You are on page 1of 37

Communications in

Commun. Math. Phys. 118, 177-213 (1988) Mathematical


Physics
© Springer-Verlag1988

The Stability and Instability of Relativistic Matter


Elliott H. Lieb 1,, and Horng-Tzer Yau 2,**
1 Departments of Mathematics and Physics, Princeton University, P.O. Box 708,
Princeton, NJ 08544, USA
2 School of Mathematics, The Institute for Advanced Study, Princeton, NJ 08540, USA

Abstract. We consider the quantum mechanical many-body problem of


electrons and fixed nuclei interacting via Coulomb forces, but with a relativistic
form for the kinetic energy, namely p2/2m is replaced by (p2c2 -~ n~2c4) 1/2 --/~c 2.
The electrons are allowed to have q spin states (q = 2 in nature). For one
electron and one nucleus instability occurs if ze > 2/g, where z is the nuclear
charge and e is the fine structure constant. We prove that stability occurs in the
many-body case if z~ < 2/re and e < 1/(47@ F o r small z, a better bound on c~is
also given. In the other direction we show that there is a critical ec (no greater
than 128/15re) such that if e > ec then instability always occurs for all positive z
(not necessarily integral) when the number of nuclei is large enough. Several
other results of a technical nature are also given such as localization estimates
and bounds for the relativistic kinetic energy.

I. Introduction
One of the early important successes of quantum mechanics was the interpretation
of the stability of the hydrogen atom. The ground state energy of the hydrogen
Hamiltonian is finite and thus the hydrogen atom is stable quantum mechanically,
even though it is unstable classically. The Coulomb singularity -ze2/r is
controlled by a new feature of Schr6dinger mechanics, the uncertainty principle.
While the stability of the hydrogen atom is clear and simple, a more subtle question
arises when many particles are taken into account. It is convenient to distinguish
two notions of stability.
Stability of the first kind: The ground state energy is finite.
Stability of the second kind: The ground state energy is bounded below by a
constant times the number of particles.

* Work partially supported by U.S. National Sdence Foundation grant PHY-85-15288-A02


** The author thanks the Institute for Advanced Study for its hospitality and the U.S. National
Science Foundation for support under grant DMS-8601978
78 E.H. Lieb and H.-T. Yau

The second kind of stability, now commonly known as the stability of matter,
was proved in 1967 by Dyson and Lenard [10] - four decades after the invention of
Schr6dinger mechanics. The Dyson-Lenard analysis clearly showed that the
stability of matter depends crucially on the Pauli exclusion principle. The ground
state energy (call if Ey) of N fermions interacting with K infinitely massive nuclei
via the Coulomb potential is bounded below by a constant time the total particle
number, i.e. E s >=- C I ( N + K). On the other hand, if all the particles considered
are bosons, Dyson and Lenard [10] showed that the ground state energy (call it Eb)
satisfies E b ~ - - C 2 ( N -t- K) s/3. Lieb [20] showed that this 5/3 bound is indeed the
correct law for infinitely massive nuclei. If the nuclei have finite mass, and are also
bosons, Dyson [9] showed by a variational calculation, that the ground state
energy of bosons is bounded above by Eb <=-- C3(N + K) 7/5. This clearly shows that
bosons are stable in the first sense, but never in the second. Dyson [9] also
conjectured a lower bound E b > - C 4 ( N + K ) 7/5 and this was finally proved 20
years later by Conlon, Lieb, and Yau [4]. They also proved a related bound for
bosonic jellium.
The Dyson-Lenard proof for fermions involved a sequence of inequalities such
that the final bound for C1 is 1014 Rydberg. New proofs were given by Federbush
[12] and Lieb-Thirring [-25] in the seventies. The Lieb-Thirring proof gave a much
better bound on C1 (23 Rydbergs) and related the stability problem to the
semiclassical picture of Thomas-Fermi theory. These matters are reviewed in [19].
The aforementioned considerations are all based on the nonrelativistic
Schr6dinger equation. The kinetic energy operator is the standard p2/2m = - A/2m
(when h = 1). One might wonder whether stability still prevails in the relativistic
case since the kinetic energy then decreases from pZ/2m to (p2+m2)~/2
- m(h = c = 1). Historically, Chandrasekhar [2] was one of the first to ask this
question, but in the context of gravitational interaction instead of Coulomb
interaction. The famous Chandrasekhar model for neutron stars or white dwarfs
consists of a semiclassical relativistic kinetic energy and classical gravitational
potential energy. This simple model remarkably predicted collapse (i.e. insta-
bility of the first kind) and gave a critical mass which is correct, at least
approximately. Despite the success of the simple semi-relativistic Chandrasekhar
theory, the kinetic energy operator,
T = (p2 + m2)1/2_ m,

which it employs is nonlocal and therefore violates a basic physical principle.


Nevertheless it is worthwhile studying this operator for several reasons. When
m = 0, T = IP] and it has the correct inverse length scaling (like the Dirac operator).
Unlike the Dirac operator it allows one to formulate a variational principle for the
ground state energy and thereby to give a rigorous definition of stability without
the necessity of filling the Dirac sea or invoking quantum electrodynafiaics. In any
event, there does not exist a truly relativistic many-body quantum theory at the
present time and it is our belief that the study of Schr6dinger operators based on T
will capture some of the essential features of "the correct theory" when it is
eventually formulated.
Stability and Instability of Relativistic Matter 179

Let us start with the Hydrogen atom by considering the one particle
Hamiltonian ~ defined by
/~1 = (p2 q_ m2) 1/2 _ m - c~z/lx[, (1.1)
where c~= e 2 is the fine structure constant (h = c = 1). This operator was studied
independently by Weder [29] and Herbst [16]. See also Daubechies' paper [7].
Since the difference between the operator (p2 + m 2 ) 1 / 2 _ m and IPl is bounded (more
precisely lpl>(pZ+mZ)l/2-m>lpl-m), the stability of (1.1) is the same as the
stability of
H I = ]p]_ L fl/lxl, (1.2)
7-~

where
fl = rcez/2. (1.3)
Note that (1.2) is homogeneous under length scaling and therefore E~ -= inf specH1
is either 0 or - o o by the scaling tp(x)--+2a/2~p(2x).
A first important fact about (1.2) is the existence of a critical tic = 1, similar to
that of the Klein-Gorden or Dirac theories. Kato [17] stated that tic> 1 and
Herbst [16] showed that tic = t. The ground state energy for the Hamiltonian (1.2)
is E1 = - o e if fl > 1 and E1 = 0 if fl < 1. (In the Dirac theory tic = re/2.)
Returning to the many-body case, suppose we have N electrons with
coordinates xl,..., xN in Na and K nuclei with coordinates R1 ..... RK in IRa and
with positive charges z 1.... ,Za. We shall consider the following relativistic
Schr6dinger Hamiltonian, HNK, for fermions with q spin states (q = 2 for real
electrons). It is the analogue of (1.2):
N
HNK =-- ~ Ipil+aV~(xl,...,xN;R1 ..... RK), (1.4)
i=1
N K
V~(xx.... ,XN;R1,...,RK) =- 2 [xl--xj1-1- 2 ~, zjlxi-R;! -1
l<i<j<N i=1 j = l
+ ~, zizj ]Ri- R~I - 1. (1.5)
l<i<j<K
Note that charge neutrality is not assumed in (1.4), or anywhere else in this paper.
Mathematically, the Hamiltonian HN~ is a quadratic form on the q-state
physical subspace ~fq of L2(IRaN). More precisely, ~pe ~/fq if and only if there exists
a partition P = {zq .... , rcq} of {1,..., N} such that Ip(xl,..., XN) is an antisymmetric
function of the variables in each rej, for all 1 Nj =<q. When q--N, there is no
restriction and the ground state energy for H~¢~ is just the ground state energy for
bosons.
Physically, the nuclear kinetic energies should be included in (1.4) since the
Born-Oppenheimer approximation (i.e. the neglect of the nuclear kinetic energies)
is inadequate in the extreme relativistic regime. For simplicity, we shall confine
ourselves to the Born-Oppenheimer approximation.
In reality, our goal is to discuss stability of the second kind for I-INK(m), which is
given by (1.4) but with IPl replaced by (p2 +m2)l/2_m there. For this purpose,
180 E.H. Lieb and H.-T. Yau

however, it suffices to study only stability of the first kind for HNK in (1.4). The
reason is the following. Let ENK(RI, ...,RK) denote the ground state energy
( = inf spec) of H~,.Kand let EN~ be the infimum of E~,~K(R1.... , RK) over all choices of
the R's. By simple scaling (~p(xl..... xN)---,23N/21p(j~x1, -.-, 2xN) and Rj ~ R i / 2 ), we see
that ENKis either zero or -- or. On the other hand, if ENK(m) is defined analogously,
then, since LP[-m<(P 2 +mZ) 1/z - m < IP], we have that EuK>EN~(m)>EN~--mN.
Thus stability of the first kind for HNK (in the sense that ENd: is bounded below
independent of the R j) is equivalent to stability of the second kind for HNK(m). Our
goal then - and that is the purpose of this paper - is to find necessary conditions
and sufficient conditions on z and e so that EN~(Ra,..., Rt~) > 0 for all N and all K
and all R1, ..., RK.
If everything is held fixed except for q, then EN~(R ~.... , R~) is a monotone
decreasing function of q. The reason is that specifying q is the same thing as
requiring that the admissable wave functions tp(x~.... , xN) are antisymmetric in
each of q sets of variables. The number of variables in each set is unimportant, zero
being an allowed number. Thus, a valid function for q is trivially a valid function
for q + l .
A further remark about (1.4) can be made. Using a convexity argument,
Daubechies and Lieb [81 proved that the stability of HNK for zl = z2 = . . . = zK = z
implies the stability of HN~: when all the nuclear charges are no greater than z, i.e.
O < z j N z for allj. With this remark, we shall assume from now on z~ . . . . . zk=z.
Let ENK(~,z) denote the dependence of ENK o n 0~ and z. We shall use the
following terminology: H(~, z) is stable means that EN~(~, z)= 0 for all N and K.
Otherwise we say that H(~, z) is unstable.
The coupling constant of the electrons to the nuclei is z~ = 2fl/rc and, from the
hydrogen atom result, it is clearly necessary to have fi< 1 for stability. It is
frequently convenient, therefore, to adopt ~ and fl as the independent variables
instead of ~ and z. When doing so we shall refer to the stability or instability of
H(~, fi) - hopefully without confusion. Indeed ~ and fi are the natural variables
from the following point of view. The electron-nuclear coupling is 2fl/zc while the
nuclear-nuclear repulsion constant is z2~ =(2/rc)zfl2/e. Suppose that K > 1 and
fl < 1, but Kfl > I. Then, if the nuclear-nuclear repulsion is ignored, the K nuclei can
come to one common point and the system will collapse - even with only one
electron. What discourages this from happening is the repulsion which is
proportional to flz/e. With fl fixed, we see that e is required to be small in order that
this repulsion prevents collapse. It is a striking fact, and it is the main theme of this
paper, that for every fixed fl < 1 and q there is a critical ~ (call it ~c(~)) so that
H(~, fi) is stable when c~< Ctc(fi). There is another critical ~ (call it ~(fi)) so that
H(a, fl) is unstable when a > g~(fl).These facts are the reason behind the contention
above that e and fi are natural. We do not know whether or not a~(fl)= 5~(fi). Note
that by the above monotonicity in z remark, stability for some (~,fl~) implies
stability for all (~, fl) with fi< fi~.
There is an additional piece of information. Suppose that stability occurs for a
pair cq, z. Then stability occurs for a pair ~, z if a<cq. The reason for the
monotonicity in e is that inf spec(~ IPit + c~Vc)>=(1 - c~/a1) inf spec(~ tpil) q- (o~/a1)
inf spec(~ IP~I + cq V~)> 0.
Stability and Instability of Relativistic Matter 181

