This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
Ilfoll
Let us now consider the complex case. We shall define two linear real valued functionals11 and f2 on Y by the equality
fo(x) =fi(x)+if2(x) Of course, IIf ll < Ilfoll and IIf2II < Ilfoll. Moreover, for x e Y,
f1(iX)+if2(iX) =fo(iX) = ifo(X) = ifi(x)-f2(x). ,(ix). Therefore fo(x) = fi(x)-if,(ix). Hence f2(x) = Let Fi(x) be a real-valued norm preserving extension fl(x) to the whole space X. Of course, by definition, IIFIII = Ilfill Letf(x) = Fi(x)-iF,(ix).
The functional f(x) is a continuous functional linear with respect to complex numbers (compare Corollary 4.1.3). Since F, is an extension off,, f is an extension of the functional fo. To complete the proof it is enough to show that I f II < Ilfoll
Let x be an arbitrary element of X. Let O = argf(x). Then If(x)I = If(e-iex)I = IF1(e-i1x)I < IIFiiflleiexII < Ilfi II IIxII < IIfthI IIxII Hence IIxII < Ilfoll.
COROLLARY 4.1.6. Let X be a normed space. Then
IIxII = sup
If(X) I
111111
where the supremum is taken over all continuous linear functionals f e X*. Proof. Let x0 be an arbitrary element of X. Let Xo be the space spanned
by x0, i.e. the space of all elements of the type axo. Let us put fo(x) = a I IxoII for x e Xo. The functional fo is of norm one. Basing ourselves on Theorem 4.1.5, we can extend it to a continuous linear function Fo of norm one. Then II xoll = fo(xo) = Fo(xo). Hence
sup I f(xo) I < I Ixol I = Fo(xo) 111111
< IUII41 sup f(xo)
Existence of Continuous Linear Functionals and Operators
193
Duren, Romberg and Shields (1969) gave an example of an F-space
X with a total family of continuous linear functionals such that X possesses a subspace N such that the quotient space X/N has trivial dual. Shapiro (1969) showed that 12', 0 < p < 1, also contain such subspaces N. Kalton (1978) showed that every separable non-locally convex F-space has this property. 4.2. EXISTENCE AND NON-EXISTENCE OF CONTINUOUS LINEAR FUNCTIONALS
THEOREM 4.2.1 (Rolewicz, 1959). Let
lim inf N(t) > 0. t t->ao Then in the space N(L(Q,E,u)) there are non-trivial continuous linear functionals. Proof. Corollaries 4.1.2 and 4.1.3 imply that it is enough to show that. in the space N(L(Q,E,p)) there is an open convex set U different from
the whole space. Let E e E, where µ(E) = a, 0 < a <+oo. Since lim inf N(t) > 0, there are a positive constant a and a positive number t-aoo
t
T such that, for t > T, N(t) > at. Let U be a convex hull of the set {x: pN(x) < 11. The set U is an open convex set. We shall show that it is different from the whole space N(L(S2,E,µ)). Let x1, ... , x.,, be arbitrary elements such that pN(xt) < 1, i = 1, 2, ..., n
Let Bs = {s: Ixs(s)l > T}. Let xi(s)
for s e BB,
elsewhere
I0
Let X0, =
X1'+ ... +x;,
o
n
and
and
x;' = xs-x;.
... xo" = xi + n
Since Ixa'(s)I < T(i = 1,2,..., n), Ix"(s)I
a,f I x:(s) I dic <.1 N(Ix E
E
Id/u
(i = 1, 2, ..., n).
Chapter 4
194
Hence f Jxo(s) Jdp
not belong to the set U.
If the measure p has an atom E0 of finite measure, then there is a non-trivial continuous linear functional in the space N(L(Q,E,p). Indeed, by the definition of measurable function, x(t) is constant on E0 p-almost everywhere, x(t) = c. Let us put f(x) = c. It is obvious that f(x) is a continuous linear functional.
If the measure p is atomless, the following theorem, converse to Theorem 4.2.1, holds : THEOREM 4.2.2 (Rolewicz, 1959). Let p be an atomless measure. If
lim inf N(t) = 0,
(4.2.1)
t
t- oo
then there are no non-trivial continuous linear functionals in the space
N(L(Q,E,p)) Proof. Let e be an arbitrary positive number. Suppose that (4.2.1) holds.
Then there is a sequence {tm} tending to infinity such that
Ntmm) -->O.
Let E be an
Let km be the smallest integer greater that
arbitrary set belonging to E of the finite measure. Since the measure p is atomless, there are measurable disjoint sets El, ..., Eke, such that km
E= U EE
and
,u(Ei) = (E)
(i = 1, 2, ..., km).
Let xm(s) =
{
tm
for s e EE (i = 1, 2, ..., km),
0
elsewhere.
Obviously, for sufficiently large m, pN(x;,) < e. On the other hand, for s e E km
ym(s) = km
x`m(s) = km Lam, ti=1
Existence of Continuous Linear Functionals and Operators
195
Since km moo, every function of the type aXE belongs to the convex hull U of the set {x: pr,(x) < e}. The set E is an arbitrary set of finite measure. Hence all simple functions with suports of finite measures belong to U. The set U is open and convex. Therefore U = N(L(Q,E,p)).
COROLLARY 4.2.3 (Day, 1940). In the spaces LP(Q,E,u), 0
Lq such that TI = f (here 1 denotes a constant function equal to one) and [Tx]q < [ f]q [x]p
for x e LP.
(4.6.9)
Proof. To begin with we shall prove the theorem for p = q. Suppose that inflf(t)I > 0. Let
f
0
f Ulq
Thus [fo]q = 1. Let T0(x) = x(F(t))fo(t ), where t
F(t) = f Ifo(t) qdt. 0
The operator TO is linear and I X(F(t))Iq dF = IIXIIq
1T'0(X)IIq = f I x(F(t))fo(t)j" dt = 0
0
It is easy to verify that TI = [f]gT01 = f(t). We recall that, in general, Lq D LP and, for y e LP, [y]q < [ylp Thus
[Tx]q < [X]q[f ]q < [X]p[f]q.
(4.6.9)
holds provided inf I f(t)I > 0. Since the functions f with this property are dense in Lq, by continuity arguments we find that (4.6.9) holds for all functions belonging to Lq.
Existence of Continuous Linear Functionals and Operators
213
Let C be a one-dimensional subspace of the space LP, 0LP/C be the quotient map associating with each x E LP the coset containing x. Of course by the definition of the quotient space IIp0fIl, = inf {Ilf--allP: a being scalar} where II II denotes the quotient norm. Since there is no danger of confusion, we shall denote II Ii; also by II IIP
The space LP/c, 0 < p < 1 is not isomorphic to the space LP (see Kalton and Peck, 1979); nevertheless, it embeds into LP, as follows from
LEMMA 4.6.3. There is a linear operator S : LP-*LP, 0 < p < 1, such that IIpofIIP < IIS(f)IIP < 2IIpoflIP
(4.6.11)
Proof. By the classical results of measure theory the space LP is isometric to the space LP([0,1] x [0, 1]). We define S : LP->.LP([0,1] x [0, 1]) by Sfl(x.y)
=f(x)-f(y)
Then
ii
i
IISf1IP = f f If(x)-fly) IPdxdy> f IlpofIIPdy = Ilpofllp 0
0
0
On the other hand, IISfIIP I
Observe that
S(f) = S(f-a) for all costant functions a. Thus by (4.6.12) IIS(f)IIP = inflJS(f--a)IIP < 2infllf-allp = 211pOf11p.
(4.6.12)
Chapter 4
214
LEMMA 4.6.4. Let {cn}, n = 0,1, 2, ..., be a sequence such that 00
1
c,112 < $
(4.6.13)
n=1
Then we may select sequences {pn}, {sn}, n = 0, 1, ..., and a sequence of finite-dimensional subspaces V. C L1" such that
en > 0, pn is increasing, po = 1/2, limp, = 1,
(4.6.14)
n---km
1eVn, nMnen < c,, n = 1, ...,
(4.6.15)
where n-1
M,dimV{,
(4.6.16)
:=o
.for each n > 0 there exists On, k:
1
k < r(n)} such that Vn,k a V,,
k = 1, ..., r(n) and r(n)
Vn,k = 1,
(4.6.17)
Ivn, k]p. < En,
(4.6.18)
[vn,k]pn+' < en"+'.
(4.6.19)
k=1 r(n)
k=1 r(n)
k=1
Proof. We select the sequences by induction. We begin by takingpo = 1/2, so = 2, Vo = C, Vo C L/1f2, r(0) = 1 and vo,1 = 1. Observe that (4.6.15)
and (4.6.17)-(4.6.19) hold. Condition (4.6.16) ought to be valid for n > 1. Suppose that (po, ..., Pm-1), (so, ..., Em-1), (Vo, ..., V,,,_1) are chosen so that (4.6.15)-(4.6.18) hold for n = m-1 and (4.6.19) holds
for n = m-2. Since r(m-1) [Vm_i,k]pm-i <em-1,
k=1
Existence of Continuous Linear Functionals and Operators
215
we can find pm sufficiently close to 1, such that pin > pm_1i p,n > 1-1/m
and (4.6.19) holds for n = m-1. Now take em such that (4.6.16) holds Since 1 belongs to the convex hull of every neighbourhood of zero in LP'", we can find vm,1, ... , Vm, r(m) E LP'" such that (4.6.18) holds for n = m. Finally, we put V. = lim (vm,1, , V M, r(m)). Let Z be a space of real valued functions f defined on (0, oo) such that co n±1
A(f)= J.f If(t)JP" dt < +oo. n=on
A is an F-norm on Z and (Z, A) is an F-space. It is easy to verify that the space Z is locally bounded and that the unit ball B = {x: A(x) < I}
is $ -convex. Thus there is a z -homogeneous norm
II
II such that
B = {x: IIxii < 1}. Write [x] = IIxii2
Let Z(a, b) be a subspace of Z supported on the interval (a, b). Let Pn, E, Qn be the natural projection of Z onto Z(0, n), Z(n, n+ 1) and Z(n, oo) respectively. Then
I = Pn+Qn = Pn+En+Qn+1 and
IIEnii = IIPnII = IlQnil = 1.
Note that if f, g Z(n, oo) then (4.6.20) [f+g]P" < [f]P"+[g]"" There is a natural isomorphism Tn: LP"->Z(n,n+l) given by
Tnf
- {f(t-n) 0
for n < t
Let Un = Tn Vn, en = Tn 1, en, k = Tn vn, k, n = 0,1, ... k = 1, 2, ... , r (n) co
Let Y = U Un, M = lin {en}. Let p be the natural quotient map Z->Z/M.a=1
p:
Thus
IIpfiI = inf Ilf--gll, pM
Chapter 4
216
where we denote the norm in the quotient space induced dy the norm II
II also byll
II.
Note that if f e Z(n, n+ 1), then Ilpfll = infllf-aenll = minll.f-aenHH. aeR
aER
LEMMA 4.6.5. Suppose that f e Z(0,n) and 11f 11 1. Then there exists a linear operator A: Z(n,n+l)->Z(0,n) such that A(en) =f and IIAII < 1. Proof. Suppose that f = ho+...+ hn_1, where hi c- Z(i,i+l), i = 0, 1, ... .... n-1. Since IIf II = 1 implies A (f) = 1, we have n-1
1 = IIAII = A(f) _
i=0
[hilp'
By Lemma 4.6.2 there exist linear operators Fi : Z(n, n+ 1)-ieZ(i, i+ 1) such that [Fi(f)] < [hi][f] and F{en = hi, i = 0,1, ...,
n-1.
Let A = Fo+...+ Fn-1. Then, of course, Aen =f
If g e Z(n,n+l) and III = 1 = A(g), then n-1
A(Ag) _
n-1
[hi]PI < 1.
[Fi(g)]P` < i=0
i=0
This implies that IIAII < 1.
Now let {B,b}, n = 0, 1, ..., be a partitioning of the set N of non-negative integers into infinite disjoint subsets with the property that
n < minB.. For each n, let {yk: k e Bn} be a dense subset of the set (f. f e Uo+... + Un, IIf I I = 1}, with the property that yk = en infinitely often. By Lemma 4.6.5 we can find operators Ak: 7,(k, k+1)-+Z(0, k) with IIAkII <1 so that Akek=yk, k = 1, 2, ...
Define T: Z-Z so that CO
T = Y ckAkEk k=o
with the convention AO = 0. By simple calculations we obtain co
IITII
1
k1/2 <
2
(4.6.21)
Existence of Continuous Linear Functionals and Operators
217
Moreover T(Z(O, k+ 1)) C Z(0, k)
(4.6.22)
T(Z(O, 1)) = 0.
(4.6.23)
and
By (4.6.21) the operators S = I-T is invertible.
Observe that T(M) C Y. Let M1 = S(M) C Y. Let it : Z->Z/M be the quotient map and let X be equal to i(Y) (isomorphic to Y/M1). We shall show that X is a rigid space. LEMMA 4.6.6. For f e Z(O,n+1). (4.6.24)
II PEnfll < 2IIitf11.
where the both norms in quotient spaces Z/M and Z/M1 are denoted in the same way by II II
Proof. Take an arbitrary number 6 > 1. By the definition of the norm in the quotient space there is a g e M such that IIf-S911 < IIitfli.
Then
IPEnf PEnSgII <6IIitfII By the definition of M, pEng = 0 for g e M. Thus by the definition of S
(4.6.25)
IPEnf+PEnTSII < 6IIitf1I.
By (4.6.22), En Tg = En TQn+1 g, and so IIPEnf II < oIIitfII+IIEnTQn+1gll (4.6.26)
<6II7rfIl+ I IIQn+1gII
Since Qn+l f = 0, Qn+l (f-Sg) = Qn+1 S g and IQn+1sgll < bJIitfJi.
Hence, by the definition of S, IQn+1gII < 6II7cf ll+IIQn+1 T9II
=
SIIitfII+IIQn+1TQn+1gll
Chapter 4
218
so that IIQn+1gII < 211icf11
Thus, by (4.6.26), IIPEnfII < 2611itfl1
Since 6 is an arbitrary number greater than 1, we obtain (4.6.24). LEMMA 4.6.7. The space X is infinite-dimensional.
Proof. By (4.6.16) en-->0, and by (4.6.17) and (4.6.18) dim V -moo. Thus For f e Un, by Lemma 4.6.6, dim
II7rf II > a lIPf11= $ minllf-aenll
Hence dim a(Un) > dim Un-1 and dimX = +oo. LEMMA 4.6.8. The set {an (en): a e R, n e N} is dense in the space X.
Proof. Suppose that f e UO+ ...+ Un and A (f) = 1 = I I f II. Then by the definition of {yk}, there is a subsequence % C N such that
limyj=f.
j By the definition of T, T(c;-lej) = yj, and so n (cj 1 ej) _ : (yj) -+ n U)
The multiples of f are dense in Y and this completes the proof. THEOREM 4.6.9. The space X is rigid.
Proof. Suppose that a linear continuous operator A maps X into itself and that IIAII < 1. We shall show that, for each n, 7r(en) is an eigenvector
of A. Fix n e N and let B. = {j e B.: yj=e1}.For jEB',Tej=cten. Hence 7r(e1) = cj n (en). Now r(j)
ej =
e,,k
k=1
and r(j)
[ej,k] < ej . k=1
Existence of Continuous Linear Functionals and Operators
219
Since IIAII < 1, there are gj,k E Y such that lt (gj,k) = Alt (ej,k) and
[gj,k] < [ej,k],
k = 1, 2, ... , r(j)
Let r(j) 1
hj =
gj,k. k=1
Then
ir(hj) = An(ej). Now
Pjgj,k E Uo+ ...
+ Uj_1,
and so, by Lemma 4.6.1 (for p = 2), r(j)
[Pjhj] <Mj
[Pjgj,k] <M5Ej
<-Cj-j
j
k=1
Similarly
Qj+lgj,ke Z(j+1, oo), and by (4.6.19) and (4.6.20) 1
r(j)
[Qj+lhj]
C
' [g3,]1+'; p1+ < Ej k=1
Thus
[hj-Ejhj] _ [Pjhj+Qj+lhj] <
(4.6.27)
This implies that, for j e B,,,
[An(en)-cj 1rEjhj] <
4
(4.6.28)
By (4.6.28) we trivially obtain for i, j e B', i <J,
[ci 1nEjhj-c; 1nEjhj] < 8(1 + i
Observe that
1I.
j
c, 17rEjhj-c: 1nE{h{e Z(0,j+1), and by Lemma 4.6.6 [c j
1PEj
hjl < 32
(i + j )
Chapter 4
220
Thus, by the definition of p, there is a 2 = 1(i, j) such that
[cc 1EjhJ-2ej] < 32 (1 + i
1 ).
j
This implies
[c; 17r(Ejhj-2ej)] < 32
+
Since, by the definition of B. c,-'Tt(ej) _ ir(en), [Air(en) - It (en)] < 64
+
As i, j e B. can be chosen arbitrarily large, we deduce that there is a real p such that Arc (en) = plc (en).
Thus, by Lemma 4.6.8, there is a dense set of eigenvectors. Using the continuity of A, we can deduce that each element x E X is an eigen-
vector corresponding to an eigenvalue A, Taking two eigenvectors x, y, we infer, by simple considerations in the two-dimensional space lin({x,y}), that x+y is an eigenvector if an only if the eigenvalues are
equal 2z=Ay= 1.Thus A=AI. Kalton and Roberts (1981) proved also that the constructed rigid space is isomorphic to a subspace of LP for all 0 < p < 1. Refining the construction described above, they constructed a rigid space such that each quotient is rigid, and formulated the following problems :
Problem 4.6.11 (Kalton and Roberts, 1981). Does LP(0
Chapter 5
Weak Topologies
5.1. CONVEX SETS AND LOCALLY CONVEX TOPOLOGICAL SPACES
Let X be a linear space over real or complex numbers. A set A C X is said to be convex if, for arbitrary non-negative, a, b such that a+b = 1, x, y e A implies ax+by c- A (see Section 3.1). The intersection of an arbitrary family of convex sets is a convex set. Let A be a convex set and let a1, ..., an be non-negative numbers such
that
at= 1. i=1
Then n 1
ajxt e A
i=1
for an arbitrary system of elements x1, ..., xn c- A. Let A be an arbitrary set. By conv(A) we denote the intersection of all convex sets containing A. The set conv(A) is called the convex hull
of the set A. It is easy to verify that the convex hull of the set A may be characterized in the following way : in
in
anxn: xn E A, an > 0, 1 an = 1 j.
conv(A) n=1
(5.1.1)
n=1
The algebraic sum of two convex sets is a convex set. The image of a convex set A under a linear operator T is a convex set. A counterimage of a convex set under a linear operator is a convex set.
Let M be an arbitrary subset of X. We say that a point p c- M is 221
Chapter 5
222
a C-internal point of the set M if, for each x e X, there is a positive number a such that p+tx e M for It I < e. A point p e X is called a C-bounding point of the set M if it is neither a C-internal point of the set M nor a C-internal point of the complement of M.
Let K be a convex set containing 0 as a C-internal point. Then we can define a functional p (x) (generally non-linear) in the following way :
p(x)=inf{t>0: t eK}.
(5.1.2)
Let us remark that the functional p(x) has the following properties (compare Section 4.1) : (1) p (x) > 0,
(2) p(x) <+oo for x e lin K (3) p(x) = tp(x) for positive t, (4) p (x+y) < p (x) +p (y), (5) if xe K, then p(x) 1, (6) if x is a C-internal point of the set K, then p (x) < 1, (7) if x is a C-bounding point of K, then p (x) = 1.
We say that a linear functional f(x) separates two sets M and N if there is a constant c such that Re f(x) > c for x e M and
Re f(x) < c
for x e N. Here by Re z we denote the real part of a complex number z. Obviously, if X is a linear space over reals, then Re f(x) = f(x). Of course, a functional f(x) separates the sets M and N if and only if it separates the sets M - N and {0}. PROPOSITION 5.1.1. Let M and N be two disjoint convex sets in a linear
space X. Suppose that M has a C-internal point. Then there is a linear functionalf(x) separating the sets M and N. Proof. Without loss of generality we can assume that 0 is a C-internal point of the set M. To begin with, we shall consider the case where X is a linear space
over reals. Let -y be an arbitrary C-internal point of the set M-N.
Weak Topologies in Banach Spaces
223
Since the sets M and N are disjoint, the point 0 does not belong to the
set M-N, Let K = M-N+y Obviously, 0 is a C-internal point of the set K and the point +y does not belong to K. Let p(x) be the functional defined by formula (5.1.2). Obviously, p(+y) > 1. Let X0 be the space spanned by the element y. We define on the space X0 a linear functionalfo in the following way fo(ay) = ap(y). Of course, for x e Xo, .fo(x) < P (x)
By the Hahn-Banach theorem (Theorem 4.1.1) the functional fo(x) can be extended to the functional f(x) defined on the whole space X such thatf(x) < p (x). This implies thatf(x) < 1 for x e K and fly) > 1. Hence the functional f separates the sets M-N and {0}. Therefore, it separates the sets M and N. Now we shall consider the complex case, i.e. the case where X is a linear space over complex numbers. The space X may obviously be considered as a linear space over reals, too. Then there is a real-valued
linear functional f(x) separating the sets M and N. Let g(x) = f(x) -if(ix) (cf. Corollary 4.1.3). The functional g(x) is linear with respect to complex numbers and it separates the sets M and N. We recall that X is a linear topological space if it is a linear space with
a topology and if the operations of addition and multiplication by scalars are continuous.
Let A be a subset of a linear topological space X, If A is a convex set, then the closure A of the set A is also a convex set and the interior Int(A) of the set A is also convex. The continuity of multiplication by scalars implies that each internal point of the set A is a C-internal point of this set. If a convex set A contains internal points, then the necessary
and sufficient condition for the point x e A to be C-internal is that x be an internal point. By conv(A) we denote the intersection of all closed convex sets containing A. The set conv (A) is called the closed convex hull of the set A. It is easy to verify that (5.1.3) conv (A) = conv(A), (5.1.4) conv (aA) = a conv (A) for all scalars a.
Chapter 5
224
PROPOSITION 5.1.2. If cony (A) is a compact set, then
cony (A +B) = cony (A)+ conv (B).
