Density Functionals For Couiomb Systems: Departments New Jersey 08544, U.S.A
Density Functionals For Couiomb Systems: Departments New Jersey 08544, U.S.A
Density Functionals For Couiomb Systems: Departments 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
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
T ( f )= ( 2 7 ~ - k21f(k)12
~ dk, (1.4)
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
(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 ( + )s ( ~ T ) ~ N [Vp1/2(x)]2dx (ferrnions). (1.12)
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
(1.13)
with
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
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
(3.1)
) 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)
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)
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
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
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
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
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
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.
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)
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
if and only if
(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,
By the above,
Z = inf
{ I
F(p)+ v,plp E X
I.
z >~(p,)+[ pov, -npl.
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 ) + & .
I I1
2 F ( N , p ) + pv S A + B .
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
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
FDMO))
= inf { C Af(pi)(
m
i=l
hipi = p, pi E 3N,
Ai I0,C hi = 1
I . (4.6)
~ , K ( p ) - t l J p(.\:)p(Y)Ix-YI-ldxdy+TKs(p)+E,c(p). (4.8)
-
But 0, 4, so D
0,-(N!)"* det [ f ' ( z , ) ] = D
= 4.
S
3Pb(X)=IS(X)12+C IDi(x)12.
i=l
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“
=&).
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
* *
i=l I
. . , 4Nare orthonormal . (4.13)
[ c hi(&, [ - A + ~ ] 4 ~ ) ( qb,.
00
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
($3 [ K + Vl4) = ZAi(fi, [-A + vlfi).
Thus E ’ ( v )? E N ( v) . But E N ( v )=g(v)by inspection.
I
T ( ~ ) Z K ‘ ( ~ T ) - p~ ( /~~) ’~dx,
/ -~ ~ / ~ (5.1)
Leiz - L 5 \ u ( x ) ( ~dx
/~
with K = (3/5)(2/5L)'I3.
(5.3)
m = D ( P ) +Em, (5.4)
with
E ( T )2 - C j p ( ~ ) ~ dx
/' (5.6)
with C = 8.5. In Ref. 24 this was improved to C = 1.68. The sharp (i.e., best)
C in (5.6) is not known, but it is larger than 1.23.
COULOMB DENSITY FUNCTIONALS 271
It is well known that in any pure, determinantal state, E ( T )< 0. For other
states, E ( r ) can be positive. Indeed, for any fixed p there is no upper bound
for E ( T )(see Ref. 24).
There is no q-dependence in (5.6) and, indeed, (5.6) holds for all statistics
(i.e., C does not depend on statistics). This is explained in Ref. 24. The Dirac
approximation has CDq-1’3in (5.6) with CD= 3 ( 6 / ~ ) ” ’ / 4= 0.93, but this q
dependence is an artifact of the particular q-dependent determinantal (c, used
to evaluate E from (5.4).
It should be noted that the bound
is nof coneex in p. It is not even positive. These two faults lead to absurd
conclusions when the right side of (5.7) is used in Thomas-Fermi-Dirac theory
(see Ref. 25).
Since r - p is linear,
cloud cannot be thought of as a simple “fluid.” This effect is somehow built into
I@),but an explicit form of f@) that will produce this effect has yet to be
displayed.
D. A Variational Principle
E ( u ) , given by (3 .3 , satisfies (by definition) the well-known variational
principle
E ( v )5 (4,Hv4). (5.9)
Can an upper bound for E(u) be given in terms of p alone? If ( 5 . 2 ) were true,
then, for any p E .9~,
I
E(v )srig h tsid eo f ( 5 . 2 ) + D ( p ) + up. (5.10)
E(u)STr y(-A+v(x))+$
I K&, z’)Ix - x ’ \ - ’ d z d z ’ , (5.11)
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)
[
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.
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
A n ( x ) = [ ~ n ( x , Y ) d y and Bn(x,y)=cCIn(x,y)*vllrn(x,y).
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 ) ~ *
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
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).