Before stating our main results in detail, let us review some recent progress with
this and related problems that also have the feature of critical coupling constants.
(1) The Chandrasekhar critical mass was established up to a factor of 4 in the
framework of the relativistic Schr6dinger equation by Lieb-Thirring [26]. Later,
Lieb-Yau [27] proved that not only is the Chandrasekhar critical mass exactly
correct, but the Chandrasekhar semiclassical equation can be derived rigorously
from the relativistic Schr6dinger equation in the limit that the gravitational
constant G~0. In particular, in the physically interesting case, the discrepancy
between the Chandrasekhar semiclassical critical mass and the quantum mechan-
ical critical mass was shown in [27] to be less than 0.01%.
(2) For the non-relativistic Schr6dinger equation, but with magnetic fields
present that couple to both the electronic orbital motion and electronic spin, the
existence of a critical nuclear charge for the stability of the hydrogen atom was
proved by Fr6htich, Lieb, Loss, and Yau 1-15, 28]. The results were extended to the
one-electron molecule and many-electron atom by Lieb and Loss [23]. The
stability criteria are very similar to that of the relativistic stability considered in
this paper. For stability, one should keep both c~2zand c~small. The general case for
this model (many electrons and nuclei) remains an interesting open problem.
(3) The relativistic stability of matter itself. For N = 1 and K arbitrary,
Daubechies and Lieb [8] were the first to note the existence of a critical c~and/~
fixed. They proved that HtK is stable in the critical case/~ = rcc~z/2= 1 if ~ < 1/3rc.
The first person to solve a general case for all N and K was Conlon [3], who proved
that the Hamiltonian H(c~,z)is stable when z = 1 provided e < 10 -z°° and q = 1.
Using a different method, Fefferman and de la Llave [14] improved Conlon's
result for z = 1 to e < 1/2.06~, and again q = 1. The Fefferman-de la Llave proof
used computer assisted proofs extensively. Without using a computer, their bound
would be worse by a factor 2.5, thereby reducing e to 1/5rc. Recently, Fefferman
[13] announced a result for the critical case /~= 1 provided some numerical
computer calculations can be made rigorous. The stability criterion announced in
[13] is that stability occurs in the critical case/~ = 1 ife < 1/20 and q = 1. A complete
proof, however, was not available when the present paper was written. Since
H(e,/3) collapses for/3 > 1 no matter how small the difference/~-1 may be, the
application of computer assisted proofs to the/~ = 1 case is delicate and difficult.
Fefferman [13] states that "arbitrarily small roundofferrors are apparently fatal."
All the results mentioned above address the situation q = 1. The methods
employed are not, in our opinion, easily generalized to treat arbitrary q, as is done
here, The ability to treat arbitrary q without increasing the complexity of the proof
as q increases is, in our opinion, one of the main advantages of our method.
Another is that we have no intrinsic need to invoke the computer. The essence of
our method is that for all q the many-body problem is reduced to a tractable one-body
problem (see e.g. Theorems 6 and 11). This method also makes it possible to prove,
for the first time, that stability occurs up to and including the critical value fl = 1.
We should point out that the main tool in proving the nonrelativistic stability
of matter, the Thomas-Fermi theory, fails to predict stability in the relativistic case.
The semiclassical kinetic energy decreases in the high momentum region from
(const) ~ 05/3 in the nonrelativistic case to (const)~ ~4/3 in the relativistic case. This
semiclassical kinetic energy, I Q413,cannot control the Coulomb singularity zc(r for
182 E.H. Lieb and H.-T. Yau

any a > 0. The fact that stability occurs only for some finite ~ > 0 and z > 0 is not a
trivial matter (see Conlon [3]). A good estimate for a, especially when fi is set
equal to its critical value 1, is very difficult to achieve and should resolve the
following subtle points:
(i) The delicate balance of charge neutrality. If, for example, the attractive term
in V~is changed from z ~ Z ~ [xi-Rjl-1 to z~(1 + e ) Z ~ f x i - R j J - i for some e>0,
then stability will not occur for any positive ~ and z. Physically, an attractive
gravitational interaction is present and it does alter the Hamiltonian in precisely
this manner - collapse does indeed occur. But the gravitational constant is small,
and this collapse happens only when N and K are extremely large - the order of a
solar mass [26, 27]. Indeed, the problem of determining the critical mass when
Coulomb and gravitational interactions are both taken into account is a difficult
open problem.
(ii) An improved version of the basic inequality lpl- 2 Ix[- 1 > 0 is needed. This
7r
is apparently crucial since each electron in general feels attractions from more than
one nucleus. One may argue that, by virtue of screening, each electron feels only
one attraction from its nearest nucleus, but it is difficult to find a simple, precise
mathematical statement about screening. Indeed, some corrections (e. g. van der
Waals force) are obviously unavoidable and can only be controlled by the kinetic
energy.
(iii) The nonlocality of the operator [p[. The technical problems caused by this
non-locality are serious, especially since the Coulomb potential is long-ranged.
Our main results are the following four theorems about stability and
instability.
Theorem 1 (Simple Stability Criterion). For any z > 0 and q, the Hamiltonian H(~, z)
is stable if
a < sup Aq(z'), (1.6)
Z'~z

where
A~(z) = (2/n)z- i [1 + ql/az - 1/3 C(z)- 1/3] - 1, (1.7)
C(z) -- 3.0844 {[1.6617 + 1.7258z- 1 + 0.9533z- 1/214 + (4/n)3 [1 + (2z)- 1/2]s} - 1.
(1.8)
Corollary. Fix fl-= z~n/2 < 1. Then stability occurs if
q~< ~0.062980(1-fl)afl-2 if fl>0.49910
= ~0.031774 /f fl<0.49910. (1.9)
Remark. There is a number zl, which is roughly 0.6, such that if z > z 1 then the
supremum in (1.6) occurs for z ' = z, while if z ~ z l the supremum occurs for z ' = z 1.
Theorem 2 (Stability criterion for fl < I). Fix fl < 1. Then the Hamiltonian H (~, fl) is
stable if
q~ < 1/47.
Theorem 3 (Instability for all z and q). There is a critical value ~1 such that if ~ > ~1
then H (~, z) is unstable for every q > 1 and every nuclear charge z > 0 (not necessarily
Stability and Instability of Relativistic Matter 183

integral), no matter how small z may be. This means that if ~ > at, one can always
choose N and K so that ENK(e, Z) = -- 00. In order to achieve this collapse, it is only
necessary to use one electron, i.e. N - - 1 . One can take e~ = 128/15n.

Theorem 4 (Instability dependence on q). Let fl = nez/2 as in (1.3). There is a critical


value o~2 such that if
~>~2q-lf1-2 , (1.10)

then H(c~,/3') is always unstable. To achieve this collapse, only N = q electrons are
needed. One can take c~2= 115, 120. Alternatively, e > 36q - 1/3z2/3.
Corollary. I f the electrons are bosons then H(c~, z) is unstable for all o~> 0 and all
f i x e d z > 0 . The number of electrons necessary to achieve this collapse satisfies
N < 4zc- 2o~2z- 2o~- 3.

Remarks. In view of Theorem 3, the number 115, 120 should not be taken seriously.
Its large value merely demonstrates how difficult it is to find simple, rigorous
bounds - even variational upper bounds - for the relativistic Coulomb problem.
These theorems, taken together, give a clear picture about the stability of
relativistic matter. The relevant parameters for stability are c~q (if/~ is fixed) and
0~qI/3 (ifz is fixed). An upper bound for 0~which is independent ofz and q is given in
Theorem 3. ~ is never larger than 1. Theorem 1 clearly fails to predict stability for
the critical case e z = 2 / n , but its proof is considerably simpler than that of
Theorem 2. It also gives the correct q dependence (when z is fixed), and its bound
on 0~for small z is better than that of Theorem 2.
To gain perspective on how good these bounds are, we specialize our results to
the following two cases. First, in the critical case, our upper bound (Theorem 2)
and lower bound (Theorem 3) differ by a factor of 128 for q = 1. Second, for z = 1
and q---1, Theorem 1 predicts stability for e__<1/3.23n, which is not appreciably
worse than the computer assisted proof bound 1/2.06n in [14]. Our bounds in
Theorem i and Theorem 2 can certainly be improved, as will become clear in the
proofs given below. We refrain from the temptation to optimize our results by
complicating the technicalities. Our goal is to give a simple conceptual proof which
has the correct q dependence and reasonable estimates.
Our proofs for Theorem 3 and 4 follow the same idea used in [23, 20].
Theorems 1 and 2 are much more difficult. Our basic strategy is first to reduce the
Coulomb potential to a one-body potential, W. Then, by localizing the kinetic
energy IP[, we can control the short distance Coulomb singularity of W,, leaving a
bounded potential W* as remainder. The last task is to bound the sum of the
negative eigenvalues of IP[ + W*, but this is standard and can be done by using
semiclassical bounds ([6]).
The following Theorem 5 is a consequence of our localization for ]Pl and
combinatorial ideas in [26]. Theorem 5 was announced in [27, Appendix B],
where it was proved for the special case q = N. Earlier, Fefferman and de la Llave
[14] proved it for q = 1. This theorem is not needed in the present work, but it is
independently interesting. (Note that the definition of c5i below is the reciprocal of
that in [27].)
184 E.H. Lieb and H.-T. Yau

Theorem 5 (Domination of the nearest neighbor attraction by kinetic energy). Let


6~= 6i(xl, . .., xN) be the nearest neighbor distance for particle i relative to N - 1 other
particles, i.e.
~ ~ min {]x~- xjI[j 4=i}. (1.11)
Let lp ~ L z ( ~ 3N) be an N particle fermionic function of space-spin with q spin states.
Then
N N
(~,]PillP)>=Caq -1/3 E (lP, 6/-1~9), (1.12)
i=1 i=1

N N
(q~,p2~)~C2q-2/a ~ (~,6/-z~), (1.13)
i=1 i=1

where
Ca =0.129, Cz =0.0209. (1.14)
The organization of the rest of this paper is as follows:
In Sects. II and III, we prove Theorems 1 and 2 assuming an electrostatic
inequality for the Coulomb potential and localization estimates for ]p[. The
theorems used in Sects. II and III are then proved in Sects. IV-VII. The
presentation has been broken up this way in order to stress the conceptual
underpinnings of Theorems I and 2.
Theorem 5 is proved in Sect. V. Some details of our numerical calculations are
explained in Sect. VIII. In the final Sect. IX we prove Theorems 3 and 4.

II. Proof of Theorem 1 (z~ < 2[n)


The proofs of Theorems I and 2 are conceptually much simpler than the following
detailed calculations and technicalities would suggest. There are three main steps
for Theorem 1 and five steps for Theorem 2. Step A is the same for both theorems.
Step A. Reduction of the many-body Coulomb potential to a sum of one-body
N
potentials plus a positive constant, namely - ~ W(xi)+ C. This reduces the
1
problem to that of showing that q times the sum of the negative eigenvalues of the
operator ]p]- W is not less than - C.
In the next step we decompose P,? into regions Bo, B1 .... , BK where the Bi are
disjoint balls centered at the Ri and Bo is everything else.
Step B. We write [Pl = fl]P] + (1 - r) lP[ with fl = zorn~2< 1. In the balls B~, i = 1.... , K
we use flip[ to control the Coulomb singularity of W and prove the operator
inequality
fl]P] - e W(x) > - U(x), (2.1)
where U = W in Bo and U is a continuous function inside each ball.Thus tP4- eW
> (1 - fl)]p]- O.
Step C. The sum of the negative eigenvalues of (1 - fl) [p[ - U is bounded by using
the semiclassical bound due to Daubechies [6].
Steps B, C, D, and E for Theorem 2 will be explained in Sect. III.
Stability a n d Instability of Relativistic M a t t e r 185

In this section we shall state the basic theorems for steps A and B. These will be
proved later in Sects. IV and V. These theorems will be combined here in step C,
thus completing the proof of Theorem 1.

Step A. Reduction of the Coulomb Potential to a One-Body Potential


This step has nothing to do with quantum mechanics or the nature of the kinetic
energy operator. It has to do with screening in classical potential theory. The total
Coulomb potential, V~,is given in (i.5). There are K nuclei located at distinct points
R 1..... R~ in ]R 3 and having the same charge, z. There are N electrons.
Introduce the nearest neighbor, or Voronoi, cells {F~}r~ 1 defined by
Fj={xllx-Rfl<=lx-Rkl for all k4:j}. (2.2)
The boundary of Fj, 0F~, consists of a finite number of planes. Another important
quantity is the distance
D r = dist (R j, 0Fj) = ½min {iRk -- Rr[ IJ=~k}. (2.3)
The following theorem will be proved in Sect. IV. Recall (1.5).
Theorem 6 (Reduction of the Coulomb potential). For any 0 < )~< 1
N ~. 1 2 K
V~(xl.... ' x N ; R i ' " " R K ) > - - i•=1 W ( x i ) + 8 z j'=21D~-I (2.4)

and, for x in the cell Fj, WZ(x)= VVjX(x)=-Gr(x) + F}(x) with


Gj{x) = z Ix -- Rrt-1 (2.5)
F~(x) - for tx-Ral<2D i (2.6)
((V~z+½)lx_Rr[-1 for [x--Rj[>).D i.
Theorem 6 says that when the electron-electron and nucleus-nucleus Coulomb
repulsion is taken into account, V~is bounded below by a positive term [the last
term in (2.4)] consisting of a residue of the nucleus-nucleus repulsion (in fact one
quarter of the nearest neighbor repulsion) and an attractive single particle part W ~.
In each cell F~, Wjz is essentially the attraction to the nearest nucleus (this is the G r
part of Wj~); there is also a small attractive error F~.
There are two essential points in (2.4). One is that the charge z appearing in Gj is
the same as in the original potential V~. The other is the existence of the positive
term. The error term F~ can certainly be improved, especially the long-range part
Ix-Ri[ > J.Dr; we have not tried to optimize F~.
It is interesting to compare our Theorem 6 with Baxter's Proposition I [1]
which says that
N
V~>-(l+2z) ~ 6(xj)-' (2.7)
j=l

with
,5(x)=min{Ix--Rrllj=l,...,K}=lx--Rjt when x~. (2.8)
Fefferman and de la Llave [14] later improved this when z = 1 from 1 + 2z = 3 to
8/3. Our proof is completely different from both proofs of (2.7), as is Theorem 6
186 E.H. Lieb and H.-T. Yau

itself. To reiterate the essential points, our bound has the correct singularity near
the nucleus (namely z and not 1 + 2z) and it also has a positive repulsive term.