The above formula is a consequence of the following lemma : LEMMA 5.1.3 (Leray, 1950). Suppose we are given a topological space Y, a compact space K, and a continuous mapping f(t, k) of the product X x K into a topological space X. Let F be a closed subset of the space X disjoint with the set f(to, K). Then there is a neighbourhood V of the point to such that the set F does not have common points with the set f(V,K).
Proof. Let k e K. The continuity of the function f implies that there are a neighbourhood V(k) of the point to and a neighbourhood W(k) of the point k such that Fnf(V(k), W(k)) = ¢. The set K is compact, hence we can choose a finite cover of the set K by the sets W(kl), ..., m
W(km), Kc U W(ki). Let z=i
m
n V(k{). v= %=I The set V is an open set with the required properties. LEMMA 5.1.4 (Leray, 1950). Let X be a linear topological space. Let F be
a closed set in X and let K be a compact set in X. Then the set F+K is closed.
Proof. If x 0 F+K, then the set F does not have common points with the set x-K. Therefore, by Lemma 5.1.3, there is a neighbourhood V of the point x such that Fn (V-K) Hence
vn(F+K) =-0.
El
The proof of Proposition 5.1.2 is a trivial consequence of Lemma 5.1.4 and of the fact that the algebraic sum of two convex sets is convex.
Let us note some further properties of convex sets. By similar considerations to those in Corollary 4.1.2 we can prove that if a functional f(x) separates two sets M and N and one of them contains an internal point, then the functional f(x) is continuous.
Weak Topologies in Banach Spaces
225
Conversely, Proposition 5.1.1 implies that if M and N are two disjoint
convex sets and one of them contains an internal point, then the sets M and N can be separated by a continuous linear functional f(x). We say that a linear topological space is locally convex if each neighbourhood of zero contains a convex neighbourhood of zero (compare Section 3.1).
THEOREM 5.1.5. Let X be a locally convex topological space. Let M and N be two disjoint convex closed sets. If, in addition, M is compact, then
there are constant c, and e > 0 and a continuous linear functional f(x) defined on the space X such that
Ref(x) < c-e
for xe N,
Ref (x) > c
for x e M.
(5.1.6)
Proof. Lemma 5.1.4 implies that the set M-N is closed. Since M and N are disjoint, the set M-N does not contain the point 0. The space X is locally convex, and therefore there is a convex balanced neighbourhood
of zero U such that Un (M-N) = 0. Proposition 5.1.1 implies that there is a continuous linear functional f(x) which separates the sets U and M-N. The set U is open, whence
sup{Ref(x) : x e U} = e > 0. Therefore, Ref(x)
(5.1.7)
e for x e M-N. This trivially implies the prop-
osition. COROLLARY 5.1.6. Suppose that in a linear space X there are two locally convex topologies (X, Tl) and (X, a2). If both topologies designate the same
continuous linear functionals, then a convex set A is closed in the first topology Tl if and only if it is closed in the second topology T2. Proof Let K be a convex set closed in the topology r1. Then for each point p 0 K there is a linear functional f(x) which is continuous in the
topology Tl and there are numbers c,e > 0, such that Ref(x) < c-e for x e K and Ref(p) > c. But the hypothesis implies that the functional f(x) is continuous in the topology T2. Hence p could not belong to the closure K2 of the set K in the topology T2. This implies that K2 = K.
Chapter 5
226
If K is a convex set closed in the topology r2, then using the same arguments we infer that it is closed in the topology r l.
5.2. WEAK TOPOLOGIES. BASIC PROPERTIES
Let X be a linear space over real or complex numbers. Let X' denote the set of all linear functionals defined on the space X. A subset T of
the set X' is called total if f(x) = 0 for all f e T implies that x = 0 (compare Section 4.2).
Let T be a total linear set of functionals. By the f-topology of the space X we shall mean a topology determined by the neighbourhoods of the type N(p,.fi,... , fx ; ar, ... , ax) = {x: l fc(x-p)I < a; (i = 1, 2,
..., k)},
where at > 0 and f¢ e T. Obviously, the space X with a f-topology is a locally convex space. We say that a set A C X is f-closed (f-compact) if it is closed (compact) in the f-topology. The closure of a set A in the f-topology will be called the f-closure. A functional f(x) is said to be T-continuous if it is continuous in the I'-topology. Let X be a locally convex topological space. Let X* be the set of all
continuous linear functionals defined on X (conjugate space). The X*-topology is called the weak topology. Obviously, the weak topology is not stronger than the original one. Let X be the space conjugate to a locally convex space X_. Let us recall that in the conjugate space we have the topology of bounded convergence. Each element x_ E X_ induces a continuous linear functional F on the space X by the formula
F(x) = x(x_).
(5.2.1)
We shall indentify the set of functionals defined by formula (5.2.1) with the space X_. The X_-topology in X is called the weak topology of
functionals or the weak-*-topology. Since we always have X* ) X_, the weak topology of functionals is not stronger than the weak topology.
Weak Topologies in Banach Spaces
227
PROPOSITION 5.2.1. Let X be a linear space. Let T be a total linear set of functionals. A linear functional f(x) is continuous in the T-topology if and only if f e F.
The proof is based on the following lemma : LEMMA 5.2.2. Let X be a linear space. Let g, fl, ... , fn be linear functionals defined on X. If
fi(x)=0,
i= 1,2,...,n,
implies
g(x) = 0, then g(x) is a linear combination of the functionals fl, ... , fn.
Proof. Without loss of generality we can assume that the functionals ... , f, are linearly independent. Let
Xo={xeX: f{(x)=0,i=1,2,...,n}.
(5.2.2)
Let X be the quotient space X/Xo. The functionals fl, ..., f, induce linearly independent functionals f1, ..., fn on X. The assumption about g(x) implies that the functional j (x) also induces a linear functional (x) defined on X.
Then space X is n-dimensional, therefore, j (x) is a linear combination off , ..., fz,
g=aifl+...+anfn. It is easy to verify that g = aifi+ ... +anf . Proof of Proposition 5.2.1. Sufficiency. From the definition of neighbourhoods in the f-topology it trivially follows that each functional f E f is f-continuous (i.e. continuous in the f-topology). Necessity. Let g(x) :f- 0 be a functional continuous in the f-topology. Then there is a neighbourhood of zero U in that topology such that sup Ig(x)I < 1. But the neighbourhood U is of the type
U={xeX: If(x)I
Chapter 5
228
This implies that g(x) = 0 for x c X0. Therefore, by Lemma 5.2.2, it follows that g(x) is a linear combination of f,, ..., f,, Since I' is linear,
gE.r. PROPOSITION 5.2.3. Let X be a linear space and c(x) a non-negative valued function defined on X. Let X' be the set of all functionals defined on X. Let K = {fE X': Jf(x)J < c(x)J. Then the set K is compact in the X-topology of the space X'. Proof. Let I(x) = {a: a scalar, lal < c(x)}
and
I =
\\ I(x). XEX
It is well known (the Tichonov theorem) that the set I is compact in the product topology. We define a mapping T of the set K into the set I by the formula
T.f= XEX /\.f(x).
(5.2.3)
It is easy to verify that T is a homeomorphism. In order to finish the proof it is sufficient to show that the image T(K) of the set K is closed.
Obviously, I may be considered as a set of functions defined on X Let
A(x, y) = {gE I: g(x+y) = g(x)+g(y)}
(5.2.4)
B (a, x) = {g e I: g (ax) = ag (x) 1.
(5.2.5)
and
The sets A(x,y) and B(a,x) are closed. Hence the set
T(K) = i
n I
1
X,ycX
A(x, y)J n [XEX n B(a, x)] a scalars
is also closed.
THEOREM 5.2.4 (Alaoglu, 1940). The closed unit ball of the space X conjugate to a Banach space X is compact in the weak-*-topology. Proof. By definition, the closed unit ball of the space X is the set
S* = {.fEX': l.f(x)l <1J4} and by Proposition 5.2.3 we trivially obtain the theorem.
Ll.
Weak Topologies in Banach Spaces
229
COROLLARY 5.2.5. A set A of the space X* conjugate to a Banach space
X is compact in the weak-*-topology if and only if it is closed in that topology and bounded in the norm topology.
Let X be a Banach space. Let X** denote the space conjugate to the space X*. The space X** is called the second conjugate. As we have seen
before, each element x e X can be regarded as a continuous linear functional on the space X* (see formula (5.2.1)). This means that there is a natural embedding n(X) of the space X into the space X**. Corollary 4.1.6 implies that the embedding n is norm preserving, i.e. that IIn(x)II = = IIxII
THEOREM 5.2.6 (Goldstine, 1938). Let X be a Banach space and let x** be its second conjugate. Let S and S** denote the unit balls in the spaces X and X** respectively. Then the set n(S) is dense in the set S** in the X *-topology.
Proof. By Sl we denote the X*-closure of the set n(S). Theorem 5.2.4 implies that S** is closed in the X*-topology, thus Sl C S**. Since Sl is a closure of a convex set, it is a convex set. We shall show that Sl = S**.
Suppose that this does not hold. Then there is an element X** e S** which does not belong to S1. Applying Theorem 5.1.5, we can find an X*-continuous functional f(x), a constant c and a positive number E such that _
Ref(x)
forxeS1.
(5.2.6)
Ref(x**) > c+E.
Since the functional f(x) is X*-continuous, by Proposition 5.2.1 we find that there is an element x* e X* such that f(x) = i(x*) for all z e X**.
Since n(S) C S,, Rex*(x) < c for x e S. The set S is balanced, which implies Ix*(x)I < c for all x e S. Therefore Ix*I1 < c. Hence
If(x**)l = Ix**(x*)I
c
and we obtain a contradiction of formula (5.2.6).
O
As an obvious consequence of Theorem 5.2.6 we find that the set n(X) is dense in X** in the X*-topology. A Banach space X is said to be reflexive if n(X) = X**.
Chapter 5
230
THEOREM 5.2.7. A Banach space X is reflexive if and only if the closed unit ball is compact in the weak topology (i.e. the X*-topology). Proof. Necessity. If X is a reflexive space, then n(S) = S** and (X**)*
= X*. Therefore the weak topology in S** is equivalent to the X*-topology. Hence, by the Alaoglu theorem (Theorem 5.2.4) S** is compact
in the X*-topology. Since S = S**, the set S is compact in the weak topology. Sufficiency. Let us now suppose that the unit ball S is weakly compact.
The natural embedding n is obviously a homeomorphism between S and n(S) in the X*-topology of both sets. Therefore n(S) is compact, and hence closed, in the X*-topology. Theorem 5.2.6 states that n(S) is dense in S in the X*-topology. Therefore, n(S) = S** and n(X) = X**. COROLLARY 5.2.8. A Banach space X is reflexive if and only if each bounded
weakly closed set is weakly compact. PROPOSITION 5.2.9. Let X be a reflexive Banach space. Let A be a bounded
closed convex set in X. Then the set A is weakly compact. Proof. Let p 0 A. By Theorem 5.1.5 there are a continuous linear func-
tional f(x), a constant c, and a positive number s such that
Ref(p) < c-s
and
Ref(x) > c for x e .A.
This implies that the point p does not belong to the weak closure of the set A. Therefore, the set A is weakly closed and, by Corollary 5.2.8, it is weakly compact.
5.3. WEAK CONVERGENCE
Let X be a Banach space. We say that a sequence {xn} of elements of X is weakly convergent to x e X if, for each continuous linear functional f(X)1
lim f(xn) =f(x). n->ao
Weak Topologies in Banach Spaces
231
THEOREM 5.3.1 (Eberlein, 1947; 9mulian, 1940). Let A be a subset of a Banach space X. Then the following three conditions are equivalent: A. The weak closure of the set A is weakly compact. B. Each sequence of elements of A contains a subsequence weakly
convergent to an element of X.
C. Each countable subset AO of the set A has a cluster point xo e X in the weak topology (this means that, for each neighbourhood of zero
U, A0n(xo+U) =0). The proof included here was given by Whitley (1967) and it is based on the following lemmas : LEMMA 5.3.2. Let X be a Banach space and let the conjugate space X contain a total sequence {x,*,} of functionals. Then the weak topology on a weakly compact subset of the space X is metrizable.
Proof. Without loss of generality we may assume that IIxnIi = 1, n = 1, 2, ... Define a metric for X by 00
d(x,y)_xn(x-Y)I.
n=1
2n
Let A be a weakly compact set. The identity mapping of A with the weak topology onto A with the metric topology induced by d is clearly continuous and thus it is a homeomorphism, since A is weakly compact.
LEMMA 5.3.3. Let F be a finite-dimensional subspace of the space Y* conjugate to a Banach space Y. Then there is a finite system of elements
y,,..., y. c- Y such that for all y*aF max 11Y*(Yf)JI
IIY*II
Proof. The surface of the unit sphere of the space F is compact. Therefore, there is a -L-net on this surface, i.e., there is a system of points Yi , , y C F such that, for any element y* e F of norm one, inf 11y*-y; J1 < 4. Let y,..... ym be elements of Y, each of norm one, 1LGi_<m
Chapter 5
232
such that ly, (yi)I > 4. Then
max ly*(yi)I = max jy*(yt)-y+ (yi)+yi (yi)j
1
1
4 - inf I.y*(yi)-Y*(yi)I > a - q = a 1 B. Let {an} be a sequence of elements of
A. Let X0 denote the space spanned by {an}. The set Xo is weakly closed as a linear subspace. Therefore, the intersection of the weak closure w (A) of the set A with the space X0 is a weakly compact set in X. By Lemma 5.3.2 the weak topology on w (A) n Xo is metrizable, because the space X0 is separable. Therefore, there is a subsequence {ank} weakly convergent to an element a e X0, i.e.
lim f(ank) =f(a)
(5.3.1)
k-*oo
for all f e Xo . Hence this is also true for f e X*. B->C. This is clear.
C-->A. Suppose that a set A satisfies condition C. Then, for each continuous linear functional f(x), the set of scalars f(A) = {f(x): x e A} also satisfies this condition. Therefore it is bounded. Hence the BanachSteinhaus theorem implies that the set A is also bounded. Let.n denote the natural embedding X into X**. Let w*(n(A)) be the closure of the .set n (A) in the X*-topology. Now we shall show
w*(n(A)) C n(X).
(5.3.2)
Let X** be an arbitrary element of the set w*(n(A)). Let xi be an .arbitrary element of the space X* of norm one. Since x** belongs to the set w*(n(A)), there is a point a1 e A such that
I(x**-n(a1))(xi)I < 1.
(5.3.3)
The space spanned by x** and x**-n(a) is two-dimensional. Hence by Lemma 5.3.3 there are points x2*, ..., xn (2) each of norm one, x; a X*,
Weak Topologies in Banach Spaces
233
i = 2, 3, ..., n(2), such that max IY**(x*)I > z IIY**II
2
(5.3.4)
n(2)
for all y** belonging to the space spanned by x** and x**-n(a,). Again using the fact that x** is in w*(n(A)), we can find a point a2 e A such that I(x**-n(a2))(xm)I
max
1<m
Then find X. (2)+11 ..., X,*, (3) of norm one in X* so that
max n(2)<m_
Iy**(x*)I > z IIY**II
for all y** belonging to the space spanned by the elements x**, n(a1), n(a2). Once more using the fact that x** is in w*(n (A)), choose a2 e A so that max I (x**-n(a3)) (x,*n)I < 1/3 and continue the construction. 1_<m_
Then we obtain : (1) an increasing sequence of positive integers {n(k)}, (2) a sequence of elements X, {x1,x2, ...}, (3) a sequence of elements of A, {a1ia2, ...} such that max 1<_m<_n(k)
IY**(xm*)I > 1IIY**II
(5.3.5)
for ally** belonging to the space Xk spanned by the elements x, n(a1), ...
..., n(ak_). And max 1-<m-
Ix**-n(ak))(xm)l < k .
(5.3.6)
Let X0 be the space spanned by the elements x, n(a1), ... By hypothesis there is a point x0 e X which is a cluster of the sequence in the weak topology of X. Since X0 is weakly closed as a closed subspace of X, x0 e X. Formula (5.3.5) implies that max Iy**(xm)I > 11y"11 for y e Xo. M
z
In particular, for x**-n(xo) we have max I (x**-n(xo)) (xn*,)I > 2IIx**-n(xo)II On the other hand, formula (5.3.6) implies that for m < n(p) < k I(x**-n(ak))(xm)I < 1/p.
Chapter 5
234
Thus I(X**-n(XO)) (Xm)I < I
P
(x**-n(ak)) (x*n)I + I x,* (ak-xo) I
+I Xm(ak-xo)I
(5.3.7)
The point x0 is a cluster point of the sequence {an} in the weak topology. Therefore, there is a number k such that
Ixm(ak-x)I < p
for m = 1, 2,..., n(p).
Obviously, we can assume that k > n(p). Therefore, (5.3.6) and (5.3.7) imply that for all p IIx**-n(xo)II
(5.3.8)
Hence x** = n(xo). This means that w*(n(A)) C n(X). Since n is a homeomorphism between X and n(X), both with X *-topology, the weak closure of the set A is exactly w*(n(A)). The Alaoglu theorem (Theorem 5.2.4) implies that the set w*(n(A)) is compact.
5.4. EXAMPLE OF AN INFINITE-DIMENSIONAL BANACH SPACE WHICH IS NOT ISOMORPHIC TO ITS SQUARE
In the majority of known examples of infinite-dimensional Banach spaces, those spaces are isomorphic to their Cartesian squares. Now we shall give an example which shows that this is not true in general. The example is based on the following LEMMA 5.4.1 (James, 1951). Let X be Banach space with a basis {Xn}. Let X. denote the space spanned on the elements en+l, en+2, ... If for any functional f belonging to X*
lim if Ix 11 = 0,
(5.4.1)
n->co
where f l y denotes the restriction of the functional f to a subspace Y, then the basis functionals {fn} (see Corollary 2.5.3) constitute a basis in X*.
Weak Topologies in Banach Spaces
235
Proof. Let f e X*. Let x e X, x = 2, fa(z)e.. Then n=1
m
O
0
f(x) = ffn(x)en) = .fn(x)f(en) = n=1
n=1
f L
f n=1
l f(en)f.](x).
00
f(en)fn is convergent to f in
Formula (5.4.1) implies that the series n=1
the norm of the functionals and that this expansion is unique. Example 5.4.2 (James, 1951)
Let x = {x1,x2, ...} be a sequence of real numbers. Let us write n
IIxII = sup
(xp=i_1-xp,t)2+(xp2n+1)2]1/2,
ti=1
where the supremum is taken over all positive integers n and finite increasing sequences of positive integers pl, ..., Pen+1 Let B be a Banach space of all x such that IIxII is finite and
lim xn = 0.
n-*,o
IIxII is a norm. Indeed, IIxII = 0 if and only if x = 0, IItxjl = It! IIxII for all scalars t. Now we shall show the triangle inequality. Let x = {x1, x2, ... }, y = {y1, y2, ... }. From the definition of the norm IIx+yiI it follows that for any positive a there is an increasing sequence of indices p1, ... , P2n+1 such that n
IIx+YII <
(xpzi_1+Yp2i-1-xpz,-Ypzi)2+(xpzn+l+ ti=1
71/2+e
+Ypzn+1)2
n 1/2
(xp2i-l-xps1)2+(xpzn+l)2,
[ i=1
(ypai-l-Yp9i)2+(Ypzn+1)211/2+E
+
ti=1
Thus the arbitrariness of a implies that IIx+YII < IIxII+IIYII
Let zn = {0, ..., 0, 1, 0, ...}. n-th place
< IIxII+IIYII+e
Chapter 5
236
It is easy to verify that the linear combinations of the elements zn are dense in the whole space B, because lim xn = 0. Moreover, for all positive integers n and p, n+p
n
aiziJI < 11 V atz= i=1
t=1
Then Theorem 3.2.15 implies that {zn}is a basis in B. We shall show now that the space B is not reflexive. For this purpose
we shall prove that the closed unit ball in B is not weakly compact. Let yn = zl+ ... +zn. Of course, IIy,II = 1. If the closed unit ball is weakly compact, then {y.} converges to a y e B. Since {zn} is a basis, y ought to be of the form (1, 1, ...). This is impossible, because, for all
xeB,limx,=0. co
Now we shall describe all functionals f belonging to the second conjugate space B**. Let {gn} be the basis functionals with respect to the basis {en}. According to Lemma 5.4.1, {gn} constitute a basis in the conjugate space B*. Let F be a functional from B**. Then the functional F is of the following form : there is a sequence of real numbers 00
00
{Fi} such that F(f) = j' Fi fi for any f e B, f = T' f gi. Let us calculate the norm of the functional F. We have n
n
i=1
i=1
= IIf!I sup[
f n
Fif{I =If(f Ftzi) <11fil I
Fizi
i=1
i=1
where the supremum is taken over all positive integers n' and increasing sequences of indices p1, ..., pen'+1 with Fpk replaced by 0 if pk > n. The arbitrariness of n implies n
IIFII < sup [ i=1
(Fptt_1_Fp2t)2+(Fp2n'+1)211/2. (5.4.2)
Weak Topologies in Banach Spaces
237
Let us now fix n and let un = S' Fizz. Let us define a linear functional i=1
f on the space Y. spanned by the elements un, zn+1, zn+2, ... in such a way that f(aun) = IlunIl
f(zi) = 0
and
for i = n+1, n+2,...
Then 00
00
f (aun+ Y a;zi)= Ilaunll < Ilunn+ f aizi i=n+1
Thus IIFII = 1. Let j=
i=n+1
figi be an extension of the functional f to
whole space B of norm one. Then fi = 0 for i > n and n
oo
F(f)=
i=1
Fifi' =I fFifi i=1
= I f(un)I = IlunIl < IIFII
Hence, calculating the norm of un, we obtain O
IIFII i f L
i=1
`Fp, 1-Fp, )2+(FPan+1)2]1/2
(5.4.3)
for all positive integers n and all finite increasing sequences of integers P1, , p2,+1 Combining (5.4.2) and (5.4.3), we obtain IIFII = sup[
i=1
(5.4.4)
(FP,i_1_FP:J2+(FPan+)2]1/2,
where the supremum is taken over all positive integers n and finite increasing sequences p1, ..., p2,+1. The norm IIFII is finite if and only if
there is a limit lim F. Since the space B is not reflexive, B** contains n=oo
an element which does not belong to n(B). Then the only possibility is B**
that n (B) is a subspace of codimension 1, i.e. that dim n(B) = 1. There(BXB)** X**
fore, dim n(B X B) = 2, and since dim is an invariant of an T, -(X-) isomorphism, the space B is not isomorphic to its Cartesian square (see Bessaga and Pelczynski, 1960b).
