ Tr yAp = Tr rMp.

Acknowledgment
This work was partially supported by U.S. National Science Foundation grant
No. PHY-7825390-A02. This paper is a revised version of a paper with the
same title that appeared in Physics as Natural Philosophy: Essays in Honor of
Laszlo Tisza on his 75th Birthday, H. Feshbach and A. Shimony, Eds. (M.I.T.
Press, Cambridge, 1982), pp. 11 1-149.

Bibliography
[l] L. H. Thomas, Proc. Camb. Phil. SOC.23, 542 (1927).
[2] E. Fermi, Rend. Accad. Naz. Lincei 6, 602 (1927).
[ 3 ] P. Hohenberg and W. Kohn, Phys. Rev. €3 136, 864 (1964).
[4] M. M. Morell, R. G. Parr and M. Levy, J. Chem. Phys. 62, 549 (1975).
[5] R. G . Parr, S. Gadre and L. J. Bartolotti, Proc. Natl. Acad. Sci. USA 76, 2522 (1979).
[6] R. A. Donnelly and R. G . Parr, J. Chem. Phys. 69,4431 (1978).
[7] H. Englisch and R. Englisch, “Hohenberg-Kohn theorem and non-v-representable densities,”
Physica A, to be published.
[8] T. L. Gilbert, Phys. Rev. B 6,211 (1975).
191 J. E. Harriman, Phys. Rev. A 6, 680 (1981).
COULOMB DENSITY FUNCTIONALS 277

[lo] M. Levy, Proc. Natl. Acad. Sci. USA 76, 6062 (1979).
[ l l ] M. Levy, Phys. Rev. A 26, 1200 (1982).
[lZ] S. M. Valone, J. Chem. Phys. 73, 1344 (1980); ibid. 73, 4653 (1980).
[13] A. S. Bamzai and B. M. Deb, Rev. Mod. Phys. 53, 95 (1981). Erratum, 53, 593 (1981).
[14] E. H. Lieb and B. Simon, Adv. Math. 23, 22 (1977). See also Thomas-Fermi theory revisited,
Phys. Rev. Lett. 31,681 (1973). See also Refs. 16 and 25.
[15] M. Hoffmann-Ostenhof, and T. Hoffmann-Ostenhof, Phys. Rev. A 16,1782 (1977).
[16] E. H. Lieb, Rev. Mod. Phys. 48, 553 (1976).
[17] N. H. March and W. H. Young, Proc. Phys. SOC.72, 182 (1958).
[18] M. Reed and B. Simon, Methods of Modern Mathematical Physics (Academic, New York,
1978), Vol. 4.
[19] S. Mazur, Studia Math. 4, 70 (1933).
[20] R. B. Israel, Convexity in the Theory ofLattice Gases (Princeton U.P., Princeton NJ, 1979).
[21] E. H. Lieb and W. E. Thirring, “Inequalities for the moments of the eigenvalues of the
Schrodinger hamiltonian and their relation to Sobolev inequalities,” in Studies in Mathematical
Physics, E. H. Lieb, B. Simon, and A. S . Wightman, Eds. (Princeton U.P., Princeton, NJ, 1976).
See also Phys. Rev. Lett. 687 (1975); Errara, 35, 1116 (1975).
[22] E. H. Lieb, Am. Math. SOC.Proc. Symp. Pure Math. 36, 241 (1980).
[23] E. H. Lieb, Phys. Lett. A 70, 444 (1979).
[24] E. H. Lieb and S. Oxford, Int. J. Quantum Chem. 19,427 (1981).
[25] E. H. Lieb, Rev. Mod. Phys. 53, 603 (1981); Errata, 54, 311 (1982).
[26] E. H. Lieb, Phys. Rev. Lett. 46, 457 (1981); Erratum. 47, 69 (1981).
[27] J. K. Percus, Int. J. Quantum Chem. 13,89 (1978).
[28] R. A. Adams, Sobolev Spaces (Academic Press, New York, 1975).
[29] W. Fenchel, Can. J. Math. 1,23 (1949).
[30] W. Kohn and L. J. Sham, Phys. Rev. A 140 1133 (1965).

Received October 19, 1982


Accepted for publication March 11, 1983

Вам также может понравиться