Step B. Control of the Coulomb Singularity in Balls


The following formula is well known. For f e L 2 with Fourier transform f,
( f lptf)=(2~)-3 j. if(p)]2 ]pldp=(2z~2)- 1 ~S If ( x ) - f (Y)]2lx - yl-4 dxdy • (2.9)
One way to derive this formula is to write
(f, IP[f ) = lim t-~ {(f f ) - (f, e- qvlf)}. (2.10)
tl0
The convergence is a simple consequence of dominated convergence in Fourier
space. The kernel of exp(-tlpl) can easily be calculated to be
e-ttPt(x, y) = ~ - 2 t [Ix -- y[2 + t 2] - 2 . (2.11)
Inserting (2.11) in (2.10) yields (2.9).
A formula similar to (2.9) can be derived this way for (p2 + m2)~/2 in place of IPl.
( f ( p 2 + m )2 1/2
f ) = ~ r1 c -2
m2 ~ [ f ( x ) - f ( y ) 1 2 [ x - y [ - 2 K z ( m [ x - y f ) d x d y , ( 2 . 1 2 )
where K 2 is a Bessel function. This follows from [11]
exp [-- t(p 2 + m2) 1/2] (x, y) = ½~- 2m2 t([x-- yf2 + t 2)- 1K2(m([x _ y[Z + t 2)1/2).
Starting with formula (2.9) we have (2.13)
Theorem 7 (Kinetic energy in balls). Let B be a ball of radius D centered at z e N 3
and let f e L2(B). Define
1
(f, lP]f)~---- 2-~2 ~ ~ I f ( x ) - f ( Y ) l z l x - Y l - 4 d x d y (2.14)
BB
and assume this is finite. Then
( f tP[f)n ~ D - 1 I Q ( l x - z[/V)[f(x)lZ dx, (2.15)
B

where Q(r) is defined for 0 < r < 1 by


Q(r) = 2/(zcr) -- Y,(r),
2 1 + 3r 2 1 - r2 4r
Y~(r) = + re(1 +r2)r In(1 +r) ~z(1 +r2)r l n ( 1 - r ) re(1 + r 2) lnr
+

< 1.56712 (2.16)


The maximum of Yl(r) occurs at r ~ 0.225975 and was computed by S. Knabe.
Note that Yt(lxt) is continuous for all ]xJ < 1.
Using (2.9) we have
Corollary. I f B1, .. ., B K are disjoint balls in ~R3 centered at R ~, . . ., R r and with radii
D 1, ..., Dr,
K K
2 j=l
IPl> re Z I x - R j l - ~ B j ( x ) - j=a
E Dj~YI(Ix-RyDj)Bj(x) (2.17)

where Bj(x) is the characteristic function of Bj.


Stability and Instability of Relativistic Matter 187

Theorem 7 is proved in Sect. V. Theorem 12, which is the analogue of


Theorem 7 with p2 in place of IPl, is stated and proved in Sect. V.

Step C. Semiclassical Bounds and the Conclusion of the Proof of Theorem 1


The problem of showing that H = ~ I p J + e V ~ > 0 has been reduced to the
N
following. In step A we showed that H > ~ ~ + C, where
1

~i = [P~t- o:WZ(xl), (2.18)


i
C---~ ~ z20~ ~ D f 1. (2.19)
j=l

If we write [pl = fl [Pt + (1 -fi)Ipt, with fl =zero~2, then step B shows that it suffices to
replace rzi in H by hi where
h i = (1 - fl)IPi] - U(xi), (2.20)

U(x)=c~F~(x)+flDflYI(lx--RyDj)B~(x)+zatxI-~(1--B~.(x)) when xeFj.


N (2.21)
Proving that Z h~+ C > 0 for all numbers, N, of q-state fermions amounts to the
1
following inequality in terms of density matrices satisfying 0 < 7 < q. [A density
matrix is a positive definite trace class operator on L2(N3).]
Tr?h>-C for all ?, (2.22)
with h = (1 - fl)IP[- U(x). (Tr?h is shorthand for ~ (fk, hfk)?k, where (f~, 7k) are the
eigenfunctions and eigenvalues of 7.) For more details see [21].
The tool we shall use to prove (2.22) is Daubechies' extension of the Lieb-
Thirring semiclassical bound from p2 to IPl.
Theorem 8 (Daubechies). Let ? be a density matrix satisfying 0 < ? < q. ( q need not be
an integer.) Let U(x) be any positive function in L4(]R3). Then for # > O,
Tr?(# [Pl - U) > - 0.0258q/~- 3 ~ U(x)4dx. (2.23)
To complete the proof we merely insert (2.21) into (2.23). A simple bound is
obtained by extending the integral over each Fj to an integral over all of N 3. This
will give K terms on the right side of (2.23) (each of which scales like D f 1) to be
compared with the K terms in C (2.19). Our condition is then (recalling that

0.0258q(1 - fl)- 3 {1~1~1 [aFZ(lxl) + flY~(]xl)]4dx

-1- f [°~F'~(lx[)q-z°~lxl-1]4 dx~ <=~z20~ (2.24)


Ixl> i )
for some choice of 0 < 2 < 1 and where
y½(1-r2) -1 for O<r<_2 (2.25)
FZ(r)= ~(l/~z+½)r-a for 2<_r.
188 E.H. Lieb and H.-T. Yau

The second integral (Ix[>l) in (2.24) (call it I+) is easy to evaluate. It is


independent of 2,
I+ = (4/re)3/34 [1 + (2z)- 1/218. (2.26)
Next, the integral of Y~ over ]xl < 1 has been done numerically by S. Knabe. The
following is actually an upper bound.
Ixl~ 1 Yl(X)4dx = 7.6245 -~ 11 . (2.27)

We shall take 2 = 10/11. Then

I ~'~(xpclx=(4~/~o) +(2z) 1/~ -x~, (2.28)


2<Ixl<l
2
1~<2~F2(x) 4 dx ~ (K/a) }. ! (1 -/. 2) - 4/•dr -~-13
Ix
= (~2/24) [(1 -- 22) - 3 __ 1] = 22.645. (2.29)
To bound the first integral (lxl < l) in (2.24) one can use the triangle inequality
S (f 4- g 4- h) 4 ~ [(S f4-)1/4 4_(o[g4)1/4 4- (o[h 4) 1/414. Thus our condition for stability is
satisfied if
0.0258q(1 -/~) - 3 {[1.6617fl + 1.0588c~({ + (22) 1/2) + 2.1815el 4
+ (4/rc)3 fi4. [ 1 + (2z)- 1/2] 8} = (2/~:)z,82/&z. (2.30)
Let us rewrite the stability condition (2.30) as
q - 1zC(z ) ~ fl3(1 __ fl)- 3 (2.31)
with C(z) given by (1.8), namely
1/C(z) - (0.0258)4re {[1.6617 + 1.7258z- 1 + 0.9533z- 1/214 + (4/7r)3 [1 + (2z)- 1/2] s}.
(2.32)
By taking the cube root in (2.31) we have that (2.31) is equivalent to the
assertion that stability occurs if
o~< Aq(z) = (2/zt) z-1 {1 + q 1/3Z- 1/3 C(Z)- 1/3} -1. (2.33)
Using the monotonicity in z for fixed c~ [8] mentioned in Sect. I, (2.33) can be
improved to the statement that stability occurs if
c~__<sup {Aq(z')[z' > z}, (2.34)
and this is precisely Theorem 1. []
Next, we address the question of finding a bound on ~ that depends only on/~
and not on z. For this purpose return to (2.31) and solve the equation q - l z C ( z )
=/~3(1 _/~)-3. Since z--* C(z) is monotone increasing, this equation has a unique
solution. Call it Zq(/~). Then stability occurs for any given/~ if

-< ~*(/~) -= (2/70 sup {fl'/Zq(/~')I/~' >/~}. (2.35)


Stability and Instability of Relativistic Matter 189

Again we have used the aforementioned fact that stability for (a, fl') implies stability
for (~, fl) if fl_<fl'.
Formula (2.35) is correct but lacks transparency. We shall now present a way to
find a function a**(fi) which is less than or equal to a*(fi) but which has the same
general features as e*(fl). It is this function, a**(fl) that is given in the corollary.
Choose an arbitrary z o. Let s o = Aq(zo) and let flo =(2/n)aoZo- Define

a**(fl)= ~(2/nq)C(zo)(1-fi)3fi-2 if fl>=fio (2.36)


((2/nq)C(zo)(1-flo)3 fio 2 if fi<=fio.
We claim that ~ < a**(fl) implies stability. First, suppose that fl >_rio. Then we have
z - (2/~)a - l fl > (2/70 [a**(fl)] -~ flo =>(2/n) [a**(rio)] - lflo (since fl > rio) = Zo.
By the monotonicity of C, we have C(z) >_C(zo). Therefore
z = (2/To)~- lfl ~ (2/re) [a**(fl)] - 1/~ ~ qC(z)- 1fi3(1 - / ~ ) - 3 (since C(z) ~ C(zo)).
This is (2.31).
Second, suppose that fi< flo- To prove the stability, we only have to verify
(2.35). For this purpose, it suffices to show that a < 2 fio/Zq(fio) with Z~(fio) solving
7[
q- 1zC(z) = f103(l-- flo)- 3. Since by definition q- ~zoC(zo) = f103(l- flo)- 3, we have
from the uniqueness of the solution of the above equations that Zq(fio ) = z 0 and
a**(fl)=(2/rcq)C(zo)(l-flo)3f102= ~-fio/Z~(fio). Hence a<a**(fl)is the same as
o
~< -~flo/Z~(fio) and thus stability occurs for (a, fl) with a<a**(fi) and fl<flo.
Let us choose Zo=10. Then (2/7[)C(10)=0.062980 and fi0=0.49910. This
together with (2.36) proves the Corollary of Theorem 1. []

III. Proof of Theorem 2 (z~ < 2/n)


In the proof of Theorem 1 we first reduced the many-body Coulomb potential to a
one-body potential in Step A. Then we split the kinetic energy IPl into two pieces.
One of them was used to control the Coulomb singularity and the other was used
to control the long range part of the potential. If the method of Theorem 1 is used
when z~ = 2/7~, all of [p[ must be used for the singularities and nothing remains to
control the long-range potential. In this section both parts of the potential will be
controlled without splitting ]PT,but this requires inventing a suitable localization
formula for [Pl. We shall henceforth take z~ = 2/~; by the monotonicity in z, this
case wilt cover all the cases z~ < 2/re.
There are five steps.
Step A is the same as before. The Coulomb potential V~is replaced by a one-body
potential W z plus a positive constant. Henceforth we shall take 2 =0.97 and omit
the superscript on W.
Step B. Here we show that if ;(l(x) is a C a function which is approximately the
characteristic function of a ball, and if7 is a density matrix with 0 < V< q and if 22(x )
190 E . H . Lieb and H.-T. Yau

is defined by Zl(x) 2 "q-)~2(X)2 = I, then


TrT(lp[- W)~ TrzlyZl([pl-potential energy correction - 141)
+ Trx27)~2(IPl - potential energy correction - W) - q. const. (3.1)
The important aspect of this inequality is this: It might have been thought that
since [PLis not a local operator, the potential energy corrections would have to be
very long range. In fact they have support only inside a ball which is only slightly
larger than the original ball. The long range nature of Ipl manifests itself in the term
q-constant which depends on 117][but not on N = T r T .
Step C. The ball referred to in step B is taken to be B1 centered at R1 (see Sect. II).
To control the first term on the right side of(3.1) we have to bound q times the sum
of the negative eigenvalues of fPl - potential energy correction -- W in a ball, where
W is the one-body potential defined in step A.
Step D. For the second term on the right side of (3.1), the localization process in
steps B and C are repeated K - 1 times for nuclei, 2,..., K. This finally leaves us
with a term Trz07Z0(lpT- potential energy corrections) where Zo is essentially the
characteristic function of the complement of the K balls. To estimate this term,
Daubechies' semMassical bound, Theorem 8, is used.
Step E. The above process leads to a lower bound on inf spec(H) in terms of
certain integrals which depend on certain parameters that remain to be specified.
These numerical facts are presented in this step. The details of the computation are
given in Sect. VIII.