Chapter 5
238
Pelczyliski and Semadeni (1960) showed another example of a space
which is not isomorphic to its square. Their example is of the type C(SC). An example of a reflexive Banach space non-isomorphic to its square was given by Figiel (1972).
Problem 5.4.3. Does there exist a Banach space non-isomorphic to its Cartesian product by the real line ? The answer is positive for locally convex spaces, as will be shown in Corollary 6.6.12. Rolewicz (1971) gave an example of normed (non-complete) space X non-isomorphic to its product by the real line. Dubinsky (1971) proved
that each Bo-space contains a linear subset X which is not isomorphic to its product by the real line. Bessaga (1981) gave an example of a normed space which is not Lipschitz homeomorphic to its product by the real line. 5.5. EXTREME POINTS
Let X be a linear space over the real or the complex numbers. Let K be an arbitrary subset of X. We say that a point k e K is an extreme point of the set K if there are no two points k,, k2 e K and no real number
a, 0 < a < 1 such that k = ak,-{-(1-a)k2.
(5.5.1)
The set of all extreme points belonging to K is denoted by E(K). A subset A of the set K is called an extreme subset if, for each k e A the existence of k,, k2, 0 < a < 1 satisfying (5.5.1) implies that k,, k2 a A. PROPOSITION 5.5.1. Let X be a locally convex topological space. Let K be a compact set in X. Then the set E(K) is non-void.
Proof. Let 2C be a family of extreme closed subsets of the set K. We can
partially order this family in the following way : we say that A - B if B D A. Since the set K is compact, the intersection of a decreasing family of closed sets is a closed non-void set, and obviously it is also an extreme set, provided the members of the family belong to 2I.
Weak Topologies in Banach Spaces
239
Then, by the Kuratowski-Zorn lemma, there is a minimal element A0 of the family W.
We shall show that the set A0 contains only one point. Indeed, let us suppose that there are two different points p, q e A0. Then there is a functional x* e X* such that Rex*(p) Rex*(q). (5.5.2) Let
A, = {x e Ao: Rex*(x) = inf Rex*(y)}.
(5.5.3)
UEAo
Since the set Ao is compact, the set Al is not empty, Moreover, formula (5.5.2) implies that the set Al is a proper part of the set A0. Let k,, k2 be points of K such that there is an a, 0 < a < 1, such that ak,+(1-a)k2 a Al. (5.5.4)
Since Ao is an extreme subset, k, and k2 belong to A0. Since (5.5.4) a Re x*(k,) + (1 - a) Rex*(k2) = inf Rex*(y). yEAo
This is possible if and only if Rex*(kl) = Rex*(k2) = inf Rex*(y). HEA,
This implies that k,,k2 e Al. Hence A, is an extreme set. Thus we obtain
a contradiction, because A0 is a minimal extreme subset. Therefore, A. is a one-point set, A0 = {x0} and, from the definition, x0 is an extreme point. THEOREM 5.5.2 (Krein and Milman, 1940). Let X be a locally convex topo-
logical space. Let K be a compact set in X. Then
cony E(K) D K. (5.5.5) Proof. Suppose that (5.5.5) does not hold. This means that there is an element k c K such that k 0 conv E(K). Then there are a continuous linear functional x* and a constant c and positive a such that Rex*(k) < c (5.5.6) and
Rex*(x) > c+e
for xe conv E(K).
(5.5.7)
K1= {x e K: Rex*(x) = inf Rex*(y)}.
(5.5.8)
Let yew
Chapter 5
240
Since the set K is compact, the set K1 is not empty. By a similar argument
to that used in the proof of Proposition 5.5.1, we can show that K1 is an extreme set. By formula (5.5.7) the set K1 is disjoint with the set E(K). This leads to a contradiction, because, by Proposition 5.5.1, K1 contains an extremal point. COROLLARY 5.5.3. If a set K is compact, then
cony K = cony E(K), COROLLARY 5.5.4. For every compact convex set K,
K = conv E(K). PROPOSITION 5.5.5. Let X be a locally convex topological space. Let Q be a compact set in X such that the set conv Q is also compact. Then the extreme points of the set conv Q belong to Q. Proof. Let p be an extreme point of the set conv Q. Suppose that p does not belong to the set Q. The set Q is closed. Therefore, there is a neigh-
bourhood of zero U such that the sets p+ U and Q are disjoint. Let V be a convex neighbourhood of zero such that
V-V C U. Then the sets p+ V and Q+ V are disjoint. This implies that p e Q+ V. The family {q+ V: q e Q} is a cover of the set Q. Since the set Q is compact, there exists a finite system of neighbourhoods of type qi+ V, n
i = 1,2, ..., n, covering Q, QC U (qi+V). i=1
Let
Ki = conv ((qi+V) n Q). The sets Ki are compact and convex ; therefore
conv(K1 u ... u Kn) = conv (K1 u ... U Kn) = conv Q. Hence n
patki, at i=1
n
0, i=1
at=1, kiaKi.
Since p is an extreme point of conv Q, all at except one are equal to 0.
Weak Topologies in Banach Spaces
241
This means that there is such an index i that
peKi C Q+V, which leads to a contradiction. REMARK 5.5.6. In the previous considerations the assumption that the space X is locally convex can be replaced by the assumption that there is a total family of linear continuous functionals I' defined on X. Indeed, the identity mapping of X equipped with the original topology into X
equipped with the F-topology is continuous. Thus it maps compact sets onto compact sets. Therefore, considering all the results given before in the space X equipped with the T-topology we obtain the validity of the remark.
5.6. EXISTENCE OF A CONVEX COMPACT SET WITHOUT EXTREME POINTS
Roberts (1976, 1977) constructed an F-space (X,
II
II) and a convex
compact set A C X, such that A does not have extreme points. The fundamental role in the construction of the example play a notion of needle points (Roberts, 1976). Let (X, II). be an F-space. We say that a point x0 e X, x0 0, is a needle point if for each E > 0, there is a finite set FC X such that II
x0 e cony F,
(5.6.1)
sup {MMxjI : x e F} < e,
(5.6.2)
cony {0, F} e cony {0, xo}+B8i
(5.6.3)
where, as usual, we denote by BE the ball of radius e, Be = {x: IIxii < E}.
A point xo is called an approximative needle point if, for each E > 0, there is a finite set F such that (5.6.2) and (5.6.3) hold, and moreover xo a conyF+B8.
(5.6.4)
Since E is arbitrary, it is easy to observe that xo is a needle point if and only if it is an approximative needle point. Let E denote the set of all needle points. The set Eu {0} is closed.
Chapter 5
242
From the definition of needle points and the properties of continuous linear operators we obtain PROPOSITION 5.6.1. Let X, Y be two F-spaces. Let T be a continuous linear operator mapping X into Y. If x0 e X is a needle point and T(x0) # 0, then T(xo) is a needle point.
x0
We say that an F-space (X, II ID is a needle point space if each x0 e X, 0 is a needle point.
The construction of the example is carried out in two steps. In the first step we shall show that in each needle point space there is a convex compact set without extreme points, in the second step we shall show
that a large class of spaces (in particular, spaces LP, 0 < p < 1) are needle point spaces. THEOREM 5.6.2 (Roberts, 1976). Let (X, II ID be a needle point F-space. Then there is a convex compact set E C X without extreme points.
Proof. Without loss of generality we may assume that the norm II II is non-decreasing, i.e. that IItxUI is non-decreasing for t > 0 and all x e X. Let {En} be sequence of positive numbers such that co
fEn < X00.
(5.6.5)
n=o
Let xo # 0 be an arbitrary point of the space X. We write E0 = conv({0,xo}). Since X is a needle point space, there is a finite set F = El = {x', ..., x,} such that (5.6.1)-(5.6.3) holds for e = e0. For each x;, i = 1, ..., n1, we can find a finite set F; such that
x; a conv({0} u F;),
(5.6.6)1
sup {IIxJI : x e Fl} < nl ,
(5.6.7)1
1
conv ({0} u F;) C conv {0, xi }+BB, .
(5.6.8)1
?it
Observe that (5.6.8)1 implies
conv({0} u E2) C conv({0} u EI)+Be,,
(5.6.9)1
Weak Topologies in Banach Spaces
243
where n,
E2=
F. 1
M
(5.6.10)1
The set E2 is finite, and thus we can repeat our construction. Finally, we obtain a family of finite sets E. such that for each x e En we have x e conv ({0} u En+1),
(5.6.6)19
sup {IIxjI : x e En} < sn,
(5.6.7)19
conv({O} u E,,+1) C conv({O} v En)+BE,.
(5.6.9)19
Let OD
Ko = conv (U En u {0}). n=0
The set Ko is compact, since it is closed and, for each s > 0, there is a finite s-net in Ko. Indeed, take no such that Co
En<E. n=n, no
By (5.6.9),, the set U E. constitute an s-net in the set Ko. 19=0
Observe that no x
0 can be an extremal point of Ko, since 0 is the Co
unique point of accumulation of the set U En, and, by construction, n=o
no x e En is an extremal point of Ko. Thus the set KO-Ko does not have extremal points.
Now we shall construct a needle point space. Let N(u) be a positive, concave, increasing function defined on the interval [0,+oo) such that N(0) = 0 and lim N(u) = 0. n-o
u
(in particular, N(u) could be uP, 0 < p < 1).
Let Q _ [0,1]' be a countable product of the interval [0, 1] with the measure µ as the product Lebesgue measure. Let E be a o-algebra induced
Chapter 5
244
by the Lebesgue mesurable sets in the interval by the process of taking product. 1
Take now any function f(t) e L°°[0,1] such that f f(t)dt = 1. We 0
shall associate with the function f a function S{(f) defined on [0,1]' by the formula
Si(f)It = f(tt), where t = {tn}. Observe that the norm of SS(f) in the space N(L(Q,E,/t)) is equal to 1
i = 1, 2, ...
IIS{(f)II = f N(f(t))dt,
(5.6.11)
0
Of course SS(f) can be treated as an independent random variable. Thus, using the classical formula n
E2
n
(Xi-E(X{)) _
E2(X{- E(Xi)) n
we find that, for at >, 0 such that Y at = 1, i=1
n
n
f [ Y at(Si(f)-1)12d/t = Y f
a
t=1 n
i=1
n
aYat f (SI(f)-1)2dp i=1
n
= a f (f(t)-1)2dt,
(5.6.12)
0
where
a = max {a1, ..., an} . By the Schwartz inequality we have n
n 2
f I atS(ft)-1 d/t < f S'at(Si(f)-1)2d/t. fd
i=1
i=1
(5.6.13)
Weak Topologies in Banach Spaces
245
The function N(u) is concave, hence the following inequality results directly from the definition (compare the Jensen inequality for convex functions) n
n
N(' aiui) > i=1
aiN(ui).
(5.6.14)
i=1
As an intermediate consequence of formula (5.6.14), we infer that for each ge N(L(SQ,2,u))r)L(SQ,E,p), we have
IIg!
(5.6.15)
n
By (5.6.12), (5.6.13) and (5.6.15) we obtain n
1
I asS{(f)-1
-
1)2dt)1,2).
(5.6.16)
0
Now we shall introduce the notion of a-divergent zone. Let f e L°°[0,1] 1
be such that f f(t)dt = 1 and let 6 > 0. An interval [a,b], 0 < a < b < 1, 0
is called a a-divergent zone for the function f if for arbitrary real numbers,
a1, ..., an, at > 0 such that a, + ... + an = 1 we have sup
2: at Si(f)
(5.6.17)
a
and
sup L,, aiSi(f) < a.
(5.6.18)
ag <_a
LEMMA 5.6.3 (Roberts, 1976). Let N(u) be as above. Then for each b > 0 1
there are a non-negative function f e L°°[0,1] such that f N(jf(t)I)dt < a 0 1
and f f(t) dt = 1 and a number a, 0 < a < b, such that the interval [a, b] 0
is a 6-divergent zone for the function f.
Proof. In view of the properties of the function N it is easy to to find
Chapter 5
246
a function f e L°°[0,1] such that f f(t)dt = 1 and 0 1
f N(If(t)I )dt < m-.
(5.6.19)
0
where m > 1/b. Take a1, ..., an such that a1+ ... +an < 1 and at > b, i = 1,2, ..., n. Thus n < m and, by (5.6.11) and the triangle inequality, we obtain m
1
fatSi(f) <m f N(I f(t)I)dt < 6. i=1
(5.6.20)
0
For chosen f, by (5.6.16) there is an a such that (5.6.18) holds.
El
PROPOSITION 5.6.4 (Roberts, 1976). A function equal to 1 everywhere is a needle point in the space N(L(S2,E,u)).
Proof. Let e be an arbitrary positive number. Let a positive integer k be chosen so that N(1/k) < e/3. Let 6 = E/3k. By Lemma 5.6.3 we can 1
choose functions f1i ..., fk r- L`°[0,1] such that fi(t) > 0, f f (t)dt = 1, 0 1
f N(I f (t)I)dt < 6 i = 1,2, ..., k and fi have disjoint 6-divergent zones 0
[ai, bi].
Let k ti=1
By (5.6.16) there is an n such that 1/n < min ai and 1-i<-k
x
f Si(f)-1l < E.
1
n =1
Thus, for the set F= {S1(f ), ... , S,(f)}, (5.6.4) holds. By the choice of fi we trivially obtain 11
Ilft!=
i=1
fil
and this implies (5.6.2).
3
Weak Topologies in Banach Spaces
247
Now we shall show (5.6.3). Take al, ..., an > 0 such that a,+... + +an = 1. For each j, j = 1, ..., k, we shall write
Lj =
atSt(fj), at>bj
Mj = I atSt(fj), aj<w
Rj =
at SS(fj) aKaj
Of course, k +=1
Lj+Mj+Rj atSS(f)= k 1=1
By the definition of the 6-divergence zone we have IILjII <6,
(5.6.21)
IIRj-cjII <6,
(5.6.22)
where
at.
cj = a,
Thus
k.
<3 and
(5.6.23)
if if k
k
1
k j=1
1
3.
cj
The intervals [aj, bj] are disjoint. Thus each at can belong to at most one interval [aj, bj]. Therefore
f
n
2:Midju = Of
j=1
kfj=1IaiE[aj.bj] atSS(fj)du k
=1 at
k t=1
k
Chapter 5
248
and by (5.6.15) k
MJ
Finally, for c =
ci, 9=1
asss(f)-c
(5.6.24)
ti=1
and (5.6.3) holds. Thus 1 is an approximative needle point and it is also a needle point.
THEOREM 5.6.5 (Roberts, 1976). For N, (Q,E,u) as above, the space N(L(Q,E,p)) is a needle point space. g 0. The function g induces a continuous Proof. Take a g e linear operator Tg, Tgf = gf mapping N(L(Q,E,,u)) into itself. Observe
that Tg(1) = g. Thus, by Propositions 5.6.1 and 5.6.4, g is a needle point. Since each element of the space N(L(Q,E,u)) can be approximated by bounded functions and the set of needle points with added 0 is closed, each element different from 0 is a needle point.
.
By the classical measure theory, there is a one-to-one mapping of the interval [0, 1] onto [0,1]' preserving the measure. This implies that the spaces N(L(E, Q ,y)) and N(L) are isomorphic. In this way we obtain COROLLARY 5.6.6. If the function N has the properties described above, then there is in N(L) a convex compact set without extreme points.
Shapiro (1977) formulated the problem of the existence of an F-space
X with a trivial dual in which each convex compact set has extreme points. Kalton (1980) showed that certain spaces N(L) have these properties. Some further information can be found in Kalton and Peck (1980).
Chapter 6
Montel and Schwartz Spaces
6.1. COMPACT SETS IN F-SPACES
The definition of compact sets implies that if K is a compact set in an F-space X, then for any neigbourhood of zero U there is a finite system of points x1i ..., x,, such that n
K C U (xi+ U).
(6.1.1)
i=1
PROPOSITION 6.1.1. Let X be an F-space. Let K be a closed subset of the
space X such that, for every neighbourhood of zero U, there is a finite system of points x1, ..., xn such that (6.1.1) holds. Then the set K is compact. Proof. Let 1
Un.=cx:IIxII
x1 such that yn e K1 = (x"+ U,). The set K1 is a closed subset of K, therefore, there are an infinite subsequence {y;,} of the sequence {yi} and an element {x2} such that y; a K2 = (x2+ U2). Repeating this argument, we find by induction that there are a family of sequences {yn} is a subsequence of the sequence {yn-1} and a sequence of points {xk} such that k / ynelxk+Uk),
n,k=1, 2,... 249
(6.1.2)
250
Chapter 6
Let {yn} = {y"}. The sequence {yn} is fundamental. Indeed, if n,m > k, then yn, y. are elements of these sequences {yn}, thus by (6.1.2) JI yn-ymII
< 2/k. The space X is complete, hence there is a limit y of the sequence {yn}. Since the set K is closed, y e K. This means that each sequence of elements K contains a convergent subsequence. Then the set K is compact.
Proposition 6.1.1 holds also for complete linear topological spaces but here we shall restrict ourselves to the metric case. Let X be a finite dimensional space. It is easy to verify that each closed bounded set in X is compact. Since a finite-dimensional space is locally
bounded, this means that there are neighbourhoods of zero such that their closures are compact sets. We say that an F-space is locally compact if there is a neighbourhood of zero U such that the closure U of the set U is a compact set. THEOREM 6.1.2 (Eidelheit and Mazur, 1938). Each locally compact space X is finite-dimensional.
Proof. Let V be such a neighbourhood of zero that the closure V of V is a compact set. Let Y be an arbitrary finite dimensional subspace different from the whole space X. Obviously, X # Y+ V. Suppose that VC Y+ V. Then Y+ V = Y+ V. Lemma 5.1.4 implies that the set Y+ V is closed. On the other hand, the set Y+ V is open. and we obtain a contradiction. Therefore, there is an a e V such that a Y+ V. Now we shall construct by induction the following sequence. y, is an arbitrary element. Suppose that the elements y,, ..., yn are defined. Let Yn be the space spanned by those elements. Let yn+, be such an element that yn+, e V and yn+10 Yn+V. It is easy to verify that if the space X is infinite-dimensional, we could construct such an infinite sequence {yn}. But this sequence would not contain any convergent subsequence, because, for k n, yk-y,,, 0 V. This leads to a contradiction since the set V is compact. Therefore, the space X is finite dimensional.
Montel and Schwartz Spaces
251
PROPOSITION 6.1.3 (Mazur, 1930). Let X be a Bo-space. If a set A C X is compact, then the set cony A is also compact.
Proof. Let U be an arbitrary convex neighbourhood of zero. Since the set A is compact, there is a finite system of elements x1, ..., xn such that n
A Ct=1 U (xi+U).
(6.1.3)
Let
K = conv({x...... xn}). The set K is bounded and finite-dimensional, therefore, there is a finite system of points y, ..., ym such that
K Ci=1 U (yi+U).
(6.1.4)
Thus, by (6.1.3),
conv Acconv(K+U)= K+U. Therefore m
n,
convA Ci=1U (yi+U)+U = i=1 U (yi+2U). The arbitrariness of U and Proposition 6.1.1 implies the proposition.
0
Since Proposition 6.1.1 holds for complete topological spaces, Proposition 6.1.3 holds for complete locally convex spaces. Therefore, in the case of complete spaces, we can omit in Corollary 5.5.5 the assumption that the set conv Q is compact.
6.2. MONTEL SPACES
In the preceding section we proved that each locally compact space is finite-dimensional. In finite dimensional spaces each bounded closed set is compact. There are also infinite-dimensional spaces with this property. Examples of such spaces will be given further on. F-spaces in which each closed bounded set is compact are called Montel spaces.
Chapter 6
252
PROPOSITION 6.2.1. Locally bounded Montel spaces are finite-dimensional.
Proof. Let X be a locally bounded Montel space. Since X is locally bounded, there is a bounded neighbourhood of zero U. The space X is a Montel space, hence the closure U of the set U is compact. Therefore, the space X is locally compact and, by Proposition 6.1.2, finite-dimensional.
PROPOSITION 6.2.2 (Dieudonne, 1949; Bessaga and Rolewicz, 1962). Every Montel space is separable. The proof of the theorem is based on the notion of quasinorm (Hyers, 1939; Bourgin, 1943), similar to the notion of pseudonorm. Let X be an F-space. By 521 we denote the class of all open balanced set. Let A E 521. The number
[x]A = inf t > 0:
x t
EA
is called the quasinorm of an element x with respect to the set A. Quasinorms have the following obvious properties : (a) [tx]A = I t I
(b) if A ) B, then [x]A < [xB], (c) the quasinorm [x]A is a homogeneous pseudonorm (i.e., satisfies the triangle inequality) if and only if the set A is convex. PROPOSITION 6.2.3. [x+y]A+B < max([x]A, [y]B).
Proof. Let us write r = max([x]A, [y])B). Let e be an arbitrary positive number. By the definition of the quasinorm, xE (1-}-e)rA and y e (1+e)rB. Hence
x+ye(1+e)r(A+B). Therefore [x+y]A+B < (1+e)r. The arbitrariness of a implies the proposition. PROPOSITION 6.2.4. Let a sequence {An} C 521 contitute a basis of neigh-
bourhoods of zero. Then a sequence {x.} tends to 0 if and only if lira [xfn]A. = 0 (n = 1, 2,...). (6.2.1) M-0
Montel and Schwartz Spaces
253
Proof. Necessity. Suppose that xm-*O; then, for arbitrary E > 0, and n, there is an mo dependent on n and a such that for m > mo, xm e eAn, whence [xm]An < 8.
Sufficiency. Let us suppose that (6.2.1) holds ; then for every n there is an mo dependent on n such that, for m > m0, [xm]A,, < 1. This means that xm a A. Since {An} constitutes a basis of neighbourhoods of zero, xm-*0. C 91 constitute a basis of neighbourhoods of zero. A set K is bounded if and only if there is a sequence of numbers {Nm} such that sup[x]An < Nn. PROPOSITION 6.2.5. Let
XEK
Proof. Sufficiency. Suppose that a sequence {Nn} with the property described above exists. Let {xm} be an arbitrary sequence of elements of K and let {tm} be an arbitrary sequence of scalars tending to 0. Then [tm xm]A, < I tm I Nn -± 0 .