Step B. Localization of the Kinetic Energy


By way of comparison we begin by reminding the reader of the IMS localization
formula (see [5, Theorem 3.2]) for p2 = _ A instead of tPf. Let )~o,Z1, -.., ZK be real
valued functions on •3 satisfying
K
Zj(X)Z=l for all x. (3.2)
j=0

Then an elementary calculation yields the following operator identity.


K K
- A = E Zj(x)(-A)zj(x)-- Y, [Vz~(x)[2. (3.3)
j=O j=0

This is a localization of - A. If we assume additionally that Zj has support in some


set Aj (which are not pairwise disjoint, of course) then for any f ~ L2(p, 3) and any
arbitrary potential V,
K
(f, [ - A + V(x)] f ) = ~ (Zj r [ - A + V(x)- U(x)] zjf) (3.4)
j=O

with
K
U(x)= Z lvzj(x)l 2. (3.5)
j=O
Stability and Instability of Relativistic Matter 191

The advantage of (3.4) is that in the fh term of (3.4) only [-V(x)- U(x)] 1Aj(x)
appears [-where 1A(X)= 1 ifx ~ A and 1A(x)= 0 ifx ¢ A] and one can utilize different
bounds on V - U according to the region Aj under consideration. Furthermore,
since z j f has support in A t one can replace - A by the larger operator - A with
Dirichlet boundary conditions on ~Aj. The price one has to pay for all this is the
negative potential operator - U(x).
For the operator [p[ the following analogue of (3.3) is much more complicated
because IPl is not a local operator. We also state its generalization to (p2 +m2)1/2.
The proof is immediate starting with (2.9) and (2.12).
Theorem 9 (Localization of kinetic energy-general form). Let )~o..... XK be Lipschitz
continuous functions satisfying (3.2). Then for any f E L2~3),
K
( f IPlf)= ~ (Zjf t p l g j f ) - ( f Lf), (3.6)
j=0
where L is a bounded operator with the kernel
1 K
L(x, y) = }~-~ I x - Yl- • E [Zj(x)- Zj(Y)]2 (3.7)
j=O

More generally,
K
( f (192+ mZ)l/zf) = • (zjf (/)2 + m2)zjf)_ ( f L¢m)f), (3.8)
j=O
where D ") is a bounded operator with the kernel
K
L(m)(x,Y)=(27z)-2m21x-yl-2Kz(mlx-Yl) E [Zj(x)-~(j(Y)] 2 (3.9)
j=0
and K2 is a BesseI function.
Formula (3.6) was proposed to us by M. Loss, to whom we are grateful.
A simple, but important corollary of Theorem 9 concerns q-state, density
matrices. As defined in Sect. II, this is any bounded operator on Lz0R 3) which
satisfies the operator inequality 0 < 7 < q and for which Tr7 < ~ .
Corollary. For any density matrix, 7,
K
Trylpl= Y. TrTilPI--TrTL, (3.10)
j=o
where 7j--Xj)'Z~, with gj being thought of as a multiplication operator.
To exploit (3.10) we now impose a condition on go,---, ZK- Let R1 ..... RK be
distinct points in N a (namely the nuclear coordinates) and let Dj be given by (2.3).
The K disjoint balls Bj = {xlx-Rjl <D~} were defined in Sect. II. Choose some
0 < tr < 1 and consider the smaller balls
B~") = {xl I x - Rjt < (1 - t~)Oj}. (3.11)
Let go,..., Zt~ satisfy (3.2) with Zj supported in B~~) for j = 1,..., K. The explicit
choice for Z~ will be made in step D.
192 E.H. Lieb and H.-T. Yau

First, consider the case K = 1. We decompose the L of (3.7) into a tong-range


part, L °, and a short-range part, L*, with L = L ° +L*. Furthermore, L*(x,y)
vanishes if x or y is not in B~ or if I x - y l > a , namely
[ ~ - 2 Ix - Yl-" I-1 - Zo(X)Zo(Y)- ):l(x)zl(Y)] B l(xlBl(y) if [ x - y I < a
L*(x, Y)
lo if Ix-y[>c~,
(3.12)
where Bi(x)= 1 if x e B i and Bl(x)=0 otherwise. Recall that Xo(X)2+ )~l(x)2 = 1 in
the K = 1 case. With these conventions, we have the following theorem which will
be proved in Sect. VI.
Theorem 10 (Localization of kinetic energy-explicit bound in the one-center case).
For K = 1, let L*i be given by (3.12) and L ° = L - L * , with L given by (3.7). For any
positive function, hi, defined on the ball B 1, let
O~(x) = h l(x)- 1 I L~(x, y) h l(y) dy. (3.13)
B1

Let (2i -=±D2TrtL°~22


1 ~, j , i.e.
~-~1 ~ 1 D 2~~ [L(x, y) - L* (x, y)] z dxdy - I (~)+ I (2), (3.14)

1(1)= ½ ~-4D~ ~ Ix-yl-8[1 -Zo(X)Zo(Y)-Xl(x)zl(y)]Zdxdy,(3.15)


x,yeB~ 'x)
Ilxl-lyll>~Ol
I~Z)=~z-'D~ J ~ , ~ Ix-y[-S[1-Zo(X)]~dxdy. (3.16)

bl-lxl>_-~o~
Then, for any density matrix 7 with [fT[[<q, and any e>0,
Tr7 IPl -> Trz1721(IPl - U*(x)) + TrZoTZo(lPl- U*(x))- q(eD O- t f21, (3.17)
where U~(x)=0 for x ~ B i and, for xeB1,
CT~(x)-(~/D~)B]~(x) + 0~(x). (3.18)
Inequality (3.17) looks complicated, but it is not vastly different from (3.3). The
first two terms in (3.17) are the localized kinetic energies (inside and outside the ball
B0. The U* term is a potential energy correction like the U in (3.4), but this
potential has support only in the ball B1. The last term is novel; it involves only the
norm of 7 and not a trace over ~. One might expect that the non-local nature of [Pl
would give rise to a long range contribution to U, but these long range effects can
be bounded by the norm of y - as is done in the last term of (3.17).

Step C. Bound on Negative Eigenvalues in a Ball


Our goal is to give a lower bound to Tr)~l)~Zl([p[- W ( x ) - U~(x)). The following is
our main tool. It will be proved in Sect. VII.
Theorem 11 (Lower bound to the short-range energy in a ball). Let C > 0 and R > 0
and let
Hce = ]Pl- 2 Ix]- 1 _ C/R (3.19)
Stability and Instability of Relativistic Matter 193

be defined on Lz(R 3) as a quadratic form. Let 0 <=7 ~ q be a density matrix as before


and let Z be any function with support in B R = {x] Ix[ =<R}. Then
Tr)~?,)~Hcg > -4.4827C4R-~q{(3/4nR3)~ I)~(x)12dx}. (3.20)
Remark. When X - 1 in BR then the factor in braces { } in (3.20) is 1.
To apply Theorem 11 to our case we take R in Theorem 11 to be (1 - ~)D ~ and
we take C to be an upper bound for ( 1 - a ) D l { a W ( x ) + g ~ ( x ) - ( 2 / n ) l x 1 - 1 }
= ( 1 - G)D1 {eFl(x)+ U*(x)} in the ball Ixl < ( 1 - a ) O 1 . This computation will be
done in Step E.

Step D. The Negative Eigenvalues for the Long Range Potential


Associated with each bail Bj of radius Dj centered at Rj will be a cutoff function Zj
defined by
Zj(X) = X(Ix-- Rji/Dj), (3.21)
where the universal ;( is given by
1 for r<_1--3a
)~(r)-- cos[~(r-l+3~)/4~r] for 1-3o-_<r_<1-o- (3.22)
0 for 1--a_<_r.
Here, it is important that o-< 1/3. We also choose a function hi(x) for x E B j,
hi(x) = h(tx - R y D j) , (3.23)
1 for r=<1-3o- and 1-o-_<r_<l (3.24)
rh(r)= 2_ _ l l r _ l + 2 a [ for 1-3o-_<r_<l-a.
Starting with Theorem 10, Eq. (3.17), we choose some ~ and compute f21, 01(x),
U~(x) using (3.13)-(3.16). We also compute some bound
C -> (1 - tr)D 1 {ccFI(X) + U*(x)} (3.25)
in B~~). By scaling, C does not depend on D1. Then, using Theorem 11, Eq. (3.20),
we have that
E = inf Tr 7([pt- eW) > - qA/D1 + inf Tr(l -)~)1/27( 1 _ )~12)1/2( t P l - ( z W - U~).
~ (3.26)
The first term, qA/D 1, is a sum of two pieces. One is the q(eDt)-a t21 in (3.17); the
other is the right side of(3.20) (call it qA2). The sum is written as qA/D1 because the
various quantities that have been introduced scale in just the right way - so that A
really is independent of D 1 and q.
For the second term on the right side of (3.26) we note the identity
(1 -- ){l(X) 2) [o;W(x) "~ U~(x)] = (I -- Zl(X) 2) [o~W(x)fll(X) + U~(x) fll(X)], (3.27)
where ill(x)=1 if tx-Rl[~l-3~rD1 and ill(x)=0 otherwise. Since
(1 -)~2)1/2y(1 -)~2)1/2 is a q-state density matrix whenever y is, the last term in (3.26)
194 E, H. Lieb and H.-T. Yau

can be bounded below by


inf Tr~ (lPl - c~W (x) fll (x) - U* (x) fl 1(x)) . (3.28)
7

Now we can apply Theorems 10 and 11 to (3.28), using the ball B 2 in place orB>
Since U*(x) = 0 for x ¢ B1 we see that (c~W(x)+ U*(x))131(x) = eW(x) for x ~ B1. This
process can be repeated until all the balls B 1.... , BK have been used. Our final result
(with U* defined as in (3.18) with R1, D1 replaced by Rj, D~) is

E>-A Z D/~+infTr7 lpl- c~W+ U*(x) 13j(x) . (3.29)


j=l ~' j=l j=

To bound the last term in (3.29) we use Theorem 8. This will result in a sum of K
integrals, one for each cell ~. As in the proof of Theorem 1, a further bound is
obtained by pretending that each Fj extends to all of ]R 3. Thus
K
E > = - q ( A + J ) 2 D]-~, (3.30)
j=l
where
J=0.0258 l [(2/~)Ix[ -1 + ~e(Ixl)+ ty*(Ix[)]'ax, (3.31)
Ixl > 1 - 3 a

and where F(r) is given in (2.25) with 2=0.97, and U*(x) is given by (3.18) with
D1 = 1 there.
F r o m (3.30) and (2.4), stability holds if
q(A + J) < -~zZe= (2:g2) - te -1 (3.32)

Step E. Numerical Results


We take a = 0.3 and ~ = 0.2077 (recall that 2 was previously chosen to be 0.97).
Since all quantities have the correct length scaling, we shall refer everything to a
standard ball of unit radius D1=1. The following are the results of the
computations given in Sect. VIII.
Starting with )~(r) in (3.22) we compute 01 - £ 2 in (3.13)-(3.16),
1~1)= 0.05529, /(2) = 0.06042,
0=I~1)+I~2)=0.1157, e- lf2=0.5571. (3.33)
F r o m the definition (3.13) and (3.24) we find that 01(x)=-O([xl) satisfies
O(r) < O*(r) and
~(3n/32)(Z-V2)a-l=0.5751 for r<__l-a
0*(r)= ~(n/64)a_5( 1 + 2 a _ r ) ( l _ r ) 3 for 1-a<r<l.
(3.34)

Using this we have, from (3.18), that


U*(r) =<eB('~(r) + 0*(r) (3.35)
with Bl~)(r) = 1 for r < 1 - a = 0.7 and B(~)(r)= 0 otherwise.
Next, we want to find some C satisfying (3.25). Since 2 = 0.97 > 1 - a = 0.7, we
need only concern ourselves with the first line of (2.25). Note that c~appears in (3.25)
Stability and Instability of Relativistic Matter 195

in the form c~Fl(x) and, since Fl(x ) does not depend on z in the region r < 2 , the
quantity o~F~(x)is proportional to ~ when zc~is fixed. Our goal is to prove stability
when e < 1/47q <=1/47, and therefore we can replace czFl(x) by Fl(x)/47 in (3.25).
Then
C = 0.7 {0.02086 + 0.2077 + 0.5751 } = 0.5629 (3.36)
satisfies (3.25) for r < 1 - a = 0 . 7 .
The right side of (3.20) (with R = 1 - o - = 0.7) can now be easily calculated. It is
qA 2 = 0.1661q. (3.37)
Adding 5-~g2 and A2 we have
A =0.7232. (3.38)
Finally, the integral in (3.31) must be computed. To bound c~F(r)we can use
(1/47)F(r) for r < 2, while for r > 2 we write z = 2/rcc~in (2.25). When r > 2 this results
in two terms in c~F, one of which is proportional to el/2 and the other to e. In both
terms we can take a = 1/47. Thus, we bound eF(r) by 0.1753/r for r > 2 and by
(1/94)(1 - r 2 ) -~ for r < 2 . We then find that
0.7
< ~ [2/72r+(1/94)(1 - r 2 ) -1 +0.2077+0.575114rZdr
(0.0258)- ~(472)-~J_
0.1
0.97
+I [(2/rcr + (1/94) (1 - r 2) - 1 q_20.20(1.6 - r) (1 - r)3] 4 r2dr
0.7
1
+I [2/7~r+ 0.1753/r + 20.20(1.6 - r) (1 - r)3] 4 rZdr
0.97