Hence, by Proposition 6.2.4, the sequence {tmxm} tends to zero. Therefore, the set K is bounded. Necessity. Suppose that there is an index n such that sup [x]4, = +oo. zEK
Let Then there is a sequence {xm} C K such that 0 < tin = 1/[xm]dn. The sequence {tm} tends to 0. On the other hand 1Xm1A.->oo.
[tmxm]A, = 1, fence, by Proposition 6.2.4, the sequence {tmxm} does not tend to 0. This implies that the set K is not bounded.
Proof of Proposition 6.2.2. Let X be a non-separable F-space with the F-norm IIxil. Since the space X is non-separable, there are a constant b > 0 and an uncountable set Z such that 11z -z'iI > S for z, z' c- Z, z 4 z'. Let K. = {x: jlxii < 1/n}, and let us write briefly [x],, = [x]gn. Since the set Z is uncountable, there is a constant M1 such that the set
Z1 = Z n {x e X: [x]1 < M1} is also uncountable. Then there is a constant M2 such that the set Z2 = Z1 n {x e X: [x]2 < M2} is uncountable. Repeating this argumentation, we can find by induction a sequence
of uncountable sets {Z,,} and a sequence of positive numbers {Mn} such that the set Z,, is a subset of the set Z,,-1 and sup [x]{ < Mn for
i=1,2,...,n.
XeZn
Chapter 6
254
Let us choose a sequence {z.} such that zn a Zn and z{ Then
zk for i * k.
sup [zn]k < max (Mk, [z,lk, ... , [4-1]0
<+oo
fork=1,2,...
Therefore, by Proposition 6.2.5, the sequence {zn} is bounded. On the other hand, zn a Z, hence jIz{-zkll > 6 for i k. This implies that the set {Z-n} is not compact. Therefore, X is not a Montel space. The following question has arised : is it sufficient for separability if we assume that each bounded set is separable? The answer is negative. Basing on the continuum hypothesis Dieudonne (1955) gave an example of a non-separable B, -space in which each bounded set is separable.
For F-spaces such an example was given by Bessaga and Rolewicz (1962). For Bo-spaces it was given by Ryll-Nardzewski.
PROPOSITION 6.2.6 (Ryll-Nardzewski, 1962). There is a non-separable Bo space in which all bounded sets are separable. Proof. Let S denote the class of all sequences of positive numbers. We
introduce in S the following relation of order 3. We write that {fn} 3 {gn} if fn < gn for sufficiently large n. A subclass S, of class S is called limited if there is a sequence {hn} e S such that { f n} 3 {hn } for all sequences {fn} e S. Let us order class S in a transfinite sequence { f;} of type co, (here we make use of the continuum hypothesis). Now we define another sequence of type co, as follows : {gn} is the first sequence (in the previous
order) which is greater in the sense of relation 3 than all {fn} for a < 8. It is easy to see that no non-countable subclass of this sequence is limited. Let us consider the space X of all transfinite sequences {xa} (a < (0,) of real numbers such that x,, vanishes except for a countable number of indices and 1J{x9}IJn = f gn jxBj < -boo,
n = 1, 2, ...
B<ml
Let us introduce a topology in the space X by the sequence of pseudonorms jj{x,,}IIn. The space X with this topology is a Bo*-space. We shall
Montel and Schwartz Spaces
255
show that the space X is complete. Let {x'n} be a fundamental sequence of elements of the space X. Let
A = la:
xa =P4- 0 for certain n}.
The definition of the space X implies that the set A is countable. Therefore, the subspace Xo spanned by the elements {xQ}, {xQ}, ... is of the type L1(am,n) Thus it is complete. Therefore, the sequence {xa} has the limit {xa} e X0 C K.
To complete the proof it is enough to show that each bounded set Z C X is separable. If the set Z is bounded, then there is a sequence of positive numbers {M16} such that sup II{xa}J1n < Mn, (z
n = 1, 2, ...
Z
Let k be a positive integer. By Ik we denote the set of all such indices that there is an x = {xa} e Z and such anindex /3 such that [xp] > 1/k. Then we have
1
k gn < jj{xa}Ijn < Mn. Hence, for /3 e Ik, we have {gn} - {Mn} and a,
by the property of the class {g,6,} the set Ik is countable. Let I = U It. k=1
Then the set I is, of course, also countable. Let y be the smallest ordinal greater than all the terms of the set I. Then from the definition of the set I it follows that if {xa} e Z, then .x5 = 0 for 6 > y. Thus the set Z is separable.
6.3. SCHWARTZ SPACES
Let X be an F*-space. We say that a set K is totally bounded with respect to a neighbourhood of zero U, if, for any positive e, there is a finite system CO
of points x1i ..., xn such that KC U (xc+eU). A set which is totally i=1
bounded with respect to all neighbourhoods of 0 is called totally bounded or precompact. Proposition 6.1.1 implies that if a set K is closed and totally bounded with respect to all neighbourhoods of zero, then it is compact.
Chapter 6
256
An F-space X is called a Schwartz space if, for any neighbourhood of zero U, there is a neighbourhood of zero V totally bounded with respect to U. PROPOSITION 6.3.1. Let X be a Schwartz space. Then its completion X also a Schwartz space. Proof. Let U0 be a neighbourhood of 0 in X. Let U = U0 n X. Since X is a Schwartz space, there is a neighbourhood of zero V C X such that for any e > 0 there is a system of points x1, ..., xn such that n
V C U (xi+EU). =1
Thus n
n
C U (xi+EU) c U (xt+2E U). 1=1
9=1
PROPOSITION 6.3.2. Every Schwartz F-space is a Montel space.
Proof. Let K be a closed bounded set. Let U be an arbitrary neighbourhood of zero. Since we consider a Schwartz space, there is a neigh-
bourhood of zero V totally bounded with respect to U. The set K is bounded. Then there is a positive a such that KC aV. The neighbourhood V is totally bounded with respect to U, and so there is a finite
system of points
y1,
, y,,,
U
such that Vc U yt+ a) . Let x{ = ay{. i=1
00
Then KC aV C U (xt+U). i=1
Since the set K is closed and U is arbitrary, the set K is compact (see Proposition 6.1.1).
There are also Montel spaces which are not Schwartz spaces. An example will be based on PROPOSITION 6.3.3. Let am, n < a.+,,.. The space M(am, n) (or LP(am, n))
(see Example 1.3.9) is a Schwartz space if and only. if, for any m, there
Montel and Schwartz Spaces
257
is an index m' such that am,n
lim InI-o am',n
= 0,
(6.3.1)
where InI = In1I+In2I+... +Inkl. Proof. Sufficiency. Let U be an arbitrary neighbourhood of zero. Let Uo be such a neighbourhood of zero that Uo+ Uo C U. Let m be such an index that the set Um. _ {x: IIXIIr < (resp. pm(x) < m m l}
is contained in Uo. Let m' be such an index that (6.3.1) holds. Let E bean arbitrary positive number. Since (6.3.1) holds, there is a finite set A of indices such that, for n 0 A, am,n
< E.
(6.3.2)
am',n
Let L be a subspace of M(a,n,,a) (resp. L2,(a,n, n)) such that {xn} e L if and only if xn = 0 for n 0 A. The space L is finite-dimensional. Let
KL = {x e L: am,,n Ixnl < 1}. The set KL is compact. Then there is n
a finite system of points yl, ..., yn such that KL C U (yi+ UO). 4=1
Let V = {x: IIxII., < 1/m}. Let x be an arbitrary element od V. By (6.3.2) there is an xo e VnL such that x-xo e EU,,, C EUa. Since VniLCKL, we have n
/'' V C V n L+EU0 C KL+EUo C U (Y{+eUo+EUo) %=1
n
C {=1 U (Yt+EU). Thus the set V is totally bounded with respect to U. Necessity. Let us suppose that there is such an m that for all m' > m lim sup am,n = 5m > 0. InI- m
am'.n
(6.3.3)
Chapter 6
258
Let U = {x: IIXIIm < 1 (resp. pm(x) < 1)}. Let V be an arbitrary neigh-
bourhood of zero. Then there are a positive number b and an index m' such that VD {x: IIXm'II < b}. Let A be the set of such indices n that
am,n > m 2
am',n
Since (6.3.3) holds, the set A is infinite. Let yn = {yk}, where
fork=n,
b
n
fork
n. It is obvious that yn e V (n = 1, 2, ...). On the other hand, if n,n' a A, 0
1
n
n', then Ilyn-yn'll > bb.,/2 (resp. pm(yn-yn') > [bb.,/2]P). Since
the set A is finite, this implies that V is not totally bounded with respect to U. The arbitrariness of V implies the proposition. Example 6.3.4 (Slowikowski, 1957) Example of a Montel space which is not a Schwartz space. Let k, m, n1, n2 be positive integers. Let k m-n, ' . ak,m,n.,n. = nl max 1, n2
Let X denote the space of double sequences x = {xn1, n2} such that IIXIIk.m = SUp ak,m,ni,n. iXn,,n.l < + 00
with the topology determined by the pseudonorms Ilxllk,m. Xis a Be-space
of the type M(am, n). The space X is not a Schwartz space. Indeed, let us take two arbitrary pseudonorms IIXIIk,m and IIXIIk',m' Let nl > m,m'. Then li m n_*
ak,m,n"n' ak',m',ni,n,
= (n,0)k-k' >
0.
Therefore
lim SUP ak ',m
,
.n.,ns
and from Proposition 6.3.2 it follows that the space Xis not a Schwartz space.
Montel and Schwartz Spaces
259
Now we shall show that the space X is a Montel space. Let A be a bounded set in X. Since X is a space of the type M(am, ), it is enough to show that 0.
lim ak.m.n.,n, Sup InHw zee
(6.3.4)
Let us take any sequence {(nl,n2)}, such that lim Ini I+Ina I = +o-o. We have two possibilities : (1) n'-goo,
(2) nl is bounded.
Let us consider the first case. Let x = {Xnl, n2} e A and let k' > k, m' > m. Since the set A is bounded, there is a constant Mk,,m, such that ak'.m',ni,n.
Mk',m' .
Then for sufficiently large nl ak,m,nr,nr .Ixnr,nr I < Mk',,n,
(nl)k-k
-> 0.
(6.3.5)
Let us consider the second case. Let m' > in and m' > nl, k' > k Then ak,m,nl,nr IXnr,nr I
< Mk',m' (n9)m-m'
0.
Therefore, by (6.3.5) and (6.3.6) formula (6.3.4) holds. This implies that X is a Montel space.
6.4. CHARACTERIZATION OF SCHWARTZ SPACES BY A PROPERTY OF F -NORMS
In the previous section we introduced the notion of Schwartz spaces. Now we shall give a characterization of those spaces by a property of F-norms. Let Y be an arbitrary F*-space with the F-norm IIxII and let s be an arbitrary positive number. We write c(Y, e, t) = inf {IItxJI : X e Y, IIxII = e}
Chapter 6
260
if there is such an element x e Y that I lxII = e and Jr
c(Y,e,t)= 0
for t=,k 0,
fort=0,
if sup Ilxll < E. xEY
THEOREM 6.4.1 (Rolewicz, 1961). Let X be a Schwartz space. Then, for every increasing sequence of finite-dimensional subspaces {Xn} such that 00
the set X * = U Xn is dense in the whole space X, the functions c (X/X,,, e, t) n=1
are not equicontinuous at O for any e. Proof. Let us write
K, = {x e X: IlxII <,}.
Suppose that the theorem does not hold. Then there are a positive e and a sequence of finite-dimensional spaces {X,,} such that X. C Xn+1, W
the set X* = U Xn is dense in the space X and the functions c(X/Xn,s,t) n=1
are equicontinuous at 0. Let 6 be an arbitrary positive number. By .1. we denote such a positive number that
c(X/Xn, e, t) < 2
for 0 < t <.lo.
Since the set X* is dense in X, this implies that
c(X*/Xn, e, t) < 2
for 0 < t <.lo.
We shall choose by induction sequences of positive integers {kn} and of elements {xn} such that (1) IIxnII < 6, (2) xn E Xk,,, (3) xn - x{ 0 AOKe12
As k1 we take 1 and as x1 we take an arbitrary element of X1. Let us suppose that for some n we have chosen the elements x1, ..., xn and the integers k1, ..., kn. By hypothesis, there is in the space X/Xk a coset Z
Montel and Schwartz Spaces
261
such that IHZII = e and IIXoZII < 6/2. By we denote an arbitrary element of the coset XOZ such that IIx,+1Jl < 6, and by Xk.+, we denote a subspace containing the element x e Xkn, E eAOZ,
xn+i-x
IIZ I I = E -
Ao
2 we have
for i = 1,2,... ,n-l,n. Thus, by (3), the set Ka is not totally bounded with respect to the set KE,2. The arbitrariness of 6 implies that X is not a Schwartz space. It is not known whether the inverse theorem to Theorem 6.4.1 holds in the general case. THEOREM 6.4.2 (Rolewicz, 1961). Let X be an F-space for which there is a positive eo such that for every x e X, x 0, sup IItxHI > Ep. If there is t>o
a sequence of finite-demensional spaces such that X C X,,+, and the functions c(X/XX., r, t) are not equicontinuous at 0 for any e, then the space X is a Schwartz space. Proof. Let us write
K,, = U tK,,. j:I
Let e be an arbitrary positive number such that KE+Ke C Kep
(6.4.1)
.
Suppose that there is a sequence of finite-dimensional spaces such that X. C and the functions c(X/XX,e, t)are not equicontinuous at 0. This means that there are a sequence {nk} of positive integers and a positive number 6 such that for each p > 0 and for sufficiently large k
sup c
o5t
6,
(6.4.2)
i.e. there is auk, 0 < /Uk < p such that C (X/Xnk, e, .k) > 6 .
(6.4.3)
This implies that in the quotient space
KaCpkKeCpKK
(6.4.4)
Chapter 6
262
Thus Kb c pK, .
(6.4.5)
Let x be an arbitrary element of a norm less than 6, IIxII < 6. Let Z denote the coset containing x. Since (6.4.5), IIZ/Full < e. Let x0 e Z be such an element that Ilxo/pII < e. This implies that xoepKE.
(6.4.6)
Let us write
x = xo+(x-x0).
(6.4.7)
Then (x-xo) e Xnk and x-xo e KK+4uKE c p(KE+KE)
(6.4.8)
Since the space Xnk is finite-dimensional, by (6.4.1) the set Xnk n (KE+KE) is bounded. Thus it is totally bounded with respect to the set UK,. There-
fore, by Proposition 6.2.3 and by (6.4.8), there is a finite system of points yl, ..., Yn such that m
Ka e
ti=1
(yi+p(Ke+Ke'+Ke)).
Hence the set Kb is totally bounded with respect to the set K = KE+K,' +K,. Since the set Ke constitutes a basis of neighbourhoods of zero, the space X is a Schwartz space. COROLLARY 6.4.3 (Rolewicz, 1961). Let X be an F-space with a basis {en}. Suppose that there is a positive ro such that sup IItxII>eo o
for every x e X, x 0. Then the space X is a Schwartz space if and only if the functions c(X,,,e,t), where X. are the spaces generated by the elements en+1, en+2, ..., are not equicontinuous at 0 for any E.
Basing ourselves on Corollary 6.4.3, we shall give an example of a Schwartz space which is not locally pseudoconvex.
Montel and Schwartz Spaces
263
Example 6.4.4 Let {p,,} be a sequence of real numbers 0 < pn < 1. By 1(p") we denote the space of all sequences x = {xn} such that Ilxll =
Ixnlp" < +oC n=1
with the F-norm Ilxll It is easy to verify that 10'") is an F-space. The sequence {en}, en = {0, ...,0,2,0,. ..}, constitutes a basis in the space n-th place
l(p"). By simple calculation we find that C(Xn, E, t) = 8jtI2n,
where X, is the space spanned by the elements en+I,en+2, ... and
Pn = inf{pt: i > n}. Therefore, by Corollary 6.4.3, the space l('") is a Schwartz space if and only if pn->0. Let us remark that if pk->0, then the space l(p") is not locally pseudoconvex. Indeed, let 6 be an arbitrary positive number. Let xn = {0, ..., 0,
(6)1jp", 0, ...}. Then Ilxnll = 6. Let p be an arbitrary positive number n-th place
and let np be such an index that, for n > np, pn < p12. Let us take the p-convex combination of the elements xnn+1, ..., x.+k. Then n,+k
xn,+i+ - - - +xn,+k
8
()PY.
n,+l
II
k (.4 p
OP )P12
= 6k 1'2 -> oo.
The arbitrariness of p and 6 implies that the space 1(") is not locally pseudoconvex.
6.5. APPROXIMATIVE DIMENSION
Let X be an F*-space. Let A and B be two subsets of the space X. Let
B be a starlike set. Let M(A,B,r) = sup{n: there exist n elements xl, ..., xn E A such that xt-xk 0 EB for f lr k}.
Chapter 6
264
The quantity M(A, B, e) is called the s-capacity of the set A with respect to the set B. Obviously, M(A, B, e) is a non-increasing function of E.
Let
M (A, B) =
ke):
cp (e) - real positive function,
lim M(A(EB,
e)
+ool.
Let us denote by 0 the family of all open sets and by J the family of all compact sets. The family of real functions
M(x)= n n M(A, B) AE6 BECJ
is called the approximative dimension of the space X (see Kolmogorov, 1958).
PROPOSITION 6.5.1. If two F*-spaces X and Y are isomorphic, then
M(X) = M(Y). Proof. Let T be an isomorphism mapping X onto Y. Then a set A is open (compact) if and only if the set T(A) is open (compact). Hence
M (X) = n n M (A, B) AeCj BE6 AdX BEX
= n n M(T(A), T(B)) = M(Y). Ae7 BE6 AcX BcX
PROPOSITION 6.5.2. Let X be a subspace of an F*-space Y. Then
M(X) D M(Y). Proof. Let U be an open set in X. Let x e U and let rx = inf {I Ix-YII: Y 0 U, y e X} .
Let us put
V=!U {zE Y: IIz-xHI < zrx}.
Montel and Schwartz Spaces
265
The set Vc Y is open as a union of open sets. On the other hand, it is easy to verify that U = Vr) X. Then
M(Y)= n n M(A, B) C n n M(AnX, B) AEg BEG
ACYBcY
AE
BEG
AcYBcY
n n M(AnX,BnX)
AEI BEY
AcYBcY
n n M(A, B) = M(X).
Acg BEG
AcX BcX COROLLARY 6.5.3. If
dimzX
M(X) D M(Y)
(resp. M(X) = M(Y)).
In order to prove a similar fact for linear codimension, we shall use the following PROPOSITION 6.5.4. Let Y be an F-space and let X be a subspace of the space Y. Let us consider the quotient space Y/X. Let K be a compact set in Y/X. Then there is a compact set Ko C Ysuch that K = {Ex]: x e Ko}, where, as usual, we denote by [x] the coset containing x.
Proof. Since no confusion will result, we denote by the same symbol II II the F-norm in Yand the norm induced by it in the quotient space Y/X. We shall say that a finite system of elements z1, ..., za constitutes a finite c-net in a set A if for any x e A there is an index i such that Ilzi-x!I
<e. Let ro be an arbitrary positive number. Since the set Kis compact, there is a finite ro-net Z°, ..., Z°no in K. Let xi, i = 1, 2, ..., no, be arbitrary elements such that Xi e Z°. Let n
Ao = U {x e Y: IIx-xill < r0} . i=1
From the definitions of an
and of a quotient space it follows that
Chapter 6
266
KC {[x]: x e Ao}. Let aZ = sup {p: there is an xZ e Z such that {x: JIx-xZIJ
2r1 = inf a° > 0. ZEE
Indeed, let us suppose that there is a sequence {Zn}, Z. a K, such that aZ, ->O. Since the set Kis compact we can assume without loss of generality that {Zn} tends to Z0 e K. Therefore, for sufficiently large n,
IjZn-Z0H < aaz This implies that a°Zn > s aZo, and we obtain a contradiction, because a4,,-->0.
Let us take a finite r1-net in K, Zi, ..., Z' . The definition of r1 implies that there are points xi e Z;, i = 1, 2, ..., n1, such that n
Al = U {x:
Ilx-x;II
i=1
Repeating this argument, we can choose by induction a sequence of systems of cosets {Zi, ..., Znk} and a sequence of positive numbers {rk} tending to 0 such that (1) Zl, ..., Znk constitute an rk-net in K,
(2) Ak C Ak_1, where Aj = U {x: Jjxx-xjj < ri}. i=1
Let nk
00
ij.
k=1 i=1
The set KO = Ko is compact. Indeed, let a be an arbitrary positive number. k
ni
Let us take an rk such that rk. The finite set U U {xx} constitutes
i=1 j=1 an rk-net in KO, because A. C Ak for m > k. Since the set K0 is closed, the arbitrariness of e implies that the set Kis compact. Let K1={[x]: x c- KO}.
Since the set K0 is compact, the set K1 is also compact. Moreover, [xn] 00
nk
= Zk and the set U U {Zk} is dense in K. Hence K7 K1. k=1 i-1
Montel and Schwartz Spaces
267
Proposition 6.5.4 implies the following fact. Let The a continuous line-
ar operator mapping an F-space Y onto an F-space X. Then for every compact set K in X there is a compact set Ko in Y such that T(K0) = K. Indeed, let Z = {x: Tx = 0}. Then the space X is isomorphic to the quotient Y/Z and the operator T induces the operator T' mapping y e Y into the coset [y] e Y/Z. PROPOSITION 6.5.5. Let X and Y be two F-spaces. If there is a continuous linear operator T mapping Y onto X, then
M(X) C M(Y). Proof. To begin with, les us remark that if T is a linear operator, then
M(A, B, e) > M(T(A), T(B), E). Hence
M(A, B) C M(T(A), T(B)). Since the inverse image of an open set is always open and in our case, by the Banach theorem (Theorem 2.3.1), the image of an open set is open,
n) M(A, B) C B G M(T(A), B).
Bcx
BcY
On the other hand, for any compact set Kc X there is a compact set Ko C Y such that T(K0) = K. Therefore
M(X) = nn BEG M(A, B) C AcXBcX
n
BE6 M(A) B) = M(Y).
AcYBcY
COROLLARY 6.5.6. If
codimi X <_ codimi Y (codiml X = codimt Y), then,
M(X) J M(Y)
(resp. M(X) = M(Y)).
By a simple calculation we obtain PROPOSITION 6.5.7. If X is an n-dimensional space, then
M(X) = {P(8): lim En(0(E) = -boo}. n_.0
Chapter 6
268
PROPOSITION 6.5.8. Let X be an F*-space. If X is not a Schwartz space, then
the set M(X) is empty. Proof. If X is not a Schwartz space, then there is a neighbourhood of zero U such that, for any neighbourhood of zero V, there is a positive number E V such that
M(V, U, cy) = +oo.
(6.5.1)
Let Vn = {x: IIxij < 1/2n}, where IlxHH denotes, as usual, the F-norm in X.
Let us write, for brevity, en = sV. Let (p (e) be an arbitrary positive function. Formula (6.5.1) implies that there is a system of points x", ..., xm. such that mn > ((en), xi E Vn, i = 1, Mn
2, ... , Mn, x7 - xk 0 En U for j = k. Let K = U U {x; }. The set K has a n=1 i=1
unique cluster point 0. Then the set K is compact. On the other hand, M(K, U, en) > mn > rp(sn). Therefore (p (e) does not belong to M(K, U). U The arbitrariness of (p (s) implies that the set M(X) is empty.
The Kolmogorov definition of approximative dimension has some disadvantages. Simply there are "too many" open and compact sets. This is the reason why in many cases it is more convenient to consider a definition of approximative dimension introduced by Pelczynski (1957). The Pelczynski approximative dimension of a space X can be defined (after certain modifications) as the set of functions
fl U
M'(X) = Ue(j VE() n positive
M(nV U)
integers
In other words, M'(X) = {cp(s): for each open set U there is an open v(s) set Vsuch that, for all n, lim _ +oo}.
s o M(nV, U, e) Let { Ui} be a basis of neighbourhoods of zero. Let M,(X) = {(p(8): for KE) = +oo}. each i, for almost all j, for all n, lim
E-0 M(nUi+ j,
UU, e)
PROPOSITION 6.5.9. For all F*-spaces, M'(X) = MM(X).
Proof. The definitions of M'(X) and MM(Y) imply that M'(Y) C MO'(Y). Suppose that cp(e) 0 M'(Y). This means that there is a.neighbourhood of
Montel and Schwartz Spaces
269
zero U such that for all neighbourhoods of zero V there is a positive
integer n such that lim
E
e-o M(n V, U, E)
oo Let Ui be an arbitrary
<
neighbourhood contained in U. Putting V = U{+,(j = 1, 2,...), we find that V(s) does not belong to M0(X) C M'(X). In applications the most useful class is the class Mo(X), because it is described by a countable family of functions. PROPOSITION 6.5.10. Let X be a Schwartz space. Then the set M'(X) is not empty. Proof. Let us choose a decreasing basis of neighbourhoods of zero { Ut} in
such a manner that M(nUt+l, Uj, E) < +oo for all positive integers n. Such a choice is possible, because Xis a Schwartz space. Let (0 (E) =
1 E
for nIl <E
M(n Uz+i, UI, E) i=1
Let us take an arbitrary index i and an arbitrary n > i. Then, for arbi-
trary j and e < 1/n,
0 < M(nUU+,, Ut, E) <M(nU;+i, U{,E)< (e)
(E)
This implies that p(E) a Mo(X). Therefore, by Proposition 6.5.10, p (e) C- M'(X).
PROPOSITION 6.5.11. Let X be an F-space. Then
M(X) D M'(X). Proof. Let K be a compact set and let V be an arbitrary neighbourhood of zero. Then there is a positive integer n such that K C n V. Therefore, M(nV, U, E) >, M(K, U, E). Thus M(nV, U) C M(K, U). This implies the proposition.
From Propositions 6.5.8, 6.5.10, 6.5.11 follows
Chapter 6
270
COROLLARY 6.5.12. X is a Schwartz space if and only if the set M(X) is not empty.
COROLLARY 6.5.12'. X is a Schwartz space if and only if the set M'(X) is not empty.
Proposition 6.5.11 shows that M(X) ) M'(X). We do not know whether the converse inclusion is true. It is so in the case where X is a finitedimensional space. For infinite-dimensional spaces only the following partial answer is known. PROPOSITION 6.5.13. Let X be an infinite-dimensional Schwartz space. Sup-
pose that, for each neighbourhoods of zero U and V, there is a neighbourhood of zero W such that for, all n,
M(nW, U, e) < M(V, U, e)
(6.5.2)
for sufficiently small E. Then M(X) = M'(X). Proof. Formula (6.5.2) implies that n M(nW, U) ) M(V, U). Then the
set M'(X) is equal to the set M'(X) = n n M(V, u). UEG VC-C)
By Proposition 6.5.11
M- (X) C M(X). Now we shall show the converse inclusion. Suppose that a function cp(e) does not belong to M(X). This means that there is a neighbourhood of zero U such that for all neighbourhoods of zero V
liminf E-.u
(Y (e)
M(V, 2U, e)
< +00.
Since the space X is infinite dimensional, M(V, 2U, e) lim E-.o M(V, U, E)
-0.
Thus
liminf E-.O
P(e) ) M(V, U, e
= 0.
Montel and Schwartz Spaces
271
This means that there is a sequence {E k} tending to 0 such that p(Ek) lim v / k->cn M(V, U,ek)
Let Vn = x:
jIxII <
2n
=0.
(6.5.3)
I , where, as usual, IIxII denotes the F-norm in
the space X.
Basing ourselves on formula (6.5.3) we can choose by the diagonal method a sequence {en} tending to 0 and such that 0 (En) lim n-,oo M(Vn, U, E-)
= 0.
This implies that for sufficiently large n there are systems of points {xi,
..., x";,n}, mn > gp(en), such that xi e V. (i = 1, 2, ..., m.,,) and x, oo
-x 0 EnU for i
Mn
j. Let K = U U {xi }. The set K has a unique clusn=1 i=1
ter point 0. Therefore, the set Kis compact. Moreover M(K, U, En) > Mn > (p(En).
This implies that the function (p(e) does not belong to M(X). Hence
M'(X) = M'(X) J M(X). Since it is not known whether the equality M'(X) = M(X) holds in general, we shall prove for M'(X) propositions and corollaries similar to Propositions and Corollaries 6.5.1-6.5.6. PROPOSITION 6.5.14. Let X and Y be two Schwartz spaces. If the spaces X and Y are isomorphic, then
M'(X) = M'(Y). Proof. The above follows immediately from the definition of M'(X) and the fact that the image (the inverse image) of an open set under an isomorphism is an open set.
Chapter 6
272
PROPOSITION 6.5.15. Let X be an F*-space and let Y be a subspace of the space X. Then
M'(X) C M'(Y).
C
Proof. The proof is the same as the proof of Proposition 6.5.2. COROLLARY 6.5.16. If
dimjX
(dimiX= dimLY),
M'(X) 3 M'(Y)
(resp. M'(X) = M'(Y)).
then
PROPOSITION 6.5.17. Let X and Y be two F-spaces. Let T be a continuous linear operator mapping X onto Y. Then
M'(X) C M'(Y). Proof. The Banach theorem (Theorem 2.3.1) implies that the image of an open set is an open set. In general, the inverse image of an open set is an
open set. The rest of the proof is the same as the proof of Proposition 6.5.6. COROLLARY 6.5.18. If
codimjX
(codimiX= codim1Y),
then
M'(X) 3 M'(Y)
(resp. M'(X) = M'(Y)).
It is easy to calculate approximative dimensions of spaces M(a,n, n). For the calculation of M'(X) it is enough to know the functions M(Ui+1, Ui, e) for a basis of neighbourhoods of zero {Un} (see Proposition 6.5.9). PROPOSITION 6.5.19. Let X = M(am,n) be a real Schwartz space. Let
Ui= {x: 1IxIIi< 1}. Then
M(Ui+5, Ui, e)
J
11 n=1
E,
(
1+
-
2 ai,n E ai+9.n
where E'(r) is the greatest integer less than r.
),
(6.5.4)
Montel and Schwartz Spaces
273
Proof. Let x = {xn} and y = be two elements of U{+j. Let x-y e EUi. Then there is an index n such that ai, n I xn -Yn l > e. On the other hand,
ai+j,nxn and ai+j,,,yn are less than
more than E' 1+
1.
Since there may exist no
2- ai, n
- numbers bk such that ai+j, nl bkl < I and
e ai+j. n
Ibk-bk,lai,n > e for k k', and systems with this number of elements exist, the proposition holds. Proposition 6.5.13 and formulas (6.5.4) and (6.3.1) imply that
M'(M(am.n)) = M(M(am.n))
If X = M(am,n) is a complex space, then q9 (E) e M(X) if and only if 4/ (E) belongs to M(X,), where X, is a real space M(am, n). Propositions 6.5.19 and Corollaries 6.5.12', 6.5.16, 6.5.18 imply PROPOSITION 6.5.20 (Pelczynski, 1957). There is no Schwartz space universal (or co-universal) for all Schwartz spaces. Proof. Since for every Schwartz space X the set M'(X) is non-empty, it is
sufficient to show that for any function f(e) there is a Schwartz space Xf such that f(e) 0 M'(Xf). Let af, = 2/k for nk_1 < n < nk, where nk = loge f(1/k)+1. Let Xf = M(am, n), where am, n = (an)'Im. By proposition 6.3.3 the space Xf is a Schwartz space. On the other hand, 1
CO
111(Ui+j,
Ui, k)
HE'
l +k
i
i+j
1
nk
JJ n=1
2nk >f I k). \
Therefore, f(e) 0 M'(Xf). Let us observe that, for any sequence {Xn} of Schwartz space, there is a Schwartz space X universal for the sequence {Xn}. Indeed, let X be the space of all sequences x = {xn}, xn a Xn with the F-norm
IIx11= n=1
1 Ilxnlln 2 n 1+I1xn11n
Chapter 6
274
where I Ix! In denotes the F-norm in X,,.It is easy to verify that X is a Schwartz-
space and that it is universal and co-universal for all spaces Xn.
6.6. DIAMETRAL DIMENSION
In this section we shall consider another definition of approximative dimension, so-called diametral approximative dimension or briefly diametral dimension (see Mityagin ; 1960, Tichomirov, 1960; Bessaga, Pelczyliski and Rolewicz, 1961, 1963). Let A, B be arbitrary sets in a linear space X. Let B be balanced. Let L be a subspace of X. We write
6(A, B, L) = inf(e > 0: L+eB > A). Let us write 6-n (A, B) = inf6(A, B, L),
where the infimum is taken over all n-dimensional subspace L. Let b(A, B) denote the class of all sequences t = (t.,,} of scalars such
that lmw
6..(A,B) =0.
The following properties of the class 5(A, B) are obvious:
if A' C A and B ) B', then b(A', B') C b(A, B) ;
(6.6.1)
S(aA, bB) = 5(A, B) for all scalars a, b different from 0.
(6.6.2)
Let X be an F-space. Let 0 denote the class of open sets and 9 the class of compact sets. Let
6(X)= U U 6(B, U). UEQQ Beg
The class 6(X) is called the diametral approximative dimension (briefly diametral dimension) of the space X. PROPOSITION 6.6.1. Let X and Y be two isomorphic F-spaces. Then
6(X) D b(Y).
Montel and Schwartz Spaces
275
Proof. The proposition immediately follows from the fact that the classes of open sets and compact sets are preserved by an isomorphism. In many cases diametral dimension is easier to calculate than approxi-
mative dimension. Unfortunately we do not know the answer to the following question : do we have 6(X) C 6(Y) is X is a subspace of an F-space Y? As we shall show later, the answer is affirmative under certain additional assumptions. PROPOSITION 6.6.2 (Mityagin, 1961). Let X and Ybe two F-spaces. Let T be
a continuous linear operator mapping X onto Y. Then
6(X) D 6(Y). Proof The definition trivially implies that, for arbitrary A, B C X and an arbitrary subspace L, 6(A, B, L) > 6(T(A), T(B), T(L)). Since dim T(L) < dim L, this implies 6 (A, B) > T(B))
and
6(A, B) D 6(T(A), T(B)).
The inverse image of an open set is an open set. For any compact set K C Y there is a compact set Ko C X such that T(K0) = K (cf. Proposition 6.5.4). Then
6(X) =AE Uf U 6(A, B) 3 U U 6(T(A), T(B)) BEO AEcf Beo AcXBcX
AcX BcX
D AEj El BEO U 6(A,B)=6(Y) AcY BcY
COROLLARY 6.6.3. If
codimjX
(codim1X= codimjY),
then
6(X) C 6(Y)
(resp. 6(X) = 6(Y)).
PROPOSITION 6.6.4 (Bessaga, Pelczyriski and Rolewicz, 1963). Let
S(X) = U n b(V, u). UE6 VC-6
UcI VcX
Chapter 6
276
Then
6(X) = 6(X). Proof. Let B be an arbitrary compact set and V, U arbitrary neighbourhoods of zero. Since the set B is compact, there is a positive number a such
that B C aV. Then by (6.6.1) and (6.6.2) we obtain b (B, U) C 6 (V, U) for all B and V. Hence
U b(B, u) C n b(V, u).
BEJ
VE(J
BcX
VcX
Therefore
U UeC) BED
S(B,U)c U n 6(v, u), UECJ VC-(J
i.e. b(X) C b(X). Now we shall show the converse inclusion. Let U0 be a neighbourhood of zero and let
{tn}e n 6(V, Uo). VEO
Let {Vk} be a decreasing countable basis of neighbourhoods of zero. For-
mula (6.6.3) implies that lim
to
nr W bn(Vk, Uo)
= 0 (k = 1, 2,...). Then we can
choose an increasing sequence of indices k1, k2, ... such that lim
to
0.
bn(Vkn,Uo)
For every Vk there is a finite Z C Vk such that
6n(Z., Uo) i
b(Vk,,, Uo)
00
Let B = U Z, . Since the set B has a unique cluster point 0, it is compact.
On the other hand, ItnI
bn (B, Uo)
_<
ItnI
bn (Zn, Uo)
_ <
21tn1 b n (Vk,,, Uo)
0.
This means that {tn} a b(B, Uo). Then {tn} e b(X).
As an obvious consequence of Proposition 6.6.4 and (6.6.1) we obtain
Montel and Schwartz Spaces
PROPOSITION 6.6.5 (Bessaga, Pelczynski and Rolewicz, 1961). Let a basis of neighbourhoods of zero in an F-space X. Then
277
be
W
Co
b(X)= V n b(Um, Un). n=1 m=1 COROLLARY 6.6.6 (Bessaga, Pelczynski and Rolewicz, 1961). If X is a Schwartz space, then there is a sequence {tn } tending to 0 such that
{tg} l b (X) .
Proof. If X is a Schwartz space, then there is a basis of neighbourhoods of zero {Un} such that lim 6n(U5 1, Up) = 0 (j = 1, 2, ...). Since we have n_ o0
only a countable number of sequences bn(U5+1i U1), it is not difficult to construct a sequence {tn } tending to 0 such that
x
to
J = 1, 2, ...,
n . an(Uj+1, Ui) - +O°' Proposition 6.6.5 implies that {tx} 0 b(X).
PROPOSITION 6.6.7 (Mityagin, 1961). Let X be a Bo space with the topology given by a sequence of Hilbertian pseudonorms, i.e. the pseudonorms MixiIn
= (x, x)n, where (x, x)n are inner products. Let Y be a subspace of the space X. Then
b(Y) C b(X). Proof Let U. = {x: 1Ix1jn < an}, where the positive numbers a are chosen in such a way that {Un} constitutes a basis of neighbourhoods of zero. Since the norms are Hilbertian, for every n and q there is a continuous linear operator Tp,q mapping X onto Y such that Tp,q(Up) = Up n Y, Tp,q(Uq) = Uq n Y.
By similar arguments to those used in the proof of Proposition 6.6.3 we obtain 6(Up n Y, Uq n Y) = 6(Tp,q(UP), Tp,q(Uq)) C b(UP, Uq) Since the sets Vp = Up n Y constitute a basis of neighbourhoods of zero in Y, 00
00
a(Y) = q=1p=1 U 1 1 b(Vp, Vq) C
up 00
co
1
b(Up, Uq) = b(X).
Chapter 6
278
We do not know whether the converse fact to Corollary 6.6.6 holds, i.e. the following question is open :
Problem 6.6.8. Let X be an F-space. Suppose that there is a sequence ft) tenping to 0 such that {tX } 0 6(X). Is X a Schwartz space?
6.6.8 is strictly connected with Problem 6.6.9. Let (X, II II) be an F-space and let A be a closed bounded
set in X. Suppose that, for each starklike open set U, the sequence {6 (A, U)} tends to 0. Is the set A compact?
The answer is positive for locally pseudoconvex spaces and spaces N(L(Q, E , y)) LEMMA 6.6.10 (Turpin, 1973). Let (X, 11 II) be an F*-space. Suppose that for each neighburhood of zero U there is a neighbourhood of zero V such that for each finite dimensional subspace L and for all a > 0 there is a finite set H C V n L such that
V r) L C H+aU.
(6.6.3)
Then each bounded set B such that lim 6.(B, U) = 0 for all open balanced w
enighbourhoods of zero U is totally bounded.
Proof. Let W be an arbitrary neighbourhood of zero. Let U be a balanced neighbourhood of zero such that
U+U C W. By our hypothesis there is a finite-dimensional space L such that B C L+
-I- U. Hence B c (B+ U) n L+ U. Without loss of generality we may assume that U n L is bounded. Hence for sufficiently small a > 0
Bc(B U)nL+Uc1Vn L+ U. a
(6.6.4)
Thus by (6.6.3) and (6.6.4) there is a finite set H such that
H+ W. B C 1a H+U+U C 1 a
(6.6.5)
Montel and Schwartz Spaces
279
Therefore, for all a' > 0, there is a finite set H such that
B C H+a'W.
(6.6.6)
REMARK 6.6.11. In Lemma 6.6.9 the hypothesis that H is finite can be replaced by the hypothesis that it is totally bounded.
Indeed, if His totally bounded, then for each a' > 0 and a neighbourhood of zero W there is a finite set Ho such that
H C H0+a' W.
(6.6.7)
Thus by (6.6.6) and (6.6.7) we obtain
B C Ho+a'W+a'W. PROPOSITION 6.6.12. Let X be an F*-space without arbitrarily short lines. Then each bounded set B such that lim 6.(B, U) = O for an arbitrary baln-. oo
anted neighbourhood of zero U is totally bounded.
Proof The proposition follows immediately from Remark 6.6.11 and Lemma 2.4.6. As an immediate consequence of Proposition 6.6.12 we obtain
PROPOSITION 6.6.13. Let X be a locally bounded space. Let B C X be a bounded set such that lim 8n(B, U) = 0 for each open balanced neighn->co
bourhood of zero U. Then the set B is totally bounded. PROPOSITION 6.6.14. Let X be an F*-space with a topology given by a sequence of F-pseudonorms {II IIn} (see Section 1.3). Suppose that for each
n there is an an > 0 such that, for all x such that IIxIIn:0, sup IItxIIn > an.
(6.6.8)
tER
Let B be a bounded set such that, for each open balanced set U, lim Sn(B, U) = 0. Then the set B is totally bounded. n_C0
Proof. Let U be an arbitrary neighbourhood of zero. Then there are n and
Chapter 6
280
a number a' > 0 such that {x: IIxIIn
(6.6.9)
Let Xn° = {x e X: 11x1 j. = 0}. Let Xn = X/Xn denote the quotient space. The F-pseudonorm II In induces an F-norm in the space X. By (6.6.9) Xn
is without arbitrarily short lines. The bounded set B induces a bounded set B. = {[x] : x e B} in X. Similarly, U induces a neighbourhood of zero in Xn. Thus, by Proposition 6.6.12, for each a > 0 there is a finite set H such that
B C B+Xn C H+Xn+a(U+Xn) = H+Xn+aU = H+aU. By Propositions 6.6.13 and 6.6.14 we trivially obtain
PROPOSITION 6.6.15. Let X be a locally pseudoconvex space. Let B be a bounded set such that, for each open balanced set U, lim 6.(B, U) = 0. n- co
Then the set B is totally bounded. LEMMA 6.6.16 (Turpin, 1973). Let X be an F*-space. Suppose that there is
a basis of neighbourhoods of zero {Um}, such that, for m = 1, 2, ..., n aUm is a linear space and there is a neighbourhood V. such that a>0
Vm+naUrC Um. a>O
Then each bounded set B such that lim Sn(B, U) = 0 for each balanced n^ oo
neighbourhood of zero U is totally bounded.
Proof. Let {Vm,n} be a sequence of neighbourhoods of zero such that Vm, 0 = Vm and Vm,n+1+ Vm,n+1 C V.'.-
Let Wm,n = Vm,n + n aUm, a>O
Of course,
naWm,n=naUm
a>O
a>0
(6.6.10)
Montel and Schwartz Spaces
281
and
Wm,n+1+Wm,n+1 C Wm,n
By the Kakutani construction (see Theorem 1.1.1), the sequence {Wm, n} induces an F-pseudonorm II IIm and, by (6.6.10), formula (6.6.8) holds. Of course, the topology determined by the sequence of F-pseudonorms
{II IIm} is equivalent to the original one. Therefore Proposition 6.6.14 implies the Lemma. LEMMA 6.6.17 (Turpin, 1973). Let X be an F*-space such that for each neighbourhood of zero U there is a neighbourhood of zero V with the following property :
if there are sequences {sn,;}, i = 1, 2,..., k such that sn,i > 0,
lim sn,i = +00, lim n-+
n-->ao
Sn,i-1
= oo, i = 1, 2, ... , k and
Sn,d
k
sn,;e{ a V, then lin(e1, ..., ek) C U.
(6.6.11)
4=1
Then the hypotheses of Lemma 6.6.10 hold. Proof. Let U and V satisfy condition (6.6.11). Let W be a balanced neigh-
bourhood of zero such that W+ W C V. Let L be a finite-dimensional subspace. We shall show that there is a bounded set H C L such that
W n L C H+ n aU.
(6.6.12)
a>O
Suppose that (6.6.12) does not hold. Then there is an unbounded sequence {xn} C Wn L such that, for each subsequence {yn} of the sequence {xn} and for each bounded set HC L.