+ ~ [2/rcr+ 0.1753r- 114 r2dr. (3.39)


1

The first integral, J1, can be bounded by replacing (1 - r 2)- ~ by (1 - (0.7) 2)- 1 and
then doing the integral analytically. The second integral, J2, was done on a
computer. In the third integral, J3, 1 . 6 - r was replaced by 1.6-0.97 and (1 - r 3)
was replaced by (1 -0.97)3; it was then done analytically. The fourth integral, J4,
can be done analytically. We find J1 < 4.435, J 2 ~ 0.17, J3 ~ 0.0135, and J4 "<0.435.
Thus
J < 1.64 (3.40)
and, from (3.32), stability occurs if c~q<1/47. This completes the proof of
Theorem 2. []

IV. An Electrostatic Inequality


Our goal here is to prove Theorem 6 about the Coulomb potential ~ given in (1.5).
A similar theorem can be derived for the Yukawa potential Ix[-1 exp(-#ix[), but
we shall not do so here. We recall the definition (2.2) of the K Voronoi cells
F 1.... , F~:for K nuclei located at R1,..., R r E]R3~ and also the radii Dj in (2.3) which
is the distance of Rj to ~Fj. Since Theorem 6 is trivial when K = 1, we shall assume
henceforth that K > 1. We set
K
V(x) = E I x - Rjl-z, (4.1)
j=l
196 E.H. Lieb and H.-T. Yau

which is the potentiM of K nuclei of unit charge located at the R j, and


6(x) = min {Ix-- Rsltl < j < K}, (4.2)
which is the distance of a particle at x to the set of K nuclei. We set
cb(x) = V ( x ) - 6(x)- 2 , (4.3)
which is the potential of all the nuclei except for the nucleus in the cell Fj in which x
is located. • is continuous but not differentiable.
Let v be any Borel measure (possibly signed) on N 3. We say that v is a bounded
measure if tvI(~3) < oo. In this case ~ cI)(x)dv(x) is well defined since • is continuous
and bounded. We define

g¢.2(v) = ~ f [x-Yl-Xdv(x)dv(.P) - z ~ q~(x)dv(x) +z2 2 ]Ri-Rj[ -1 .(4.4)


l<i<j<_K

The first term on the right side of (4.4) is well defined (in the sense that it is either
finite or + oo) since I x - Yl-a is a positive definite kernel. The following is basic to
our analysis.
Lemma 1. Let v be any bounded measure, let z > 0 and let 4) be given by (4.3). Then
K
go,~(v) >=~ z 2 j~=l Dr l" (4.53

Proof. There is a (positive) measure/~ that satisfies the equation


fxt - j * p = z ~ (4.6)
K
and p has support on 0 F - ~) 0Fj. In fact, # can be computed explicitly as
j=l

# = -(z/4z) A ¢ . (4.7)
More precisely, 0F consists of pieces of 2 dimensional planes separating some F~
from some Fj; on 0~
d#(x) = - (z/2n)no 17Ix - R~I- ~d2x, (4.8)
where d2x is two-dimensional Lebesgue measure on 0Fj, and n is the unit normal
pointing out of Fi. Let
A = - ½z ~ 3(x) -1 dl~(x). (4.9)
Then
K
1 If I x - yl- 1 d#(x)d#(y)= ~1 ~ ~(x)d#(x)= ~z j=l
Z f fx-- Rjl - 1 dfa(x) + A

Z2 K
-- ~ ~(Rj) + A = z 2 2 ]RI-Rj[ -~ + A . (4.103
2 i-~ ~<=i<j<=K
On the other hand, if each part of OF is counted twice we obtain
K
A=(z2/8~) Z ~ I x - R j l - ~ n ' V l x - R j [ -~dzx. (4.11)
j= 1 ~r~
Stability and Instability of Relativistic Matter 197

Let Ij denote the integral in (4.11). The integrand is ½n. VIx-Rj1-2. With Aj
denoting the complement of Fj in ~ 3 (so that OAj = OFj) we have
1
I j = 5 ~ n.Vlx-ajl-2a2x= 1_ ~ Alx_ajl_2dx= - I tx-aj1-4dx.
or~ 2 Aj Aj (4.12)
For convenience in evaluating (4.12) we can take R j = 0 and assume that A i
contains the half-space ((x, y, z)lx > D j}; the reason for this is that (assuming Dj
4: oo) there is another nucleus at some R~ such that the midplane between Rj and Ri
is given (after rotation of coordinates) by {(x, y, z)lx = D j}. Thus

Ij<=- ~ ~ dydz~ d x ( x 2 + y Z + z 2 ) - 2 = - u / D j , (4.13)


-oo --c~ Dj

and therefore
A _ z2 o;1 d=l
Using (4.6) and (4.10) we have that
~.,z(v) = -i1 j f Ix-- y]- 1d ( v - #) (x)d(v - #) (y) - A. (4.15)
The integral in (4.15) is nonnegative (since I x - y [ - 1 is positive definite), and the
lemma follows from (4.14). []
Proof of Theorem 6. There are N points x 1..... x N. If x i is in some cell F~ we shall
replace the unit point charge at x i by a unit charge distributed on a sphere Si but, in
general, the center of S i will not be x ~and the charge distribution on S~ will not be
uniform. Also, S~ is not always contained entirely in F~. (If x~ is in more than one Fj
then an arbitrary choice can be made.) The definition of S~ and the charge
distribution vi on S~ is the following:
(i) If Ix i - Rjl < 2Dj, then S~ is the sphere ~Bj = {x[ Ix-- Rjl = D j}. The charge v i is
determined so that its (continuous) potential Vii-Ix[- a * v~ satisfies

Vi(x)= [[x--xi] -1 for ] x - R j l > Oj (4.16)


IIx-x*I-1Ixi-RjI-1Dj for Ix-RjI<Dj,
where x* is the image of xi with respect to S~, namely
x* -- Rj = O } Ix i - R j]- 2(x i - R j). (4.17)
The potential V~(x)is harmonic inside and outside B j, and v~can be computed from
the formula - A V~= 4rove,but we shall not need this. It is important to note that vi is
nonnegative.
(ii) If ]xi- R~[ > 2Dj and xi ~ Fj, then Si is a sphere centered at xi and of radius ti
given by
h = Ixe-Rjl (1 + If2~) - t (4.18)
The charge distribution v~ on S~ is the uniform one with unit total charge.
Now we apply Lemma I with
N
v= Z vi. (4.19)
i=1
198 E.H. Lieb and H.-T. Yau

In order to utilize inequality (4.5) it is necessary to relate ge. ~(v) to V~.The last term
in (4.4) is, of course, exactly the nuclear repulsion. The first term on the right side of
(4.4) (call it I) satisfies

I= ~, JS l x - y l - ldv,(x)dvk(y)+ ~1 i=l
~ i~ix_yl_~dv~(x)dv~(y). (4.20)
l <=i<k<=N

Each v,vk integral in (4.20) is less than or equal to I x , - xk[- 1. This is so because, by
construction
(lx1-1 *vi)(x)<_lx-xit -1 , all x, (4.21)
and hence
(Ixl-,, vi) (x)dvk(x) ~ (lxl-,, v0 (x3 ~ lxk- x,L- 1. (4.22)
The vy, integral in (4.20) is just the self energy of v,. Call it e,. There are two cases.
(i) I x , - R j[ < 2Dj. Then, from (4.16)
e, = ~I i x - yl -~ av,(x)dvi(y) = f I x - x,l-1 dye(x)= V~(~0
={x,-x*I-11x,--RsI-1Ds=Dfl(1-Df2fxi--RjI2) -1 (4.23)
(ii) Ix, - Rjl > 2Dj and x, e Fj. Here e, = 1/h since v, is uniformly distributed on a
sphere of radius t,.
To summarize,
1 ~ {Eq.(4.23)incase(i)~
I< ~ Tx,- Xkl - 1 + . . (4.24)
- l<-,<k<-N 2 i=1 1/t, in case (ii)J
The second term on the right side of (4.4) is a sum of z ~ cbdv,. Again, there are
two cases.
(i) I x , - R j l < 2Dj. From the definition of W and the fact that (lxl-l* vi)(x)
= Ix - xil- 1 when x ¢ Fj, we have
K
• (x)dvi(x) = Z l x i - Rkt-1 _ Ix,-- Rf1-1 (4.25)
k=l

(ii) Ixi- Rfl > 2Dj and xi ~ Fj. By the definition of


K
[~(x)dvi(x)= 2 IlX--Rkl-ldv,(x) - f a ( x ) - l d v i ( x ) , (4.26)
k=l

where 6(x) is the distance to the nearest nucleus. Since every R k (including Rj) is
K
outside S,, the first term in (4.26) is merely 2 tx,--Rkt -~. The difficulty in
k=l
estimating the second term in (4.4) stems from the fact that v, can have support in
several cells - not just Fj. We have, however, that for Ix-x,l = t, and any k,
Ix - - e k [ q- ti = lx - Rkl + Ix - xil > IRk-- Xi[ > [R j - xil . (4.27)
Hence 6 ( x ) ~ IR a - x i l - t,, and therefore in case (ii),
K
J ~(x)dv,(x) > Y. I x , - Rkl- 1 _ (iRa_ x,l - t,)- ~ (4.28)
k=l
Stability and Instability of Relativistic Matter 199

Using these inequalities and the definition (4.18) we find that


N
e.,~(r) <=v~ + 2 W~(xO , (4.29)
i=i
with Wa(x) given in (2.5), (2.6). This, together with Lemma 1, proves
Theorem 6. []

V. Simple Localization of the Kinetic Energy


Here we shall prove Theorem 7, but before doing so let us motivate Theorem 7 by
stating the analogous Theorem 12 below for p2 instead of [Pl. This latter theorem is
simple to prove, but we have not seen it in the literature.
Theorem 12 (The energy ofp 2 in balls). Let B be a ball of radius R centered at z ~ IR 3
and let f ~ L2(B) and Vf s L2(B). Define
(f, p2f) B-- [. [Vf(x)l 2dx. (5.1)
B

Then
(f, p2f) B>=R - 2 1 H((x - z)/R)]f(x)] z dx, (5.2)
B

where H(x), for lxi < 1, is any function of the jbrm H(x) = - h- l(x)Ah(x) and where
h is a smooth, strictly positive function with vanishing normal derivative on the
boundary Ix[ = 1. In particular, by taking h(x) = (]x[2 + t) - 1/4 exp ['¼]x[2/(1 + t)], and
then letting t--O (using Fatou's lemma) we have that (5.2) holds with
H(x) = ¼]xl- 2 _ Y2(Ix]), Y2(r)= 1 + ¼r 2 . (5.3)
Remark. It is important to note that ¼, the coefficient of the Ix1-2 singularity, is
precisely the sharp constant for the uncertainty principle in all of IR 3, (f, p2f)
~ ¼ I [f]2 ] x ] - 2 d x .

Proof Write f(x)=g(x)h(x) so that Vf = hVg+gVh. Then


IVfl 2= I h21ggl 2+ ~ tgt2(Vh)2+ I (Vg2) hVh. (5.4)
B B B B
Integrating the last integral by parts
I lVf[ a> - I g zhAh= I f zH. (5.5)
B B B

Equation (5.3) is merely a calculation. []


We turn now to the problem of proving Theorem 7 which is the analogue of
Theorem 12 for
( f tPl f)n = (27rz)-1 l I If(x) - f(y)[2 Ix - Y[- 4 dxdy. (5.6)
BB

If B is IR3 then this is just (f, tpl f); see (2.9).


Proof of Theorem 7. Without loss of generality we can take z = 0 and R = 1. First,
we regularize I x - y 1 - 4 to Lt(x,y)=(lx-yl2+t) -z. The theorem will follow by
letting t ~ 0 and using dominated convergence and Fatou's lemma.
200 E . H . Li e b a n d H.-T. Y a u

With L t in place of Ix--y[ -4 we have


(f, tP[ f)B.,= n - 2 t If(x)] 2 Kt(x)d x _ n - 2 I ~ f ( x ) f(y)Lt(x, y ) d x d y , (5.7)
13 BB

K,(x) = ~ L,(x, y)ay. (5.S)


B

The second integral in (5.7) can be bounded above using the Schwarz inequality as
follows. Choose a real valued function h with h(x)> 0 for all Ix] < 1. Then
f f f L t = f I [f(x)h(Y)l/2/h(x) 1/2] [f(Y)h(x)l/2/h(Y) 1/2] Ct(x, y)dxdy
BB

<- f If(x)12th(X) dx (5.9)


B

with
q,(x) = h(x)- ~ ~ L,(x, y)h(y)dy. (5.10)
B

We make the choice that h is radial, i.e. h(x) = h(r) with r = txl. To compute Kt
and ~h we can do the angular y integration. With [y[ = s we have
l,(r,s)- f L,(x,y)do~,=[n/rs] {[(r-s)2 +t]-t-[(r +s)2 + t ] - l } . (5.11)
Combining (5.7)--(5.11) we have that
( f [P[ f)B,t > I If(x)[ 2 Qt([xl) dx , (5.12)
B

with
t
Q,(r) = n - 211,(r, s) [1 - h(s)/h(r)] s2 ds . (5.13)
0

Finally, we choose
h(r) = (1 + r2)/r. (5.14)
(Note that dh/dr = 0 at r = 1.) The integrand in (5.13) is then
nr-l(l+r2)-l(s-r)(1-rs){[(r-s)2+t]-l-[(r+s)2+t]-l}. (5.15)
At this point we can let t ~ 0 by recognizing that the integral in (5.13) becomes a
principal value integral in the limit, i.e. Qt-~Q with
Q(r) = 4 n - 1(1 + r 2) -1 :~(s - r) - 1(r + s) - 2 (s - rs z) ds. (5.16)
To do this integral (call it I) we set
1
11 = ~ ( s - r)- 1 (r + s)- 2sds = [2r(l + r)] - 1 - (4r)- 1 In [(1 + r)/(1 - r)], (5.17)
o

The remainder of I (namely the rs 2 term) is


1
- ~ rs(r+s)-Zds-rZI1 = - rln[(1 + r ) / r ] + r ( r + 1) -1 - r a i l . (5.18)
0

By combining (5.17), (5.18), Eq. (2.16) is derived. The m a x i m u m of Y~(r) was


computed numerically by S. Knabe. []
Stability and Instability of Relativistic Matter 201

With the help of Theorems 7 and 12, the proof of Theorem 5, which was stated
in Sect. 1, can now be given.
Proof of Theorem 5. Fix 0 < L < N and M = N - L and consider any partition
P = (re1, zc2)of {1.... , N} into two disjoint sets with L integers in rq and M integers
{
L,] such partitions, For each P we de fil n e
in n2. There are ~ N'~

6i(n2)=min{[xi--xj[[j6n2 and j+-i if i@7C2}. (5.19)


First the operator IPl will be considered. Define the N-particle operator
he = E IP,I- 2 Y. 6i(22)- 1 + c~ Z 61(~2)- 1 (5.20)
iEZgl iE~ 1 iE;~2

for some 2, e > 0 to be determined later. Let the N-particle operators H a n d / ~ be


given by

H= .£ ~ hp, (5.21)

N N
I1= Z tP,l-Clq -a/3 Z 671. (5.22)
i=1 i=1

If H a n d / ~ are compared we observe that the IP~t terms are identical. The
potential energy terms are more complicated, but we wish to choose 2 and ~ so that
n > H . To this end, fix xl, ...,xN and let x~i) be a nearest neighbor of xi, that is
Ixj(i)- xit = rain {Ixk-- xil]k # i}. It is obvious that 6i(rCz)- 1~ 6 i - 1 SO that the last
N
term in (5.20), when summed on P, is at most az ~ 6/- 1, where
1

z= ~- - L

To bound the middle, or 2, term in he we note that for each i E {1.... , N} there will
/

be (N-L_I2) partitions in which i~n I and j(i)En 2. Therefore this middle sum in
\ /
N
he, when summed on all partitions, is at least )~v~ 671, where
1

(N)-IN(N-2) N-L (5.24,


v= -L L - I - N-1
Consequently,/~ >_>_H if
C1 q-1/3 < ( N - L ) I-2(N - 1)- 1 _ ~L-1]. (5.25)

Assuming (5.25), Theorem 5 will be proved if we show that 0P, hs,~P):>0 for
every P. Since permutation of the labels in zq and rc2 is irrelevant, it suffices to
prove this for any one P. To this end we henceforth change notation so that
xl .... ,xLelR 3 are the variables in the rq block and Rt . . . . . R M e ~ 3 are the
variables in the zc2 block. Obviously we can assume that the R~ are fixed and
distinct and that q~is then a function of xl,..., xL with q-state Fermi statistics. We
202 E.H. Lieb and H.-T. Yau

shall also drop the subscript P on he. Thus, we want to show that h__>0 for all
choices of the R~. Since h is a sum of one-body operators, we have to show that for
any density matrix 7 with 0 =<'s,--<q,
M
rrT(lpl- V) => - c~ 2 (2Dj)- 1, (5.26)
j=l

where V(x) and Dj are defined by


V(x) = -- 26(x)- 1, (5.27)
2Dj = rain {IRj- Rkl[k = 1,..., M but k #:j}, (5.28)
6 (x) = rain { Ix - Rjtlj = 1,..., M}. (5.29)
Under the assumption that 2 < 2/re, we write lP[ as the sum of two pieces [p[
=(2rc/2)lpl+(1-2n/2)lp[. We also introduce the Voronoi cells rj={xllx-Rjl
<tx--Rkl for all k=t=y} and the balls BjCFj defined by Bj={xsFjIIx-RjI<Dj}.
Obviously M
( f [PI/)> ~ ( f [P[f)Bj, (5.30)
j=l

where the right side is the sum of the kinetic energies in the balls Bj defined in
Theorem 7, (2.14). Using Theorem 7, we have that
M
()~rc/2)(f, lplf)~(Rrc/2) ~ D~ 1 ~ [f(x)lZQ(lx-RyDj)dx, (5.31)
j= 1 Bj

with Q given by (2.16). Hence


Tr 7([p[- V) > Tr7 [(1 - 2n/2)[Pl - 2 W], (5.32)
where W is given in each Fj by

Wj(x) = ~ [ x - RjI- 1 if [ x - R j[ > Dj (5.33)


((n/2)D]-~Y~(Ix--RyDj) if ]x-Rjl<Dj
with Y1 given in (2.16).
Next, we use the Daubechies bound, Theorem 8,
Tr7 [(1-2zc/2)lpl-2W]>_ -0.0258q[1 -).re/2]- 324 ~ W(x)4dx. (5.34)
The integral in (5.34) is a sum of integrals over each Fj. To obtain a bound we
shall merely integrate each I x - Rjl term in W [see (5.33)] over all Ix - Rjl > Dj and
omit the restriction that x E Fj. The integral outside each ball Bj is thus
S I4~4=4rc/Dj • (5.35)
~Bj

The integral inside Bj is (see (2.27))


I WJ4=(rc/2)4Dj -1 S Yl(x)4dx=g6.418/Dj . (5.36)
Bj ]x I < 1

Combining (5.34)(5.36) we find that (5.26) is satisfied provided


qA24(1 - 2rc/2)- 3 < ½e (5.37)
Stability a n d Instability of Relativistic M a t t e r 203

with
A = 0.0258 [4z~+ 46.418] = 1.522 (5.38)
and provided 2 < 2/re. We shall choose ~ so that (5.37) is an equality. We shall also
write 2=Sq -1/3. Then (5.25) is satisfied if C 1 satisfies the following for some
O<<.X<2/rc and some 0 < L < N :
C1 < ( N - L) I X ( N - 1)- i _ AX4(1 --Xrc/2)- 3L- 1]. (5.39)
(Here we have used the bound that 2rc/2 < Xrc/2, which holds since q > 1.)
Consider the case N > 3. To utilize (5.39) we make the following choices
X = 1/5 and L = {(B/X)I/2N}, (5.40)
where B---AXa(1-Xrc/2)-3=O.O075486 and where {a} denotes the smallest
integer > a. Write L = l + e with l = N(B/X) 1/2 and 0 < e < 1. We claim that when
N > 3,
( L - I) X/(N - 1) + BN/L ~ IX/N + BN/l. (5.41)
Assuming this for the moment, we would then have that (5.39) is satisfied with
C 1 ~---( X 1/2 - B1/2)2 ~ 0.129, (5.42)
which proves Theorem 5 when N > 3. If N = 1 there is nothing to prove. If N = 2,
Theorem 3 is trivial because it asserts that
[Pli q- [Pz]=
> 0.129q- 1/3 IX1 -- X2i- 1 , (5.43)
but we already have the simple bound [PlI > (2/re)Ix1- x21-1 for all Xz.
To prove (5.41), insert L=l+e in the left side and multiply by N(N-1)L1
(recalling that I=N(B/X)I/z). Then (5.41) is equivalent to
N t - l(1+ 2~) + N~(1 - e) ~ 0. (5.44)
Since l < N/5, (5.44) holds for N > 3.
The proof for p2 in place of [p[ follows the same route, but using Theorem 12 in
place of Theorem 7 and using the Lieb-Thirring [25] bound in place of the
Daubechies bound. This is
Try(kLpz - 2W) > - qa#- 3/2~5/2 S W(X)5/2dx.

The best bound for o-is obtained in [22] and is o-=0.040305. We split the operator
p2 into 42p 2 + (1 -- 42)p 2, and take the # above to be (1 - 44). Using Theorem 12, W
is given in each cell Fi by
flx-Rj1-2 if [x-R~[ >Oj (5.45)
W(x) = [4Dj- 2 Y2(lx- Rjl/Dj) if I x - R~I < Dj.
The analogue of (5.35), (5.36) using Y2(r)= 1 + r2/4, is
1
w=~D2~3 W~{x)S/ZKx=27~4-128=!(1 +rZ/4)S/2rZdr.
Using (1 + r2/4)~/2~ 1 4-r2/8 in the above integral we find w < 198.2.
204 E.H. Lieb and H.-T. Yau

Setting 2 = Xq -2/3, the analogue of (5.39) is


C2 < ( N - L) [X(N -- 1)- 1 _ AXS/2(1 _ 4 X ) - 3/2L- ~] (5.46)

with A = o-w = 7.988. F o r N > 3 we make the following choices:


X=1/20 and L={(B/X)~/ZN}, (5.47)

with B = AX5/2(1 - 4 X ) - 3/2 = 0.006241. Again, setting L = 1+ e with 1 (B/X) 1/2N,


=

we have to verify (5.41), which is equivalent to (5.44). This inequality is true for
N > 4 s i n c e / = 0 . 3 5 3 3 N . With (5.41) satisfied we have that
C 2 ~ ( X I/2 -- nl/2) 2 ~ 0.0209. (5.48)

This proves T h e o r e m 3 for N >4. W h e n N = 1 there is nothing to prove, while for


N = 2 we require
p2 + pZ > 0.0209q- 2/3 tx~ - x21- 2 (5.49)

Since p~ > l [ x 1 - x 2 [ - 2 for all x2, inequality (5.49) is satisfied. F o r N = 3 it suffices


to have
p2 =>0.0209q- 2/3 {ix1 _ x21 - 2 + ix 1 _ x31 - 2}, (5.50)

and this is clearly true by the inequality just mentioned. []


Remarks. In the above proof, the inequality for p2 was proved in a fashion
analogous to that for lPt by substituting T h e o r e m 12 for T h e o r e m 7. However,
a n o t h e r p r o o f for p2 can be given by using the IMS localization [see (3.3)] instead
of T h e o r e m 12.