{yn}tH+naU. a>0
(6.6.13)
The existence of such a sequence follows from the fact that L is finitedimensional. We shall show that (6.6.13) does not hold. Namely, we shall show that each unbounded sequence {xn} C Wn L contains a subsequence {yn} C W n L such that there is a bounded sequence {zn} such that Yn E zn+
n aU. a>O
Chapter 6
282
Since {x,,} is unbounded and the space L is finite-dimensional, we can find e, a L, e1 54 0 and a subsequence {y.} of the sequence {xn} such that yn = sn,1e1+z , where sn,1-moo and ?n -a0 and, moreover {zn} belongs to a subspace Sn,i
L1 of the space L such that el 0 L1. Either the sequence {z} is bounded or it is unbounded. In the second case we repeat our process. Finally we can choose a subsequence {yn} of the sequence {xn} which can be represented in the form k'
Yn =
1
Sn,iei+zn,
(6.6.14)
i=1
where k' < k, sn, i > 0, sn, i-*oo, Sn,i-i -aoo and {zn} is a bounded seSn, i
quence.
Since {zn} is bounded, there is a number b, 0 < b < 1, such that {b zn} C Wn L. The sequence {yn} is a subsequence of the sequence {xn} ; thus {b yn}C WnL. Therefore, by (6.6.14), k'
bfsnieiE V i=1
and, by (6.6.11), lin (e1, ..., ek) C U. This implies that lin (e1, ..., ek) C n a U and (6.6.13) does not hold. a>o
Thus we have (6.6.12) and since L is finite-dimensional the hypotheses of Lemma 6.6.10 hold. THEOREM 6.6.18 (Turpin, 1973). In the space N(L(Q, L', u)) each bounded .set B such that lim 6.(B, U) = 0 for each balanced neighbourhood of zero U is totally bounded. Proof. We shall show that the hypothesis of Lemma 6.6.17 holds. Let E be
an arbitrary positive number. Let f,i ..., fk be measurable function s. Suppose that there are sequences {s,,,i}, i = 1, ..., k such that sn,i >0,
Montel and Schwartz Spaces
sn,i-> CIO,
sn,i_i Sn, i n
283
-*oo and for all n k
PNi=1 (f Sn,ifi)
f J
n
N (i=1 f sn,Ji
d-p < E.
(6.6.15)
Let k
A = U {t: f;(t)#0}. i=1
k
For each t e A, I sn,i f (t) I tends to infinity. Thus, by (6.6.15) and the Fatou lemma,
f
sup
N(u)dµ-<e,
A o
i.e.,
sup
0
N(u) < E
(6.6.16)
p(A)
k
By (6.6.16), for each linear combination f =
ci f , i=1
PN(f)<S <2e. The hypothesis of Lemma 6.6.17 and the above formula imply the theorem.
Turpin (1973)'showed in fact a stronger result, namely that Theorem 6.6.18 is valid for some generalizations of spaces N(L(S2, E, p)). Suppose we are given a space M(ana,n), where m are positive integers and n are non-negative integers (see Example 1.3.9). We say that the space
M(am,n) is regular if, for m < m', the sequence am,n/am,,n is non-increasing. PROPOSITION 6.6.19. If a space M(am,n) is regular, then S(M(am,n ))
_ {{tn}: lim to n-oo
aq,n ap,n
= 0 for some p and all q}. I
Proof. Let Up = {x: IIxjj < 1}. Since the space M(am,n) is regular, we have, for q > p, Sn_1(Uq, Up) = ap,n/aq,n and Proposition 6.6.5 trivially implies the proposition.
Chapter 6
284
COROLLARY 6.6.20. Let {a,}, {bn} be two sequences of reals tending to in-
finity. Then b (M(an )) = M(an) b (M(bn "na)) = M+(bn h/m) = {{tn}: tnbn 1I' --> 0 for some m}. COROLLARY 6.6.21 (Bessaga, Pelczynski and Rolewicz, 1961). There is an
infinite-dimensional Bo space X which is not isomorphic to its product by the one-dimensional space. Proof. Let X = M(expm22°+'). The sequence belongs to p b(M(expm221)), but it does not belong to {exp(-22°+111)}
b(M(expm22"+')).
Now we shall introduce a class of sequences in a certain sense dual to the class b(X). We denote by 6'(A, B) the set of all sequences {tn} such that lim to bn(A, B) = 0. The class b'(A, B) has the following properties :
if A' C A, B' D B, then b'(A', B') j b'(A, B),
(6.6.1') b' (aA, bB) = 6'(A, B) for all scalars a, b different from 0. (6.6.2')
Let
b'(X)=KE9 n nUEQb'(K, u). By a similar argument to that used for b(X) we find that b'(X) is an invariant of linear codimension, which means that if
codim1X
(codimjX= codimjY),
then
b'(X) j b'(Y)
(resp. b'(X) = b'(Y)).
Let
s'(X) = U U b'(V, U). UEG VEO
By a similar argument as to that used in the proof of Proposition 6.6.4 we obtain PROPOSITION 6.6.22. b (X) = b'(X).
As a consequence of Proposition 6.6.22 we obtain the following proposition, in a certain sense dual to Proposition 6.6.20. ,
Monte] and Schwartz Spaces
285
PROPOSITION 6.6.23. If a space M(am,n) is regular, then
S'(M(am,n)) _ {{tn}: for all p there is a q such that tn apn aq,n
0 }JI.
COROLLARY 6.6.24. Let {an}, {bn} be sequences of reals tending to infinity. Then '
6' (M(bn "m)) = M(bn 11 ),
6'(M(an)) = M-(an) = {{tn}: tnan'n -> 0 for some m}. In the same way as in Corollary 6.6.7 we can prove COROLLARY 6.6.25. Let X be a B, -space with the topology given by a se-
quence of Hilbertian pseudonorms. Let Y be a subspace of the space X. Then
6'(X) j 6'(Y). The following, natural question arises : when does the equality of diametral approximative dimensions imply isomorphism? We shall show that this holds for an important class of spaces called Kothe power spaces. Let an be a sequence of positive numbers tending to infinity. The space M(an) is called a Kothe power space of the infinite type, the space M(a,-,uIm) is called a Kothe power space of the finite type.
To begin with, we shall show that two Kothe power spaces of infinite type (of the finite type) induced by sequences {an} and {;n} are equal as the sets if and only if there are two positive constant A, B such that
A < logan
(6.6.17)
logan
Indeed, x = {xn} a M(an), (M(an11m)) if and only if
log jxnJ+mlogan --> -oo,
m = 1, 2,...
(6.6.18)
(resp.
logIxnJ-
1
m
logan -+ -oo,
m = 1, 2,...).
(6.6.18')
Chapter 6
286
If (6.6.17) holds, then (6.6.18) (resp. (6.6.18') holds if and only if
m= 1,2,...
1ogIxnj+m1ogan-*-oo, (resp.
logIxn1- 1 logan -* -oo, M
m =1,2,...).
Thus (6.6.17) implies that the spaces as the sets are equal. In a similar way we can show that the sets M -(an) and M-(am) (resp. M+(anlim) and M+(anhlm) , where M-(a,,) is defined in Corollary 6.6.24
and M+(anlIm) is defined in Corollary 6.6.20, are equal if and only if (6.6.17) holds. Thus, by Corollaries 6.6.20 and 6.6.24, we obtain COROLLARY 6.6.26. Let X, Y be two Kothe power spaces of the finite type (of the infinite type). Then the following three conditions are equivalent.
X and Y are isomorphic,
(6.6.19.i)
the diametral approximative dimensions of X and Y are equal,
6(X) = 6(Y), 6'(X) = 6'(Y),
(6.6.19.ii)
(6.6.19.iii)
Corollary 6.6.26 shows that KSthe power spaces are distinguishable with respect to the diametral approximative dimension. Dragilev (1970b) described a class of spaces L(an,n) (see Example 1.3.9) distinguishable with respect to the diametral approximative dimension. Let [am,n] be a matrix of positive numbers such that for each p there is a q such that Jim aP,n
,-. aq,n
= 0.
Let N denote the set of all subsequences v = {nk} of the sequence of positive integers. Let
NlP,q,g = jv e N: liminf as.n2aP,nk > 0 nkEV
aq,nk
and let N2
,q,8 =
jv a N: li m sup nkEV
a8,fZaP,nk < aq,nk
+
Montel and Schwartz Spaces
287
Let
N'(L(am,n)) =p u nq n N:,q,s and
N2(L(am,n)) p= n u n q
Np,q,s.
s
The following facts hold N'(L(am,n)) n N2(L(an=,n)) = o, N' (L (am,n)) U N2(L (au+,n))
o.
Moreover, N1(L(am,n)) and N2(L(am,n)) are invariants of isomorphisms inside the class of all spaces L(am,n), which means that if two spaces L(am,n) and L(bm,n) are isomorphic, then NL(L(am,n)) = Nt(L(bm,n)),
i = 1, 2.
We shall introduce an order in Nin the following way v, -3 v2 if almost all elements of v, belong to v2.
The class N'(L(am,n)), i = 1, 2, satisfies one of the following three conditions :
there is a maximal element in Nt(L(a,n,n)),
(6.6.20.1)
there is no maximal element in Nt(L(am,n)),
(6.6.20.2)
the class Nt(L(am,n)) is empty.
(6.6.20.3)
We say that a space L(am,n) belongs to the class. Et,5, i, j = 1, 2, 3, if N'(L(am,n) satisfies condition (6.6.20.1) and N2(L(am,n)) satisfies condition (6.6.20.2). The class E,,, is empty. Dragilev (1970b) showed that no other class is empty. Let E0 = E13 u E3,1 u E1 1. Dragilev (1970b) showed that if two spaces belonging to Eo have the same approximative diametral dimension,
then they are isomorphic. On the other hand, if Et, p e E0, then there are two spaces belonging to the union Et, j u Eo such that they are not isomorphic but they have the same approximative diametral dimension. It is not known whether two spaces L(am,n) and L(bm,n) having the same approximative dimension and equal sets N1 and N2 are necessarily isomorphic.
Chapter 6
288
6.7. ISOMORPHISM AND NEAR-ISOMORPHISM OF THE CARTESIAN PRODUCTS
The results of this section for locally convex linear topological spaces was obtained by Zahariuta (1973). We shall formulate it for metric linear spaces not necessarily locally convex. Let (X, II IIx) and (YII IIy) be two F*-spaces. We recall that a linear
continuous operator T mapping X into Y is called compact if there is a neighbourhood of zero U in X such that the set T(U) is totally bounded. We say that an ordered pair of F-spaces (X, Y) satisfies condition R (briefly (X, Y) e R) if every linear continuous operator T mapping X into Y is compact. PROPOSITION 6.7.1. If Y is a Montel space, then (X, Y) e R for each locally bounded space X.
Proof. Let T be a linear continuous operator mapping X into Y. Let U C X be a bounded neighbourhood of zero. By Theorem 2.1.1 the set
T(U) is bounded. Since Y is a Montel space, the set T(U) is totally bounded. PROPOSITION 6.7.2. Let Y be a locally p-convex space. If (X, Y) e R for all
locally bounded spacesXwith p-homogeneous norms, then Y is a Montel space. Proof. The topology in Y is given by a sequence {ll Iln} of p-homogeneous
pseudonorms. Let A be an arbitrary bounded set in Y. Then there is a sequence of numbers such that
P(y)=supmnllylln<1 n for all y e A. Let (X,p(y)) be the space of such y that p(y) < -boo. Of course p (y) is ap-nomogeneous norm on X. The operator T equal to identity maps X into Yin a continuous way. Observe that A C U = {y: p(y)
1}. Thus A C T(U). Since T(U) is totally bounded, the set A is also totally bounded.
Montel and Schwartz Spaces
289
PROPOSITION 6.7.3. Let X be a Schwartz space. Then, for each locally bounded space Y, (X, Y) e R. Proof. Let T be a continuous linear operator mapping X into Y. Since the space Y is locally bounded there is a neighbourhood of zero UC X such
that T(U) is bounded. Take any neighbourhood of zero WC Y. Since the set T(U) is bounded, there is a positive number a such that aT(U) C W. The space X is a Schwartz space, thus there is a neighbourhood of zero V totally bounded with respect to U, i.e. for each e > 0 there is a such that finite system of points {x1, ...,
V C U ( xi {
U) .
(6.7.1)
Thus
T(V) C U (T(x+ e T(U)) C U i=1
a
(T(xi)-FeW).
i=1
The arbitrariness of e and W implies that the set T(V) is totally bounded.
PROPOSITION 6.7.4. Let X be a locally pseudoconvex (locally p-convex)
space. If (X, Y) e R for each locally bounded space Y (locally bounded space Y with a p.homogeneous norm,) then X is a Schwartz space.
Proof. Take an arbitrary absolutely p-convex neighbourhood of zero U. The set U induces a p-homogeneous pseudonorm II II. Take as X0 the quotient space Xo = Xl{x: IIxii = 0}. The space X0 is a locally bounded space with the norm induced by II II. The canonical mapping 7ca: X->X0 is a continuous linear operator. Thus, by our hypothesis, there is a neighbourhood of zero V such that 7r,(V) is totally bounded with respect to 7ca(U). This implies that V is totally bounded with respect to U. Since there is a basis of absolutely p.-convex (p-convex) neighbourhoods of zero in X, V is totally bounded. COROLLARY 6.7.5. (X, Y) e R and (Y, Z) e R does not imply (X, Z) e R.
Proof. Let X be a Schwartz space. Let Y be an arbitrary Banach space. Then (X, Y) e R and (Y, X) e R, whereas (Y, Y) 0 R.
Chapter 6
290
LEMMA 6.7.6. Let (X, Y) e R. Let XO be a subspace of X and let Yo be a subspace of Y. We assume that XO is complemented, i.e. that there is a continuous projection of X onto X0. Then (XO, Yo) e R.
Proof. Let To be an arbitrary linear continuous operator mapping XO into Yo. Since (X, Y) e R, the operator TOP mapping X into Y is compact. Thus there is a neighbourhood of zero U in X such that TO(P(U)) is totally bounded. Therefore the set TO(XO n U) C TOP(U) is also totally bounded. LEMMA 6.7.7. If (X, Y) e R and X is isomorphic to Y, then X is finite-dimensional. Proof. The lemma is a trivial consequence of the Eidelheit-Mazur theorem (Theorem 6.1.2).
Combining Lemmas 6.7.6 and 6.7.7., we obtain PROPOSITION 6.7.8. Let (X, Y) e R. Let XO be an infinite-dimensional subspace of X. Let XO be complemented. Then XO is not isomorphic to any subspace YO of Y. Let (X, II IIx) and (Y, II IIY) be two F-spaces. A linear continuous operator T mapping X into Y is called a 0-operator (or a near isomorphism) if T(X) is closed and
dim{x: T(x) = 0} = a(T) < +oo
(6.7.2)
dim Y/T(X) = 3(T) < -boo.
(6.7.3)
and
By the index of a 0-operator T we shall mean the difference
x(T) = a(T)-#(T).
(6.7.4)
If there is a 0-operator mapping X into Y we say that the spaces X and Yare nearly isomorpic. The theory of 0-operators in Banach spaces is presented in a fundamental paper by Gohberg and Krein, 1957. The results concerning non-locally convex spaces can be found in Przeworska-Rolewicz and Rolewicz (1968). We shall formulate those results without proofs.
Montel and Schwartz Spaces
291
Let X, Y, Z be three F-spaces. Let T: X->Y and S: Y-->Z be two 0-operators. Then the superposition ST is a 0-operator and we have the following equality for indexes
x(ST) = x(T)+x(S).
(6.7.5)
Formula (6.7.5) for Banach spaces was proved by Atkinson (1951, 1953).
Basing ourselves on the Riesz theory of compact operators on non-locally convex spaces (Wiliamson, 1954), we find that if T: X-*Yis a 0-oper-
ator and S: X-> Y is a compact operator then T+S is a 0-operator and
x(T+S) = x(T).
(6.7.6)
Let T: X->Y be a 0-operator, then there is a b-operator t such that TT = I-}-B,
(6.7.7)
TT = I+ C,
(6.7.8)
where B and C are compact operators. Let X = X1 x X2i Y = Y1 x Y2 be F-spaces. Let T be a continuous linear operator mapping X into Y. Of course T can be represented in the matrix form
T=
Ti,i T1,2J T2,1 T2,2
where Ti, p: X1->Y5, i, j = 1, 2.
Now we shall formulate a lemma proved by Douady (1965) for Banach spaces and by Zahariuta (1973) for locally convex spaces. LEMMA 6.7.8 If T is a 0-operator and (X1, Y2) e R, then the operator Ta
[Tii 0
Ti.zl
is a c-operator. If moreover (X1i Y2) e R then
T2,2 J
x(T) = x(T1,1)+x(T2,2)
(6.7.9)
0
Proof. Since (X1, Y2) a R, the operator S = 0 is compact. Hence I T2,1 0
the operator T-S =
T1, 1
T1, z
0
T2,2
is a 0-operator.
Chapter 6
292
If (X2, Y1) e R, then the operator S1 =
x(T-S-S1) = x([OTi,l
TZ 2J)
O Ti,z is compact and 10 0
= (T1,1)+x(T2,2)
As a consequence of Lemma 6.7.8 we obtain PROPOSITION 6.7.9. Let X = X1 X X2 and Y = Y1 X Y2. Let (X1, X2) e R
and (Y1, X2) e R. Then X is nearly isomorphic to Y if and only if X1 is nearly isomorphic to Y1 and X2 is nearly isomorphic to Y2.
Let X be an F-space. By X M we shall denote an arbitrary subspace of codimension i if i > 0 (observe that all such subspaces are isomorphic) and an arbitrary space X x Z, dim Z = I i 1, if i < 0. THEOREM 6.7.10. Under the hypotheses of Proposition 6.7.9 the space X is isomorphic to the space Y if and only if there is an integer s such that Y1 is isomorphic to X() and Y2 is isomorphic to X2 8>
Proof Sufficiency. If Yl is isomorphic to X(") and Y2 is isomorphic to X2-8), then Y1 x Y2 is isomorphic to X() x XZ 8) and it is isomorphic to X1 X X2.
Necessity. Let T: X1 x X2--> Y1 X Y2 be an isomorphism. Then, by Lem-
ma 6.7.8, T1,1: X1- Y1 and T2,2: X2->Y2 are 0-operators. Then Y1 is isomorphic to X(',) and Y2 is is isomorphic to X. By formula (6.7.9) 0 = x(T) = x(Ti,l)+x(T2,2) = S1+s2 and
Now we shall apply the results given above to a certain class of locally convex spaces. Let X be a Bo-space with a topology defined by a sequence of pseudonorms {11
_
JIk}. Suppose that {en} is a basis in X such that, for each x
M
00
Ixnj 1 JenJIk are convergent for k = 1, 2, ... We
xne,,, the series n=1
n=1
shall call a basis with this property an absolute basis. We say that X e d1 (is of type d1) if there are an absolute basis {en} in
X and an index p such that for each index q there are an index r and N
Montel and Schwartz Spaces
293
= N(p, q) such that
for n > N.
IlenJIq < IlenIIr, IlenjI,
(6.7.10)
We say that X e d2 (is of type d2) if there is an absolute basis {en} in X
such that for each p there is a q such that for each r there is an N = N(p, r) such that
for n > N.
lien II q > I Ienl lp I I enllr
(6.7.11)
Example 6.7.11. Let {an} be a sequence of positive numbers tending to in-
finity. The spaces LP(a;,, ), 0
and Ye d1. If Y is a Montel space, then each continuous linear operator trapping X into Y is compact. Proof. By definition there are absolute bases {en} in X and {fn} in Ysuch that (6.7.11) and (6.7.10) hold. Since the bases are absolute, we may assume without loss of generality that the topology in X (in Y) is given by a sequence of pseudonorms {I I
I Im} (resp. {I
I m})such that for x = Jxn en e X(resp. y =
00
yn fn e Y)
n=1
n=1 00
m= 1,2,...,
IIxlIm = f Ixnl Ilenllm, n=1 (resp. oo
IIyAIM =
Iynl Ilenlim,
m= 1,2,...)
n=1
Let T be a continuous linear operator mapping X into Y. Let hn 00
Co
= T(en) = I ti,n f . Of course for each x = j=1
00
oD
T(x)
xn e,, n=1
YxnT(en) =f xn n=1
00
Chapter 6
294
The continuity of the operator T implies that for each p there is a q = q(p) such that cc
C(P)
=
keaq)Ilp IITII(eekll
sup
T Iti,kllllillp
= sup IIekIIq
< +oo
(6.7.12)
We have assumed that Y is a Montel space. Thus to prove the theorem it is enough to show that the operator T maps some neighbourhood of zero UQ. = {x: IIxIlq. < 1 } into a bounded set.
Since Y e d1, there is a p, such that for each p there are io(p) and p2(p) such that
for i > io(p) (6.7.13) On the other hand, X e d2. Take q = q(p1). Thus there is a qo such that (IIf ll;)2 < Ill{II , Illill ,,
for each q2 there is a ko(g2) such that IIekIIq, >
IIekIIq,
for k > k(q2).
IIekIIq,
(6.7.14)
Take q2 = q(p2). Of course, k(q2) depends implicitly on p. By (6.7.13) and (6.7.14) there is a constant L(p) such that Illillp
(
IIekIIq.
IP2
(6.7.15)
Ilekllq,llekllq, /
fori,k= 1,2,... Hence, by the Cauchy inequality, we obtain 00
Ill{IIp ti.k I IIekIIq. 1/2
L(p)
i11=1d
(
Iti,kl
HJ lIp,
1/2
I ti kI
IIekIIq, /
IIsIIPa
IIekIIq: I 1/2
IIIiIIP,
IekII, I
11/z
(y' Iti,kl Illdlla IIekIIq,/
liT(ek)IIP
< M(p), where M(p) =
IIekIIq.
= 1, 2, ... By the form of the pseudonorms, this implies that II T(x)l < M(P) I IxIIq.
Therefore the set T(U) is bounded.
L(p)C(p1)112C(P2)1/2,
p
Montel and Schwartz Spaces
295
COROLLARY 6.7.13. If X is a Kothe power space of the finite type and Y is
a Kothe power space of infinite type, then each continuous linear operator mapping X into Y is compact. Later we shall give an example of a Kothe power space of infinite type, which is a subspace of a Kothe power space of the finite type (see Section 8.3).