VI. Refined Localization of the Kinetic Energy


Proof of Theorem 10 (Sect. III). Starting from the Corollary of T h e o r e m 9, we see
from (3.10) that our task is to find an upper b o u n d to TrTL with L= L ° + L* and
with
L(x, y) = rc - 2 Ix - yi 4 [1 - )~o(X)Zo(Y)- )~(x)gl(y)] (6.1)
and

L,(x,y)=SL(x,y)BI(x)BI(y) if Jx--yl<=a
(6.2)
if ]x-y]>a.
Recall that B 1 is a ball of radius D1 centered at the origin. By simple scaling we can,
and shall take D ~ = I ; we shall also write BI=B. We have Z l ( x ) = 0 unless
[x[ < (1 - a), i.e. unless x s B (~).
We first bound TrTL °. Notice that when txl < lyl, L°(x, Y) = 0 unless Ixl < (1 - ~).
Using the symmetry of L o we can write
Tr7 L ° = 2 R e S ~ ~l/2(x,z)71/2(z,Y)L°(x,y)B(~)(x)dxdydz, (6.3)
txl < ly[
Stability and Instability of Relativistic Matter 205

where ~1/z is the operator square root of y. We do the y integration first and then
apply Minkowski's inequality to the x integration. For any ~ > 0,
TryL ° < a j]" lyl/2(x, z)12B(~)(x)dxdz
+ ~-1 j S I,, >S,x, 71/2(z, Y)L°(x, Y)dy 2B(~)(x)dxdz. (6.4)

The first integral is just


J ?(x, x)B('~)(x)dx. (6.5)
In the second integral we do the z integration before the x integration and obtain
JJ Y(Y' Y')(Ja L°(x" Y)L°(x' y')dx)dydy', (6.6)

where A is the region Ixl <min((1 - a), tYl, lY'I). The factor in parentheses in (6.6) is
the kernel of a positive definite operator, so we can bound (6.6) by
11711j aj L°(x, y)2 dxdy, (6.7)

where A is the region rxl < (1 - G) and [Yl> rx[. In view of the fact that L°(x, y) is
symmetric and L°(x, y) = 0 unless at least one of lxl or lyl is less than (1 - a), and
given that I1~II=q by assumption, (6.7) is just ½qTr(L°) 2. Thus,
Tr 7L ° < ~ j 7(x, x)B('~)(x)dx + qe- 1~21 (6.8)
with Q1 = ½Tr(L°) 2. The verification of the two integrals for O~ in (3.15), (3.16) is
evident if one recognizes that Zo(X)= 1 and ~ 6 x ) = 0 for lxl > ( 1 - ~ ) .
Now we turn to Tr yL*. Since 7 is a positive operator, its kernel satisfies lY(x, Y)I 2
< 7(x, x)y(y, y). Hence, since L*(x, y) > 0 and hi(x) > 0,
TrTL* = j j 7(x, y) L*(x, y)axay
N ~J [y(x, x)hl(y)/hl(x)] 1/2 [7(Y, Y)hl(x)/hl(y)] 1/2 L~(X,,y)dxdy
j j [7(x, X)hl(y)/hl(x)] L~(x, y)dxdy
= J 7(x, x)Ol(x)dx. (6.9)
The second inequality in (6.9) is the Schwarz inequality, together with the
symmetry in x and y. The idea of using the Schwarz inequality in this fashion goes
back to Hardy and Littlewood; see [18] for another application.
When inequalities (6.8) and (6.9) are inserted into (3.10), the Corollary of
Theorem 9, the result is Theorem 10. []

VII. Estimates of Negative Eigenvalues


Proof of Theorem lI (Sect. III). It obviously suffices to consider the case q = 1. Let
the kernel of ? be
7(x, y) = • z~f~(x) f~(y) (7.1)

with 0 < z ~ < 1 and Z z~< oo and with the f~ being orthonormal. Let g~(x)
-= ;~(x)Z(x). We want to prove that, with V(x) = 2/(~z txt) + C/R,
E = Z (g~, (IPl- V)g~)> -4.4827(3/4~R3)C4R -1 IiziI22. (7.2)
206 E.H. Lieb and H.-T. Yau

By scaling it clearly suffices to prove the theorem for R = 1, which we assume


henceforth.
It is convenient to use Fourier transforms. Let
~(p, q) = ~~ ]~(x)z(y)7(x, y) exp(ip • x --iq. y)dxdy. (7.3)
Since Z~'Z is positive semidefinite, so is 6, and hence
]O(P, q)l < Q(P,p)U2o(q, q)1/2 ~ #(p)#(q) (7.4)
with #(p)= e(P, p)1/2. F r o m (7.3) and the fact that 0 < 7 < 1 as an operator,
#(p)2 =(np, 7np) <(np, nv) = ~ [Z(x)[2 dx-- M 2 , (7.5)
where np(x) = Z(x) exp( - ip. x) and M = 1t)~]t2. Using the Fourier transform of Ixl- 1,
namely
4rc[p[-2 = $ ix[- 1 exp(ip- x)dx, (7.6)
E can be written as
E=(27z)-3{~Q(p,p)([pt-C)dp-~z-3~(p,q)lp-q[-Zdpdq}. (7.7)
Using (7.5) we have that
E>(2rc) -3inf{/~(#)10<#(p)<M for all p}, (7.8)
where/~(#) is defined by
/~(#) = S It(p)2 (lpl- C)dp - ~-3 ~~#(p)it(q)]p _ qt- 2 dpdq. (7.9)
To b o u n d the second integral in (7.9), let
A 2 if lPl<=a
(7.10)
h(p) = tPl-2 if IPl > A ,
where A is some constant to be determined later. Employing the same strategy as
in (6.9) we have
~ It(P)#(q)tP -- q]- 2 dpdq
= ~ it(p) (h(q)/h(p))a/2#(q) (h(p)/h(q))~/2 IP- ql- 2dpdq < ~ #(p)2 t(p)dp, (7.11)
with
t(p)= h(p) - ~~ tp- ql- Zh(q)dq
=h(p)-l {~ [p-ql-Zq-Zdq-s(p)}=h(p)-t {~alpl-l-s(p)}, (7.12)
and with
s(p)= ~ tp-ql-Z(q-2-A-2)dq. (7.13)
Iql<A

TO calculate s(p) we use bipolar coordinates, i.e. for any functions f and g

Sf(JP-- qJ)g([qt)d3q = (2r(jpj) ~o flf(fl) ~tI~l-pt


f lPl~lJ°~g(a)d°~}dfi" (7.14)
Stability and Instability of Relativistic Matter 207

Thus,

s(p)=(2n/lpI) ~ (3- 3A-2)LtIpI~


Ivl;a_¢l c~-~d°~} d/3
~1¢ ( l + u_~
= (2rc/Ipl) ~ (u-: -- u~ z) In du (7.15)
o \11-~II
with ~ = IpI/A.
We claim that
~(8n/3)A. for [p[ > A
(7.16)
s(P)>g(P)=[4~lPl-:[x~-+~2-2r'+!~2+~;:3~8
,~ 6 ~ 36~ J for tpl<A.
We shall prove (7.16) later. For now, let us insert (7.16) into (7.12), and then into
(7.11) and (7.9),
E(y) ~ iplI>A ~t(p)z 1-8A(3~2) - a_ C] dp + Ipt~<A y(p )e liP[ -- AZ ]Pl -~
+ 7:- 3A Zg(p) _ C] dp. (7.17)
We choose
A= 3nZC/8 (7.18)
so that the first integral in (7.17) vanishes. Then, using (7.18) and performing the
angular integration,
/~(y) > 4hA4 i y(Aw) 2 { w + 2wZ/3zc2 + 5w3/9n 2

-I~-40/(9zc2)]w-:-32/(3zcZ)}wZdw. (7.19)

As is easily seen, the factor { } in (7.19) has its maximum at w = 1 and it is negative
there. Therefore the infimum of the right side of (7.19) over the set #(Aw)<M
occurs for tt(Aw)= M for all 0 < w < 1. The right side of (7.19) with # = M is
-(598/1357:)A4M 2 . (7.20)
Returning to (7.8) and using (7.18) and (7.20) (with 598 replaced by 600) we have
that

Since M = tlZil2, (7.21) is the same as (7.2).


To complete the proof we must bound (7.15) by (7.16). When u =<1/~, the factor
u-l--u~2>O. When ~>1 (i.e. IpI>A), u<l and we have the bound
in [(1 + u)/(1 - u)] ~'__2u. (7.22)
Inserting (7.22) into (7.15) yields the first part of (7.16).
If IPl <A, then ~ < 1. The integral in (7.15) from 0 to 1 can be done explicitly,
1
(u-: - u~ 2) In [(1 + u)/(1 - u)] du = n 2 / 4 - ~ 2 (7.23)
0
208 E.H. Lieb and H.-T. Yau

To b o u n d the integral from 1 to 1/4, use the fact that for u > 1,
ln[(1 + u)/(u- 1)] > 2u- 1 + ~ u - 3. (7.24)
Then

(u- 1 _ u~2) In [(1 + u)/(u -- 1)] du


1
1/~ ( 2)
=> ! ( u - l - u ~ z) 2 u - a + s u - 3 du=20/9-4~+4~z/3+4~3/9. (7.25)

When (7.25) is combined with (7.23) (and the 443/9 term is replaced by the smaller
quantity 543/18) the result is the second part of (7.16). []

VIII. Some Numerical Calculations


Our goal here is to derive the bounds (3.33) for (2 and (3.34) for O(r).
(A) Evaluation of (2. Q is defined as the sum of the two integrals in (3.15), (3.16).
Recall that a = 0.3 and Z ~(x)= X([x[) is given in (3.22) while X0(x)2 = 1 -Xx(x) 2. We
already set D 1 = 1.
To evaluate I (1) we use the spherical symmetry of Z and first do the angular
integration on x and y. This integral is

ix-- Yl - 8 dooy - 2re i ( x2 "j- y 2 __ 2xy cos 0)- 4 sin OdO


0
= (7~/3)([x] ]yl) -~ {(txl-lyl)-6-(lxl + lyl-6)}. (8.1)
Thus,
1-2o" 1 -a
I(1)=4(3rc2) -1 I sds I tdt[(t-s) -6
0 s+a

- ( t + s) -6] [1 - ( 1 -Z(S)Z)l/z(1 -)~(t)z) 1/z - Z(s)z(t)] 2 . (8.2)


(Note that we integrate over t > s + o- and s, t < 1 - a, and then multiply by 2. Since
s < t - a and t < 1 - a, we have that s < 1 - 2a.) This integral is not elementary, but
because it is an integral of a continuous, bounded function over a bounded domain
in N z it can be confidently evaluated on a computer. The result is (3.33).
To evaluate i(2), the angular integration over y is done first as before, with the
result (8.1). Then 1(2) is the sum of three integrals according as Ix] < 1 - 3a, 1 - 3a
___<]xl<l-2a, 1 - 2 a < [ x ] = < 1 - a . Thus,

I(2)=(4/3n2) S sds ~ tdt[(t-s)-6-(t+s) -6]


0 1--a

+(16/3nz) f sds tdt[(t-s)-6-(t+s)-6]sin 4 0-a-s)


1-3a 1-a
+(16/3~ 2) S sds ~ tdt[(t-s)-6-(t+s)-6]sin 4 (1-a-s) . (8.3)
1-2a s+cr
Stability and Instability of Relativistic Matter 209

In each case the t integration can easiIy be done analytically. This transforms (8.3)
into three integrals over the b o u n d e d intervals 0 =<s < 1 - 3o., 1 - 3o'__<s < 1 - 20-
and 1-2o._<s_< 1 - a . T h e integrands are again b o u n d e d and continuous so
numerical integration can be used. The result is (3.33).
( B ) Bound on 0(r), Eq. (3.34). The function 0 =- 01 is defined in (3.13) with h defined
in (3.24). Again we take D 1 = 1. The kernel L* is given in (3.12) with )~1 ----)~given in
(3.22) and X~ - 1 --)~2
We want to c o m p u t e
I(r) = f L*(x, y)h([yl)dy (8.4)
with r = Ix[, Since the angular integral of tx - y[ - 4 is less than z~(rs)- 1(r - s) 2, with
s = lyl, we have that
1
I (r) <_(1/nr) f (r -- s)- 2 h(s)m(r, s)sds , (8.5)
0

where re(r, s)= m(s, r) and, for r < s, re(r, s) is given by

m(r,s)=
{ 1-cos[~(s-~)/4a]
1-cos[rc(s--r)/4o]
l--cos[rc(2a+'c-r)/4o]
0
for O<_r<_z<_s<_r+o
for z<_r<_s<_min(~ + 2o, r +a)
for s--o.<_r<_z+2o.<s<<_l
otherwise.
(8.6)

In (8.6), z = 1 - 3~.
The arguments of the cosines in (8.6) are all at most rc/4 and one can use the
inequality cosb > 1 - b2/2 for [b[ < re/4. if we use this inequality in (8.6) and then
insert the result in (8.5), the integral (8.5) is seen to be elementary but tedious [recall
(3.24)]. Finally, O(r)=I(r)/h(r).
Let us verify (3.34) when r > 1 -o-. Then rh(r)= 1 and thus
1--O"
O(r) = (rc/32a 2) ~ sh(s) ( r - s)- 2 (1 - ~ - s) 2 ds. (8.7)
y-o"

The second line of(3.24) is appropriate for this region. In the region r - ~r < s_< 1 -
the function ( r - s ) - Z ( 1 - o . - s ) 2 is m o n o t o n e decreasing in s and so has its
m a x i m u m at s = r - o . Thus,
1-o-
O(r)<(~z/3Zo'2)o'- ~(1-r) 2 ~ { 2 - o - ( s - - 1 + 2o')}ds, (8.8)

and this agrees with (3.34) for r > 1 - a .


The verification of the r < 1 - a case of (3.34) is elementary and we omit the
details.