Now following Zahariuta (1973) we shall introduce a new topological invariant. Let E1, e2 be two classes of spaces such that (X1, X2) C- R for X, 'c- e1 and X2 e E2. With each X which is a product X = Xl x X2 we shall associate the set I'(X) of all pairs (S(XlBt), 6(XZ' )), s = 0, ±1, ±2, ..., where 6(Y) denotes the diametral approximative dimension of Y. As a consequence of Theorem 6.7.10 we imediately obtain THEOREM 6.7.14. I'(X) is an invariant of an isomorphism, i.e. if X = Xl X X2i Y = Y1 X Y2, X1i X2 e el, X2, Y2 a E2 and the spaces X and Y are isomorphic, then T(X) = I'(Y). If, moreover, the classes e1 and 62 are such that Z1, Z2 e e,, i = 1, 2, and 6(Z1) = 6(Z2) implies the isomorphism of Z, and Z2, then I'(X) = .1(Y) implies that X is isomorphic to Y. As a consequence of Theorem 6.7.14 and Corollary 6.6.26 we obtain
PROPOSITION 6.7.15 (Zahariuta, 1973). Let X, X1 be two Kothe power spaces of finite type and let Y, Y1 be two Kdthe power spaces of infnite type.
If I'(X x Y) = I'(X1 x Y1) then the space X X Y is isomorphic to the space X1 X Y1.
Chapter 7
Nuclear Spaces. Theory
7.1. DEFINITION AND BASIC PROPERTIES OF NUCLEAR SPACES
Let X be an F*-space. The space X is called nuclear if, for any neighbourhood of zero U, there is a neighbourhood of zero V such that lim nbn(V, U) = 0,
(7.1.1)
n-00
the definition of bn(V, U) being given in Section 6.6. PROPOSITION 7.1.1. If a space X is nuclear, then for any neighbourhood of
zero U and for any positive integer k there is a neighbourhood of zero V such that lim nkbn(V, U) = 0. (7.1.2)
The proof of Proposition 7.1.1 is based on LEMMA 7.1.2 (Mityagin and Henkin, 1963). For arbitrary U, V, WC X
bn+m(W, U)
(7.1.3)
provided bn(W, V) and bm(V, U) are finite. Proof. Let
a=bm(V, U),
b=bn(W, V).
Let a be an arbitrary positive number. Then, by the definition of bm(V, U)
and bn(W, V), there are subspaces Ll and L2, dimL1 = in, dimL2 = n, such that
V C Ll+(a+e) U,
W C L2+(b+e) V.
Nuclear Spaces. Theory
297
Therefore
W C L1+L2+(a+e)(b+s) U. Since dimL1+L2 < n+m, bn+m(W, U) < (a+e)(b+e). The arbitrariness of a implies the lemma.
Proof of Proposition 7.1.1. Let U be an arbitrary neighbourhood of zero. Since X is a nuclear space, there is a system U = Vo, V1, ..., Vk of balanced neighbourhoods of zero such that limnSn(VV, Vi_1)=0 (i = 1,2,...,k). Then, by formula (7.1.3), lim nkbkn(Vk, U) = 0.
(7.1.4)
Since bm(Vk, U) is monotonic,
0 < mkbm(Vk, U) < m
E(k)
k
[E(.)kE()(vk, U)1->0, J
where, as usual, E(a) denotes the greatest integer not greater than a. This implies the proposition. Further on, other equivalent definitions of nuclear spaces will be given. PROPOSITION 7.1.3 (Ligaud, 1973). Let X be a nuclear F*-space. Then its completion X is also nuclear. Proof. Let U be a neighbourhood of zero in X. Let U1 be a neighbourhood
of zero in X such that Ul+ U1 C U. Let V = U1 n X. V is a neighbourhood of zero in X and VC U1i where V is a completion of V. Since X is nuclear, there is a neighbourhood of zero W, WC X, such that bm(W, V) < m+1 I. By the definition of bm, there is a subspace L, dimL = m, 1
such that WC L+ m+ 1 V. Hence, for the completion W of W we have W C L+ m+1 V+ m+1 U1 C m+1 U+L. Thus bm(W, U) <
1
m+1 and this implies the nuclearity of X.
Chapter 7
298
THEOREM 7.1.4 (Ligaud, 1971). Every locally pseudoconvex nuclear space is locally convex.
The proof is based on several lemmas. The set n
of
I'P(A)[x:xatxt,xteIatIP<1,n= 1,2,...}. t=1
ti=1
is called the absolute p-convex hull of the set A. Observe that I'p(I'p(A)) = I'P(A). If 11 11 is ap-homogeneous norm, then
I'p({x: IIx!I < r}) = {x: IIx!I < r}. LEmtn1A 7.1.5. Let X be an n-dimensional space. For any set A C X,
f1(A) C np-1Pp(A). n
n
Proof. Let x e f1(A), i.e. x = E atxi, where 27 jail < 1. Observe that i=1
i=1 IailP)-1'xefp(A).
y i=1
Hence n
r1(A) C
maxll atIP)'1P:
Iatl < 11rp(A)
r
t=1
i=1
= np-1Pp(A). LEmtiA 7.1.6 (Auerbach, 1935). Let (X, II II) be an n-dimensional real Banach space. Then there are elements e1, ..., en of norm one and functionals f1i ..., fn of norm one such that 1
fi(ej) =
10
fori=j, for i
j.
(7.1.5)
Proof. Let x1i ..., xn be elements of norm one and let vol (x1, ..., xn) denote the volume in the Euclidean sense of a parallelepiped with the vertices (e1x1i ..., enxn), where et = 1 or -1 (i = 1, ...). Let e1, ..., en be such elements of norm one that
vol(el, ..., en) = supvol(x1, ..., xn).
Nuclear Spaces. Theory
299
Since the unit sphere in an n-dimensional space is compact and the volume
is a continuous function of x,..... x,,, such el, ..., e exist. Let Hi be a hyperplane passing through the point ej and parallel to vectors e,, ..., ei+,, .. ., en. The hyperplane Hi does not have common points with the interior of the unit ball in X. Indeed, suppose that Hi
has common points with the interior of the unit ball. Then there is a point ei of norm one such that Hi separates 0 and ei and moreover ez
Hi. Therefore
vole
)
(
)
which contradicts the definition of e,, ..., en.
Let fi be such a functional that {x: fi(x) = 1} = Hi. Since Hi do not have common points with the interior of the unit ball, the functionals fi have norm equal to one. The definition of Hi implies (7.1.5). LEMMA 7.1.7. Let (X, II II) be an n-dimensional complex Banach space.
Then there are a system of elements e,, ..., e and a system offunctionals ,f, such that Ileill = Iifill = 1 (i = 1, 2, ..., n) and (7.1.5) holds. .fig
Proof. Let x,, ..., xn be arbitrary linearly independent elements of X. Let X, be a real space spanned by x,, ..., xn. By Lemma 7.1.6 there are in X, elements e,, ..., en and real-valued linear functionals g,..... gn such that IIeill = Ilgili = 1 (i = 1, 2, ..., n), and (7.1.5) holds. Let h5 be a real-valued norm-preserving extension of gj on the whole space X considered as a real 2n-dimensional space. Let fi(x) = hj(x)-ihj(ix), j = 1, 2, ..., n. The functionals f1 are linear (see Corollary 4.1.3). In the same way in the proof of Theorem 4.1.5 we can show that II f I I = 1 (j = 1, 2, ..., n). By the definition of fp, formula (7.1.5) holds. LEMMA 7.1.8. Let X be an n-dimensional space. Let A be a balanced closed convex set. Then there are points e,, ..., en e A such that
A = I',(A) c nl'({e,, ..., e.}). Proof. Without loss of generality we may assume that X = lin A. Since A is convex and balanced, it induces the norm IIxII = inf {t: x/t e A}. By Proposition 7.1.6 (or 7.1.7) there are elements e,, ..., en and functionals f,, ...,fn of norm one such that (7.1.5) holds.
Chapter 7
300
Observe that
A C {x: If(x)I <1,i= 1,2,...,n}C n
n
Ifi(x)I <
C {x:
n}
= n {x:
i=1
Ifs(x)I <
1}
i=1
= nF1({e1, ... , en}) .
LEMMA 7.1.9. Let X be an n-dimensional space. For any set A C X, there are e1, ..., en a F'p(A) such that
A C nPI'p(e1i ..., en). Proof. Of course, A C F'1(A). Thus, by Lemma 7.1.8, there are el, e P1(A) such that
..., en e
A C nrp({ei..... en}). By Lemma 7.1.5, el, ..., e' e nP-'I'p(A). Let
ei = n pei,
i= 1,2,...,n.
Then et a F'p(A), i = 1, 2, ..., n and, by Lemma 7.1.5,
A C n nP-'Pp({ei, ... , en}) = n''PPp({ei , ... , e;}) a
1
s
= nn F'p({e1, ..., en}) C nPI'p({e1, ..., en}). LEMMA 7.1.10. Let (X, 11 I) be a locally bounded space with a p-homogeneous norm II II. Let A be a balanced bounded set in X. If there is an n-di-
mensional space L such that
A C L+SU,
(7.1.6)
where U is a unit ball in X, U = {x: llxl I < 1}, then for all8' > b there are el, ..., en a rp(A) such that 9
Y
8
A C 2PnP6'U+nPPp({el, ..., en}).
Proof. Let A' = {x a L: dist(x,A) < 6'}, where dist(x,A) = infllx-yiI yEA
The set A' is bounded and, by (7.1.6), A C A'+b'U. By Lemma 7.1.9 there are e', ..., en e I'p(A') such that A' C
Nuclear Spaces. Theory
301
C n2Ipl'(p{e'1, ..., en}. From the definition of an absolute p-convex hull we can represent e in the form n
ei
ai,jxi,j,
I ai,jI p < 1.
xi,j e A',
9=1
i=1
Let yi, j e A be chosen so that Ilyi,j-xi.jll < S'. Let n
ei = f aijyi.j j=1
Then n
Ilei-eill <
j=1
(7.1.7)
Iai,jlp lxi,j-yi,jll <S'.
r
Let x =
ai ei e I'p({e1, ..., en}). Then for y =
ai ei E I'p({e,, .. . i=1
e=1
.., en}) by (7.1.7), IIx-yII < S'. This implies I'p({ei, ..., e,'}) C 1'p({el, ..., en})+h'U. Finally, A C n2/P(I'p({el, ..., en})+S'U)+8'U C nZ1'Pp({e,, ..., en})+22fpn2/p8'U.
C
LEMMA 7.1.11. Let (X, II II) be a locally bounded space with a p-homogeneous norm 11 11. Let U = {x: I IxHI < 1} denote the unit ball in X. Let A be
a balanced set in X. If there is a > 0 such that lim n4a+PSn(A,
U) = 0,
n-* co
then there is a sequence {xn} such that lim najIxnII = 0
n-i co
and A C I'1({x1, x2, ... }) .
Proof. Let {sn} be such a sequence of positive numbers that
lim nksn = 0, n-. co
k = 1, 2, ...
(7.1.8)
Chapter 7
302
Let
dn(P, U) = bn(P, U)+sn for every set P. Let mn = 22". We shall construct by induction a sequence of finite systems {xn, t, 1 < i < m,,} and a sequence of sets {Bn} in the following way, We put Bi = A. Suppose that the set B. is defined. Then by Lemma 7.1.10
we can choose a system of points {xn, {, 1 < i < mn} such that xn, t e rp(B,,) and B. C m l'I'p({xn,i,
..., xn,m"})+22/Pmxi/Pdm"(Bn, U)U.
(7.1.9),,
Let
Bn+1 = [B.-mnlvl p({Xn,i, ..., xn,m"})l (7.1.10),
2/pmn 2n'pdm"(Bn, U) U.
Of course, by (7.1.9), and (7.1.10)., Bn C Bn+i+mn/prp({Xn,1, ..., xn,m"}) Since bk(rp(A), U) = bk(A, U), we have, by (7.1.10),, bk(Bn+1, U) < bk(Bn+n2/9Bn, U)
22/Pn21pbk(Bn, U).
Then by induction
n-
n-1 1
I m,lpbk(A, U).
(7.1.11)
n- n-1 dk(Bn, U) < 22 P 11 m,lPdk(A, U).
(7.1.12)
bk(Bn, U) < 22 P
1
Hence i=1
Now take any positive integer r. We can represent r in the form
r = mo+...+mn-i+i, where we put mo = 0 and 0
i < m.. Let
n 2/p zr=2mn xn,i.
We shall show that A C r1({z1i z2, ...}). Let x e A. Basing ourselves on formulas (7.1.10)., we can represent x by induction in the form
X = 2 tj+...+ 2n_i to-i+Yn,
Nuclear Spaces. Theory
303
where i
2
p
t{ E 2 m i IP((X{,1, ..., Xd,mi})
and yn E B. Observe that J
2 to+...+
2n-1
to-1 E F21*1, Z2, ...})
r1({Z1, ...})
and
Yn E dmn(Bn, U)U
and by (7.1.12) n
mi/Pdmn(A,
22n/P
yn e
U)U,
i=1 n
But ]I m2 t=1
22+...+2^ = 22n+1-1
=
< 22,t+1 -
171n,
Hence 22IPmnlPdmn(A, U)U C mn/Pdmn(A, U)
yn c-
and, by (7.1.8) and the definition of Sn, yn- O. Therefore X E 1'1({z1, z2, ...}). Now we ought to investigate the converge nce of razr. By definition
r aZr = (m0+. .. +mn-1 +l )a2ninn Xn,d E (m0+ ... +mn-l f i)a2nmr2,IPI'p(B..) .
By (7.1.10).. and (7.1.12), putting k = mn_1 we obtain n-1
razr E
namn22(n-1/P)
J7
mil Pdmn
,(A, U)U
i=]
U5 C mnPldmn_,(A, U)U
and, by (7.1.8), razr tends to zero.
Cl
Proof of Theorem 7.1.4. Let X be a nuclear locally pseudoconvex space, with topology determined by a sequence ofpl-homogeneous pseudonorms II II1. Let Al = (x: IIxII1 < 1). We shall show that each Al contains an
open convex set. Observe that, by the nuclearity of the space X, for each j there is an index k such that lim nl3IP,dn(A1+k, A1)
a-.
= 0.
Chapter 7
304
Let X? = {x: IIxjI; = 0} and Xp = X/X,. The pseudonorm II Ill induces a pp-homogeneous norm on X1. Observe that
i= 1,2,...
As=At+X°,
Putting A = Ap+k+XX and U = A1, by Lemma 7.1.11 we find that there is a sequence {zn} of elements of Al such that lim n2lp'll znll i = 0 and
(7.1.13)
Al+k+X, C 1'i({zi, z2, ...})+X; .
Since II Its is pj-homogeneous by (7.1.13), there is an M > 0 such that
I7i({z1, z2, ...}) C MAi. Therefore AJ+k C M I'1({Z1, z2, ...}) C A5
and the set
IntI'1({z1iz2, ...}) is the required open convex set.
11
M The existance of non-locally convex nuclear spaces follows from PROPOSITION 7.1.12. A space 111"I (see example 6.4.7) is nuclear if and only
if lim supp logn < +oo,
(7.1.14)
is the sequence obtained from the sequence {pn} by ordering it in a non-increasing sequence. where
Proof. Let
Kr={x:Ilxli
By a simple calculation we obtain for r < s 6n(Kr, Ks) =
r s
1Pn
)
.
Nuclear Spaces. Theory
305
Hence
r p`nlogn+log s nbn(Kr, K,,) = exp
(7.1.15)
PC.
r
pnlogn+log-s (7.1.16)
pn
and by (7.1.15) the space X is nuclear. Conversely, if X is nuclear by (7.1.15), formula (7.1.16) holds for a certain r° < s. Thus (7.1.14) holds.
Another example of a non-locally convex nuclear space was given by Fenske and Schock (1970). The above-mentioned examples of non-locally convex nuclear spaces have the property that the dual spaces are non-trivial. Ligaud (1973) constructed an example of a nuclear space without non-trivial linear continuous functionals. Example 7.1.13 (Ligaud, 1973) Let E be the linear subset of the space L°[0,1] spanned by characteristic functions of intervals with rational ends. The algebraic dimension of the space E is countable. Let V be an arbitrary neighbourhood of zero in E in the topology of L°[0,1]. Observe that
E = r1(
>o
V).
Indeed, for each e > 0, every function belonging to L°[0,1] can be represented as a finite sum of functions with supports contained in intervals of measure less than e. Now we shall introduce a new topology on E. We shall construct it in the following way. Let {e1,e2, ...} be a Hamel basis in E. Let Vn be a sequence of balanced neighbourhoods of zero in the primal topology such that Vn+1+ Vn+1 C Vn .
Chapter 7
306
We define by induction new neighbourhoods Wn as follows : let Wn = V. and let Wn+1 =
n 1
m2+
Wn+Lm
1
m=o
1
where Lm = lin({el, ..., em}). The neighbourhoods Wn define a new topology on E. By induction we trivially infer that the sets Wn are balanced, W.+1+ WP+1 C Wn,
ifn
Wn CWn W.' CWW
Hence Wmax(p,P') max(n,n')
WP n n WP n'
Now we shall show by induction that the sets Wn are absorbing. Of course, Wn = V, is absorbing. Suppose that Wn is absorbing. Let x E E. Then there is an m0 such that, for m > m0, x e Lm. Since Wn is absorbing for m < m0, there are Am > 0 such that
.lmxem+lWn. Let A = min(1, A1, Ax e
m+ 1
..., Am. ). Then
Wn+Lm
for all m. Thus Ax e Wn+l and Wn+1 is absorbing. Therefore the family {Wn, n = 1, 2, ..., p = 0,1, ...} defines a linear topology on E stronger than the primal topology. The family Wn is countable, and thus the topology in question is metrizable. Since for all m > 0 W.+1
C m-{ 1
l
WP +L 1
an\W'P'+1' Wn) < M+1
Therefore the space E with this new topology is nuclear. By the defini-
Nuclear Spaces. Theory
307
tion of WP
n=1,2,...,p=0,1,...
n2Vn.c W' z>o
Hence
r1(
)
I'1(n 2V)=E. A>o
Thus there are no non-trivial linear continuous functionals on E. The space E is not complete, but its completion E is, by Proposition 7.1.3, also
nuclear, and of course there are no non-trivial linear continuous functionals on E. The example described above has been specially constructed to show the existence of nuclear spaces without non-trivial linear continuous functionals. Burzyk (1980) investigating the completeness of the Mikusinski operator field, has introduced in a natural way an algebra which is a Montel space without non-trivial linear continuous functionals. It is not clear whether the space constructed by Burzyk is also nuclear. Kalton (1979) has proved that, if a strictly galbed infinite-dimensional F-space Xis not locally bounded, then it contains an infinite-dimensional
nuclear locally convex space. For a locally-convex X this has been proved in Bessaga, Pelczyriski and Rolewicz (1961). The notion of nuclearity can be refined by the notion of 2-nuclearity.
Let 1 be a linear space of sequences of numbers. We assume that e2 s imply {,qn} e 2) additive (i.e. if {rn} e 2, for C2.-1 = fin, 42n = ?1n the sequence C,,( ,,)belongs to 2 for all bijections it (n) of positive integers onto themselves) and decreasing rearangement invariant (i.e. if E A then the sequence {En} obtained from the sequence is non increasing, by a rearangement such that also belongs to 2). We say that an F*-space X is 2-nuclear if, for each neighbourhood of U)} e 2. zero U, there is a neighbourhood of zero V such that A is normal (i.e.
Taking as 2 _
e A, Ipnn1
sup 1 nj nk < +oo, k = 1, 2, ...} we obtain
nuclear spaces. Investigations of A-nuclear locally convex spaces have been carried out in Ramanujan (1970), Dubinsky and Ramanujan (1972), Dubinsky and Robinson (1978), Moscatelli (1978), Dubinsky (1980).
Chapter 7
308
7.2. NUCLEAR OPERATORS AND NUCLEAR LOCALLY CONVEX SPACES
In this section we shall begin investigations of nuclear operators in Banach spaces. Let (X, II IIx) and (Y, II IIy) be two Banach spaces. By Bx,By we shall denote the unit balls in X and Y, Bx = {x e X: IIxilx < 1}, By = {y e Y: Ilyll < 1). Write d%(T) = Sn(T (BB) , By) for any linear continuous operator T mapping X into Y. It is obvious that an operator T is compact if and only if d a(T)->O. Let three Banach spaces X, Y, Z be given. Let B be a continuous linear operator mapping X into Y and let A be a continuous linear operator mapping Y into Z. Then by Lemma 7.1.2 do+m(AB) = dn(A)dm(B).
Let X be a Bo space and let {IIxII{} be an increasing sequence of homogeneous pseudonorms determining the topology. Let
x° = {x: IIxIIti = 0} and let Xi be the quotient space X/X°. The pseudonorm IIxlkk induces a homogeneous norm in the space Xi. Since no confusion will result, we shall denote this norm by the same symbol IIxII{ Of course, Xi with the norm IIxlli is a normed space. Since the sequence of pseudonorms {IIxII{} is increasing, there is a natural continous embedding of the space X;+1 into the space X1. We shall denote it by Ti. The definition of nuclear spaces implies PROPOSITION 7.2.1. Let X be a Bo-space. The space X is nuclear if and only if there is an increasing sequence of pseudonorms {IIxjli} determining a to-
pology such that
limndn(Ti) = 0
(i = 1, 2, ...).
(7.2.1)
n. co
A continuous linear operator T mapping a normed space X into a normed space Y is called nuclear if it can be represented as the sum of
Nuclear Spaces. Theory
309
one-dimensional operators Pn(x) = fn(x)en (en - an element,f - a continous linear functional) Co
T = f Pn n=1
such that 00
IITII^ = f IIPnII n=1 ^
The number I I TII is called the nuclear norm of the operator T.
This definition trivially implies that the sum of two nuclear operators is a nuclear operator and that a superposition of a nuclear operator with a continuous linear operator (also a superposition of a continuous linear operator with a nuclear operator) is a nuclear operator. PROPOSITION 7.2.2. If
lim n4dn(T) = 0, n-+ ao
then the operator T is nuclear. The proof is based on the following LEMMA 7.2.3. Let Y be an n-dimensional subspace of a Banach space X. Then there is a linear projection P, with the norm not greater than n, of the whole space X onto Y.
Proof. Basing ourselves on Lemma 7.1.6 and 7.1.7, we can find in Y such elements e1, ..., en andfunctionals f1, ..., fn that IIej1I = IIfiII = 1 (i = 1, 2, ..., n) and (7.1.5) holds. By the Hahn-Banach theorem we can extend each functional fj to a functional Fj of norm one defined on the whole space X. Let n
P (x) _ I Fj(x) e j .
(7.2.2)
i=1
The operator P is a continuous linear projection of X onto Y. Moreover, n
IIPII <
=1
IIFj!I IIej!I = n.