IX. The Occurrence of Collapse for Large


In the previous sections it was shown that the H a m i l t o n i a n Hmc (1.4) under
consideration is stable if ~ is small enough. There are two parameters in the
problem, zg and ~. F o r stability of one electron and one nucleus it is necessary and
210 E.H. Lieb and H.-T. Yau

sufficient that za < 2In, but, assuming this condition, there is stability in the many-
body case ira < eo/q with ~0 > 1/47. In this section we shall prove that this stability
b o u n d is not just an artifact of our proof but that instability definitely occurs if ~ is
too large. Theorems 3 and 4 will be proved here.
Proof of Theorem 3. The method of proof here is the same as the method employed
in [23] to prove the instability of one-electron molecules in a magnetic field. Let
~bff L2(~ 3) be real with It4 l[2 = 1 and let ~ = (~b, [p[ qS) which is assumed to be finite.
Then
K
E=(~b, HNKd?)=~--zaI~2(x) Z Ix--R~l -ldx+zza 2 tRi-RjI-I'(9-1)
j=l l<=i<j<K

With ~b fixed let us try to position the R, so as to minimize the right side of (9.1).
This minimum (call it e) is less than any average of E over positions of the R,. In
K
particular, we use tp = I] ~(R~)z as a probability density for such an average. Then
j=l
A~(E) =, - ~ [z~/~ - ?~K(/~ - 1)/23 = • + ~ ~ {z:c, [ K - ½ - z - q : - ¼ ~ : ~ - ~ - z ~ } ,
(9.2)
where
o'= f q~(x)2q~(y)2 Ix--Yl- ldxdy. (9.3)
Now K can be chosen so that ] g - ½ - z - l l < ½ . Using this K, we have
e <=Av(E) <=z - ½a~. (9.4)
If we set ~ =2~/a, then when a > e l , e < 0 , and we can drive e to - ~ simply by
dilation, i.e. ~(x)--*23/24(;tx) and R~2RH2 with 2 ~ o o .
T o obtain a numerical value for el, choose q~(x)= re-1/2 e x p ( - r ) with r = Ixt.
The Fourier transforms of q~ and ~bz are
q~(p)= 8~1/2(1 + p 2 ) - 2 , ~'~(p) = 1 6 ( 4 + p 2 ) - 2 . (9.5)

Then
v=(2n)-3 ~~(p)21ptdp=8/3~, a=(2~)-3 ~~(p)(4~/lp]2)dp=5/8, (9.6)
and 2v/a = 128/15rc. []
Proof of Theorem 4. The method of proof here is similar to that used in [20] to
prove that the energy of N nonrelativistic bosons interacting with fixed nuclei via
Coulomb forces diverges as - N 5/3.Again, let q~e Lz(R 3) be real with tlq~II2 = 1 and
z = (~b,IPt ~b). Since there are q spin states, we can put N = q electrons into the state
q~. The energy is then
K 1
E=q~c--z°~q~2(x) ~ Ix-Rjt -ldx+z2~ Z IRi-Rit-x+ q(q-1)a
j=l l<i<j<=K
(9.7)

with tr given in (9.3). Let us first prove the theorem under the condition q/z > 1; at
the end of the proof we shall show how to handle the case q/z < 1.
Stability and Instability of Relativistic Matter 211

To construct ~b we first define g e L2(n:~3) by


g(x, y, z) = f(x) f(y) f(z), (9.8)
where f ~ L2(N 1) is given by f(x) = 3V~(1 -Ix[) for txl < 1 and f(x) = 0 for [xt > 1.
This f has flf l12= 1, and thus [Jg 112= 1. Let h e L20R 3) be some other function with
compact support and with (h, Ipl h) < oo and J]h IIz = 1. Define the integers n and K
and the positive number 2 by
n=[(q/z)l/3]> l, K = n 3,
2 = n3z/q = Kz/q, (9.9)

where [b] means integral part of b. Clearly, 1 ~ 2 ~ 1/8. Finally, we construct a


sequence of functions ~b°)(x), x ~ l R 3, by
(~(S)(X)2 -~--2g(X) -}- (I -- 4)S- 3 h(x/s -~ (0, 0, $2)) 2 . (9.10)
Now choose some fixed locations R1 .... , RK of K nuclei. Because of the scaling
ofh by s- 1 and translation by (0, 0, s2), we have that E converges to the following E'
as S--* oo :
K 1
E'=q2z-z°~2q~g2(x) 2 [x-ej] -ldx+z2~ 2 ]Ri--Rj] -1+ 22q(q--1) ac~,
j=l l <-i<j<-K
(9.11)
where z now means (g, ]pig) and a is given in (9.3) with g in place of 4~.
We claim that it is possible to choose the locations R 1..... Rr~ so that

~. ]Ri--Rjl-~--gygZ(x) ~ ] x - - R j l - l d x + K2tT<--K4/3/6. (9.12)


l<=i<j<=K j=l
If (9.12) holds then, recalling (9.9),
E' < q 2 z - -~zZc~(2q/z)4/3 . (9.13)
Recalling that 2 > 1/8 we have that E' < 0 whenever
~2~q >__8 (6~) 3 (~/2) 2 . (9.14)
We also have that z = (g, IP[g)< (g, p2g)l/2= 3 (by the Schwarz inequality). Thus,
collapse occurs if c~> ~2q- 1t - 2 with ~2 = (~/2) 2 8 (18) 3 -- t 15, 120, provided q/z > 1.
If, on the other hand, q/z<l and if ~>o~2q-lfl-2=o~2q-lz-2~-2(2/Tr,) 2, w e
have that (7c/2)2(zc~)3 >c~2z/q>c~z. Since ~2>>(2/rc)s, we are in the situation that
z~ > 2/~, which certainly entails collapse. Therefore, the theorem is proved for all
ratios q/z with the c~, given above.
There remains to prove (9.12). Choose n - 1 numbers fl~, .. ., /3,_ ~ satisfying - 1
- t o <ill < ... <ft,-1 < f l , ~ l such that
flj+ 1
f(x)Zdx=l/n for a l l j .
P1
Let Lj be the interval I t j_ 1, flj] in ]R 1 and, with m denoting a triplet (i,j, k), let F(m)
C R 3 be the rectangular parallelepiped L~ x Lj x Lk. Then, for each m,
gZ(x)dx = 1/n 3 = 1/K. (9.15)
r(m)
212 E.H. Lieb and H.-T. Yau

There are n 3 of these paralMepipeds. To prove (9.12) we shall place one of the R~'s
in each F(m)and average its location with respect to the density g2(x) restricted to
F(m). If the average satisfies (9.12) then there is surely some choice of the R~'s that
satisfies (9.12). Apart from a self energy contribution from each parallelepiped, the
average of the left side of (9.12) is zero. Thus the average of the left side is given by
the self energy terms

W = - ~I n6 ~ ~ S g(x)2g(y)2lx-Y[-ldxdy. (9.16)
F(m) x F(m)

Each integral is the self energy of a charge density g2 in F(m).However F(m)lies


inside a ball B(m) of radius r(m)=(sa+t2+u2)1/2, where 2s, 2t, and 2u are the
lengths ofF(m), namely (fii- fli- 1), ( f i j - fij- 1), (ilk -/~k- 1). The self energy is greater
than the minimum self energy of a charge 1/K distributed in B(m); the minimum
occurs for a uniform charge distribution on the boundary of B(m) and is r(m)- 1/K2.
Thus,
n
W<--~ ~=1 ~ ~ (s2+t2+u2) -1/2. (9.17)
j=1 k=l

NOW (S 2 + t 2 +u2)-1/2 >(S-t-t+u) -1. Substituting this latter expression in (9.17)


and then using the convexity of the function (s, t, u)~(s + t + u)- 1and recalling that
K = n 3, we have that
W>-½K(a+b+c) -1, (9.18)

where a, b, a n d c are the averages of s, t, and u. But a = b= c= 1/n, and thus (9.12) is
proved. []

Acknowledgements.The authors thank Michael Loss for helpful discussions and comments and
they thank Stefan Knabe for performing numerical calculations.

References

1. Baxter, J.R.: Inequalities for potentials of particle systems. IlL J. Math. 24, 645-652 (1980)
2. Chandrasekhar, S.: Phil. Mag. 11, 592 (1931);Astro. J. 74, 81 (1931); Monthly Notices Roy.
Astron. Soc. 91, 456 (1931); Rev. Mod. Phys. 56, 137 (1984)
3. Conlon, J.G.: The ground state energy of a classical gas. Commun. Math. Phys. 94, 439-458
(1984)
4. Conlon, J.G., Lieb, E.H., Yau, H.-T.: The N7/5law for charged bosons. Commun. Math. Phys.
116, 417-448 (1988)
5. Cycon, H.L., Froese, R.G., Kirsch, W., Simon, B.: Schr6dinger operators. Berlin, Heidelberg,
New York: Springer 1987
6. Daubechies, I.: An uncertainty principle for fermions with generalized kinetic energy.
Commun. Math. Phys. 90, 511-520 (t983)
7. Daubechies, I.: One electron molecules with relativistic kinetic energy: properties of the
discrete spectrum. Commun. Math. Phys. 94, 523-535 (1984)
8. Daubechies, I., Lieb, E.H.: One-electron relativistic molecules with Coulomb interaction.
Commun. Math. Phys. 90, 497-510 (1983)
9. Dyson, F.J.: Ground state energy of a finite system of charged particles. J. Math. Phys. 8,
1538-1545 (1967)
Stability and Instability of Relativistic Matter 213

10. Dyson, F.J., Lenard, A.: Stability of matter I and II. J. Math. Phys. 8, 423-434 (1967); ibid 9,
698-711 (1968). See also Lenard's Battelle lecture. In: Lecture Notes in Physics, vol. 23. Berlin,
Heidelberg, New York: Springer 1973
11. Erdelyi, A., Magnus, W., Oberhettinger, F., Tricomi, F.G.: Tables of integral transforms,
Vol. t. New York, Toronto, London: McGraw-Hill t954, p. 75, 2.4 (35)
12. Federbush, P.: A new approach to the stability of matter problem. II. J. Math. Phys. 16,
706-709 (1975)
13. Fefferman, C.: The N-body problem in quantum mechanics. Commun. Pure Appl. Math.
Suppl. 39, $67-S109 (1986)
14. Fefferman, C., de la Llave, R.: Relativistic stability of matter. I. Rev. Math. Iberoamericana 2,
119-215 (1986)
15. Fr/Shtich, J., Lieb, E.H., Loss, M.: Stability of Coulomb systems with magnetic fields. I. The
one-electron atom. Commun. Math. Phys. 104, 251-270 (1986)
16. Herbst, I.: Spectral theory of the operator (pZ+mZ)l/Z-ze2/r. Commun. Math. Phys. 53,
285-294 (1977); Errata ibid 55, 316 (1977)
17. Kato, T.: Perturbation theory for linear operators. Berlin, Heidelberg, New York: Springer
1966. See remark 5.12, p. 307
18. Kovalenko, V., Perelmuter, M., Semenov, Ya.: Schr6dinger operators with L~2(R *)potentials.
J. Math. Phys. 22, 1033-1044 (1981)
19. Lieb, E.H.: Stability of matter. Rev. Mod. Phys. 48, 553-569 (1976)
20. Lieb, E.H.: The N 5/3 law for bosons. Phys. Lett. 70A, 71--73 (1979)
21. Lieb, E.H.: Density functionals for Coulomb systems. Int. J. Quant. Chem. 24, 243-277 (1983)
22. Lieb, E.H.: On characteristic exponents in turbulence. Commun. Math. Phys. 92, 473-480
(1984)
23. Lieb, E.H., Loss, M.: Stability of Coulomb systems with magnetic fields. II. The many electron
atom and the one electron molecule. Commun. Math. Phys. 104, 271-282 (1986)
24. Lieb, E., Simon, B.: Thomas Fermi theory of atoms, molecules and solids. Adv. Math. 23,
22-116 (1977)
25. Lieb, E.H., Thirring, W.: Bound for the kinetic energy of fermions which proves the stability of
matter. Phys. Rev. Lett. 35, 687-689 (1975). Errata, ibid 35, 1116 (1975); see also their article:
Inequalities for the moments of the eigenvalues of the Schr6dinger Hamiltonian and their
relation to Sobolev inequalities. In: Studies in Mathematical Physics, Essays in honor of
Valentine Bargmann. Lieb, E.H., Simon, B., Wightman, A.S. (eds.). Princeton, NJ: Princeton
University Press 1976
26. Lieb, E.H., Thirring, W.: Gravitational collapse in quantum mechanics with relativistic kinetic
energy. Ann. Phys. (NY) 155, 494-512 (1984)
27. Lieb, E.H., Yau, H.-T.: The Chandrasekhar theory of stellar collapse as the limit of quantum
mechanics. Commun. Math. Phys. 112, 147-174 (1987). See also Lieb, E.H, and Yau, H.-T.: A
rigorous examination of the Chandrasekhar theory of stellar collapse. Astro. J. 323, 140-144
(1987)
28. Loss, M., Yau, H.-T.: Stability of Coulomb systems with magnetic fields. III. Zero energy
bound states of the Pauli operator. Commun. Math. Phys. 104, 283-290 (1986)
29. Weder, R.: Spectral analysis of pseudodifferential operators. J. Funct. Anal. 20, 319-337 (1975)

Communicated by A. Jaffe

Received May 12, 1988

You might also like