O
Chapter 7
310
Proof of Proposition 7.2.2. Let Ln be an n-dimensional subspace such that
Ln+2dn(T)BY) T(B)x and let Pn be a projection, with the norm not greater than n, onto L. The existence of such a projection follows from Lemma 7.2.5. Let x e Bx. Then T(x) = y+z, where y e Ln and IIzil 2dn(T). Thus IIPn(T(x))-T(x)II = IIPn(z)-zIl < (IIPnHI+1)IIzll 2 (n+ 1) dn(T) .
(7.2.3)
Now let
T(x) _ (T(x)-P1(T(x)))+(P1(T(x))-P2(T(x)))+ ...
(7.2.4)
On the other hand, dim (PnT-Pn+1T) (X) < 2n+2.
Therefore, by Lemma 7.1.6 and 7.1.7 and by the Hahn-Banach theo-
rem, there is a system of one dimensional operators Ki, . . . , Ken+1, IIK;11 = 1, i = 1, 2,..., 2n+1, such that the operator 2n+1
P'=Kj j=1
is a projection of the space X onto the space (Pn T-Pn+1T) (X). By (7.2.4) co 2n+1
T= Y
(PnT-Pn+1T)KJ".
n=Lj=1
The operators (Pn T-P.+, T) Kj are one-dimensional and, moreover, by (7.2.5) oo 2n+1
ao
II(PnT-Pn+1T)KjIJ
n=1
(2n+2) (11P. T- T11+ IIP.+, T- T11) n=1 oo
f (2n+2) (4n A- 3)dn(T) < +00. n=1
(2n+2)II(PnT-Pn+1T)II
Nuclear Spaces. Theory
311
PROPOSITION 7.2.4. Let T be a nuclear operator mapping a Hilbert space H into itself. Then
lim ndn(T) = 0.
(7.2.5)
n-- oo
be eigenvalues of Proof. The operator T is of course compact. Let the selfadjoint operator T*T, and let {en}, IIen11 = 1 be the eigenvectors corresponding to {An}. Let us choose en in such a way that they are orthogonal even if they correspond to the same eigenvalue. The operator T can
be written in the form oo
Tx = f 2n(x,en)fn n=1
where ),nfn = Ten and 11f,11 = 1. Let us order all A. in a non-increasing sequence
A,>A2-... Then dn(T) = An+1.
On the other hand, the operator T is nuclear, and thus, by definition, it can be written in the form 00
Go
T(x) = f (x, xn)yn,
where f IIxnII IIYn11 < +oo
n=1
n=1
Hence 00
00
n = (Ten,.fn)
(en, xj) (Yj,fn)
(en, xj)Yj, fn) _ j=1
j=1
Thus 00
00
00
00
: dn(T) = I A. = I I (en, xj) (yj,.fn) n=0
n=1j=1
n=1 .0
00
00
C j=1 LJ In=1 U I (en, xj)12)1/2 n=1
I (Yj,f)I2)1I2
= I IlxjII IIYj11 < +0o. j=1
and, since {dn(T)} is a non-increasing sequence, (7.2.5) holds.
Chapter 7
312
PROPOSITION 7.2.5. Let X and Y be two Banach spaces. Any nuclear operator T mapping X into Y can be factorized as follows :
X\\
T Y
//T
S2
S1
H where H is the the space 12.
Proof. Let us write Tin the form 00
2nfn(x) en,
(Tx) n=1
00
where II fnll = IIenII = 1, 2 < 0 and ,' A. < -f oo Let us put n=1
S,(X)
fn(X)lJ ,
x e X,
S1(X) E 12,
W
l
S2({ht}) _
n hnen,
{ht} E 12,
S2({hI}) E Y.
n=1
It is easy to verify that S2 S1 = T. PROPOSITION 7.2.6. Suppose we are given four Banach spaces Xl, X2, X3i X4
and let T{ be nuclear operators mapping Xi into X;+1(i = 1, 2, 3). Then lim ndn(T3 T2 Tl) = 0.
(7.2.6)
Proof. Let us factorize T1 and T2 as follows : T3
T2
T1
---k X2
-X2------
Xl
-X4 S2
S1
H1
-'H2 To
where H1=H2=12, The operator To is a nuclear operator mapping H1 into H2 as a super-
Nuclear Spaces. Theory
313
position of the nuclear operator T2 with continuous operators. Hence, by Proposition 7.2.4, 1im
ndn(To) = 0.
n-o0
On the other hand, T3 T2 T1 = S2T0 S1. Therefore 0 ( ndn(T3T2T1) < nIIS2II dn(To) IIS1II--HO.
This implies (7.2.6).
THEOREM 7.2.7 (Dynin and Mityagin, 1960). A Bo-space X is nuclear if and only if there is an increasing sequence of pseudonorms {IIxlIi} determining a topology such that the canonical mappings Ti are nuclear operators. Proof. Sufficiency. Suppose that a sequence of pseudonorms {IIxIIi} satisfies the properties described above. Let IIxIIi = IIxIIs, and let T; denote the canonical mappings with respect to the pseudonorms IIxIji. Then, by Proposition 7.2.6,
limndn(Ti)=0,
i=1,2,...
Thus, by Proposition 7.2.1, X is a nuclear space. Necessity. Suppose that X is a nuclear space. Then, by Proposition 7.2.1, there is an increasing sequence of pseudonorms {IIxIIi} such that the canonical embeddings satisfy (7.2.1). Let us put IIxIIi = IIxIi4, Then, by Proposition 7.2.2, the canonical mappings T; with respect to the pseudonorms jxlli are nuclear. PROPOSITION 7.2.8 (Mityagin, 1961). Let X be a Bo nuclear space. Then there is a sequence of Hilbertian pseudonorms IIxIIi = j/(x xx)j determining a topology equivalent to the original one.
Proof. Theorem 7.2.7 implies that there is a sequence of pseudonorms {IIxIIi} such that the canonical mappings Ti of Xi+1 into Xi are nuclear. This means that T{ can be written in the form Co
Ti(x) = I 'li ys n(x)yi,n n=1
where yi, n E Xi, yti n E Xi+1, I Iyi, nII = I Iy;,nll = 1, Ail > 0 and the series
Chapter 7
314 W
P, is convergent. Let us introduce now inner products in X by the n=1
following formula OD
(x, y)i =
?Y n(x)Y n(Y) n=1
On the one hand, we have Go
(x, x)i < f
IY*n(x)12 < A, IIxII +1,
n=0
CD
where A
On the other hand, n=1 Co
IIx4I1= 1 yz n(x)Yj,n <.E A; IY*n(x)I n=1
n=1 [co
CO
I Y%(x)1a,1/2
n=1
n=1
= A{(x, x)c,
Then Ai 11211xlli < (x,
X)1 12 %
<
<1,IIxII:+1
Hence the Hilbertian pseudonorms (x, x)112 yield a topology equivalent
to the original one. As an obvious consequence of Proposition 7.2.8 and Proposition 6.6.7, we obtain PROPOSITION 7.2.9 (Mityagin, 1961). Let X be a nuclear Bo-space and let Y
be a subspace of the space X. Then
6(Y)C 6(X).
7.3. UNCONDITIONAL AND ABSOLUTE CONVERGENCE
In Section 3.6 we have considered unconditional convergence in F-spaces. Now we shall consider other kinds of convergence of series in F-spaces.
Nuclear Spaces. Theory
315
xn of elements of an F-space X is absolutely
We shall say that a series n=1
convergent if for each continuous quasinorm [x] the series00I [xn] is conn=1
vergent. The classical Riemann theorem shows that if X is a finite-dimensional OD
space, then a series ' xn is unconditionally convergent if and only if it n=1
is absolutely convergent. Let {Um} be a basis of balanced neighbourhood of zero and let
[x]m = inf{t > 0: t e Um} be the quasinorm with respect to Um. It is easy to verify that a series W
00
[xn]m are con-
xn absolutely convergent if and only if the series n=1
n=1
vergent for m = 1, 2, ... If X is a locally bounded space with a p-homogeneous norm IIxHI, then
a series j' xn is absolutely convergent if and only if the series 7 [xn]lIp n=1
n=1
is convergent. PROPOSITION 7.3.1. An F-space X is locally convex if and only if each absolutely convergent series is unconditionally convergent.
Proof. Necessity. Let X be a locally convex space and let {U,n} be a basis of balanced convex neighbourhoods of zero. Let us denote by IIXIIr the
xn be an absolutely convergent
pseudonorm generated by Um. Let n=1 00
series in X. Then the series 2; IIxnIIm are convergent for m = 1, 2, ... Let n=1
{en} be a sequence of numbers equal either to 1 or to -1. Then 00
0<1 f Enxn M < n =k
11x-11M-->0
n=1
for k tending to infinity and for m = 1, 2....
Chapter 7
316
Therefore, by definition, the series 2' xn is unconditionally convergent. n=1
Sufficiency. Let X be a non-locally convex E-space. Let { Um} be a basis
of balanced neighbourhood of zero such that U,,,,+, C z Um. Since the space X is not locally convex, there is a neighbourhood of zero V such that cony U. V (m = 1, 2, ...). This means that there are elements xm,1, , xm, n,, of Um and non-negative reals am,1, ... , am, n,, such that nm
am,t = 1
(7.3.1)
%=1
and nm
I am,ixm,i 0 V.
(7.3.2)
i=1 Go
Let us order the elements am, i xm, i in the sequence {yn}. The series
yn n=1
is absolutely convergent. Indeed, let us denote by [x]k the quasinorm with respect to the set Uk. Then for j, j' > k we have
''1 nm
[am,ixm,i]k = m=j i=1
[am,ixm,ijk
m=j %=1
< Ym=sup [x]k < j zEU,,,
2m-k
m=j
1
2j-k-1
On the other hand, formula (7.3.2) implies that the series ' yn is not n=1 unconditionally convergent. If a space X is infinite-dimensional, then unconditional convergence does not imply absolute convergence. Dvoretzky and Rogers (1958) have
shown that in each infinite-dimensional Banach space there is an unconditionally convergent series which is not absolutely convergent. This
theorem has been extended to locally bounded spaces by Dvoretzky (1963).
In general, the problem when unconditional convergence implies absolute convergence is open. For locally convex spaces such characterization is due to Grothendieck.
Nuclear Spaces. Theory
317
THEOREM 7.3.2 (Grothendieck, 1951, 1954, 1955). Let X be a Bo space. The space X is nuclear if and only if each unconditionally covergent series in X is absolutely convergent.
The proof of this theorem, the main theorem of the present section is based on several notions, lemmas and propositions. We say that a linear continuous operator T mapping a Banach space X into a Banach space Y is absolutely summing if there is a positive constant C such that, for arbitrary x1, ..., Xn E X, IIT(x#1 < C
(7.3.3)
x{
%=1
i=1
PROPOSITION 7.3.3. An operator T satisfies (7.3.3) if and only if n
`n
Eixi II T(xi)ll < C sup ei=f1 i=1 i=1
(7.3.4)
I.
Proof Necessity. Let e{ = 1. Then n
Y Xi i=1
i=1
Thus (7.3.4) implies (7.3.3). Sufficiency. Let x1, ..., xn be arbitrary elements of X. Let s,"., ..., E.' be
arbitrary numbers equal to + 1 or -1. Then putting yi = e°Xi, i = 1, .. . ..., n, and applying (7.3.3) to yi, we obtain n
i=1
n
n
IjT(xi)II =
i=1
IIT(.v )Ij < C
n
f E{xill < Csup i=1
et =±1 i=1
Eixi
.
CI
PROPOSITION 7.3.4. A linear operator T satisfies (7.3.4) if and only if n
n
.Y IIT(xi)II < C sup i=1
I If
If(xi)I i=1
(7.3.5)
Chapter 7
318 00
Proof. Sufficiency. Let y = E 8°x{ be such an element that j=1
n
1E{x{ 11 . IIYII = SU 'Sup ,=p1 ii=1
Let f' be a functional of norm one such thatf'(y) = IIYII Then n
sup
n
= IIYII =f'(Y) _
e{ x{
ei=±1 i=1
f'(E°xi) {=1
n
If'(x{)I < sup i=1
If(xi)I
IIfIX'1 i-1
Therefore (7.3.4) implies (7.3.5). Necessity. Let e° = signf(xi) for a functional f e X* of norm one. Then n
n
f(x) =
n
f(E°xi) =f(E e°xi) i=1
i=1
i=1
n
n
E°xill {=1
Y E{x{ < sup ei=f1 i=1
Therefore, (7.3.5) implies (7.3.4).
Formulae (7.3.3)-(7.3.5) give us three equivalent definitions of absolutely summing operators. The infimum of those T which satisfy (7.3.3) will be denoted by a(T). Let T be an absolutely summing operator belonging to B(X-Y). Let A e B(Y-- Z) (or A e B(Z->X)). Then the operator AT (resp. TA) is absolutely summing. PROPOSITION 7.3.5. A linear operator T mapping a Banach space X into a Banach space Y is absolutely summing if and only if it maps unconditionally convergent series into absolutely convergent series.
Proof. Let I00x be an unconditionally convergent series. Then n=1
k'
lim sup k.k'.-
e,
±1
enxn = 0.
(7.3.6)
Nuclear Spaces. Theory
319
Let T be an absolutely summing operator. Then by (7.3.4) and (7.3.6) k'
lim f IIT(xn)II = 0,
k,k'- oo ,=k
and the series Y T(xn) is absolutely convergent. n=1
On the other hand, if we suppose that an operator T e B(X--Y) is not absolutely summing, then, by definition, for any k there are elements {xk....... xk, nk} of X such that nk
sup
ei=f1 {=1
Etxk,i
(7.3.7)
and nk
(7.3.8)
II T(xk,{)II > 1. {=1
Let us order all xk,i into a sequence {yn}. Formula (7.3.7) implies that the Co
series
7 yn is unconditionally convergent. Formula (7.3.8) implies that
n-
the series Ico T(yn) is not absolutely convergent. n=1
PROPOSITION 7.3.6. Each nuclear operator is absolutely summing.
Proof. Let T e B(X- .Y) be a nuclear operator. This means that the operator T can- be written in the form
T(x) _
ingn(x)Ym n=1 m
where An > 0, C = 2' A. < +oo, gn e X*' Y. e n=1
(n = 1, 2,
Y, IIgnjj = IIYnI I = 1
.).
Let x1, ..., xN be arbitrary elements of X. Let fs, i = 1, ..., N be a continuous linear functional of norm one defined on Y such that f (T(x{))
Chapter 7
320
_ JIT(xi)Il. Then N
N
N
fi(T(Xi)) _ I fi(
JIT(xi)Il _ i=1
t=1
i=1
n=1
2ngn(xi)yn)
i=1
N
co
<
I N
N
A-1 I gn(xt)I If (Y.)I < C sup
gCX* i=1
i=1
Ig(xi)I
Hence, by (7.3.5), the operator Tis absolutely summing.
We say a continuous linear operator T mapping a Hilbert space H1 into a Hilbert space H2 is a Hilbert-Schmidt operator if, for any orthonormal sequence {en} in the space H1, CO
I I T(en)I I2 < + 00 . n=1
This definition is clearly equivalent to the following one. An operator T e B(H3--.H2) is called a Hilbert-Schmidt operator if, for an arbitrary orthonormal sequence {en} in H1 and an arbitrary orthonormal sequence {fn} in H2, 00
I(T(et),fi)I2 < +oc 7.i=1
This implies that an operator conjugate to a Hilbert-Schimdt operator is also a Hilbert-Schmidt operator. PROPOSITION 7.3.7. If an operator T e B(H1-*H2), where H1 and H2 are Hilbert spaces, is absolutely summing, then it is a Hilbert-Schmidt operator. Proof. Let {ei} be an arbitrary orthonormal set in H1 and let {ai} be an arbitrary sequence belonging to 12. Let xt = atet. Then, by (7.3.5), n
(1auIITfrh1
)C 2
n
ti=1
o0
t=1 Go
Thus, by the arbitrariness of n, we find that the series 7 aiJIT(ei)Il is coni=1
Nuclear Spaces. Theory
321 cc
vergent. Since this holds for all sequences {an} e 12, the series I II T(ei)II2 is i=1
convergent. This means that T is a Hilbert-Schmidt operator. PROPOSITION 7.3.8. The superposition of two Hilbert-Schmidt operators is a nuclear operator.
Proof. Let H1, H2, H3 be Hilbert spaces. Let T e B(H1->H2), and let Se B(H2-*H3) be Hilbert-Schmidt operators. Let {en} be an arbitrary orthonormal set in H2. Then co
co
ST(x) = f (T(x), en)S(en) = L, (x, T*(en))S(en), n=1
n=1
where T* a B(H2-->H1) denotes the operator conjugate to the operator T. The operator T* is also a Hilbert-Schmidt operator. Thus 00
IIT*(en)II IS(en)II n=1
w
(n=1IIT
*(en)1/2
IS(en)I2,1/2
<
oo.
,a=1
Hence ST is a nuclear operator. THEOREM 7.3.9 (Pietsch, 1963). Let T e B(X-*Y) be an absolutely summing oprator. -Then there is a probability measure (i.e. a regular positive Borel measure with total mass 1) u on the unit ball S* of the conjugate space X* such that
IIT(x)II < a(T) f Ix*(x)I dp(x*). s*
Proof (Lindenstrauss and Pelczyriski, 1968). Let n
W = {g e C(S*):g = a(T)
n
I.fx,(x*)I with
IIT(x{)II = 11,
where fx(x*) = x*(x) for x* e S* and x e X. We shall show that the set W is convex. Let n
m
g1 = a(T) I, I fx,,a(x*)I ,
g2 = a(T) I I fxa.,(x*)I ,
i=1
i-1
Chapter 7
322
where n
m
IIT(xi,l)II =
IIT(xi,2)II = 1.
(7.3.9)
Let
a+b = 1.
a, b >_- 0,
(7.3.10)
Let
for j = 1, 2, ..., n, for j = n+1, ..., n -t-m.
laxi,i YJ =
bxi_n,2
Then, by (7.3.9) and (7.3.10) n+m
m
n
IIT(YJ)II = a
IIT(xi,2)II = 1.'
IIT(xi,l)Il+b
j=1
i=1
i=1
Moreover, n+m
g(t) = a(T) Y Ify,(x*)I J=1 n
m
= a(T)
Ifa,,,(x*)I+ i=1
I fbx,,,(x*)I i=1
n
= a(T) [a
m
Z I fx,..(x*)I +b
i=1
I ff,,.(x*)I ] = agi+bgi. i=1
Thus the set W is convex. The definition of a(T) implies that if
IIT(xi)II = 1, then
i
n
sup x*ES* ti-1
Ix*(xi)I = sup
x*eS* i=1
I fx,(x*)I >' 1
(see Proposition 7.3.4). Therefore, the set W is disjoint from the set
N = {fe C(S*): f(x*) < 11. The set N is open and convex. Therefore, there is a continuous linear functional F defined on the space C(S*) such that
F(f) > 1
forfcW
(7.3.11)
Nuclear Spaces. Theory
323
and
F(f) < 1
for f e N.
(7.3.12)
The general form of continuous linear functionals on the space of continuous functions implies that there is a regular Borel measure po defined on S* with its weak-*-topology such that
F(f) = f .f(x*)d uo(x*). S*
Since the set N contains the cone of negative functions in C(S*), by (7.3.12) the measure po is positive. Thus it is of the form po = ap, where p is a probability measure and a = IIFUI. The set N contains the unit ball in C(S*), hence, by (7.3.12), a = IIFII < 1.
Let x E X and T(x) :y 0. Then g = a(T)
1
IIT(x)I
I fx(x*)I e W. There-
fore, by (7.3.11)
f gdp > f gdpo > 1. S*
S*
Thus
IIT(x)II < a(T) f Ifx(x*)dp(x*) = a(T) f Ix*(x)I dp(x*) S*
S*
and this completes the proof. THEOREM 7.3.10 (Pietsch, 1963). Let T be an absolutely summing operator mapping a Banach space X into a Banach space Y. Then the operator T can be factorized as follows
X
T
Y
i
C(M)->H 1
I
where H is a Hilbert space, M is the unit ball S* in the conjugate space X* with its weak-*-topology, and i is the natural embedding of X into C(M).
Proof. Let p be a probability measure defined in Theorem 7.3.9 on the
Chapter 7
324
set M. Let L1(p) denote the completion of C(M) with respect to the norm IIxII = f I x(t)I dp, and let L2(p) denote the completion of C(M) with rem
spect to the norm
IIxII = [f Ix(t2)Id z]"2
if
Let C(M)_%L2(4u)->L'(p)
be natural injections and let Z be the closure of ja i(X) in the space L'(p). The theorem follows from the diagram
T
X
-* y
Z C L'(p).
i
C(M)
H = L2(p) a
Theorem 7.3.9 implies that the operator y is continuous. THEOREM 7.3.11 (Pietsch, 1963). A composition of five absolutely summing
operators is a nuclear operator. Proof. Let us consider the diagram Ti
Xl
'X3
\ /
.\ \'//
i
T3
T2 -->X2
Hl
T4 -->X4
--X6 \
- H3
-- H2 a
T5
_XB
/ J
rg
The existence of such factorization follows from Theorem 7.3.10. The
operators a, j9 are absolutely summing as compositions of absolutely summing operators with continuous operators. Therefore, by Proposition 7.3.8, the operator #a is nuclear. Thus the operator TS T4 T3 T2 Tl = jflai is nuclear.
Nuclear Spaces. Theory
325
Proof of Theorem 7.3.2. Sufficiency. Let X be a nuclear B, -space and let the topology in X be given by an increasing sequence of homogeneous pseudonorms {IIxIIr} such that the canonical mappings T{ from X,+1 into X{ are nuclear. By Proposition 7.3.6 the operators Tj are absolutely summing.
xn be an unconditionally convergent series in X. This means
Let n=1
that oD
lim sup
e,=±1 n=k
snxn
r
= 0,
i = 1, 2, ...
Since the canonical mappings Ti are absolutely summing, this implies that 00
the series S I Ixnl a- are convergent for i = 2, 3,... n=1
Necessity. Let X be a Bo space and let {IIxIr} be an increasing sequence of pseudonorms determining the topology. Theorem 7.3.11 implies that it is sufficient to show that for any pseudonorm IIxIIr there is a pseudonorm IIxIII such that the canonical embedding Xj into Xr is an absolutely summing operator. Suppose that the above does not hold. This means that there is a pseudonorm IIxIIr0 such that the canonical embedding Xr into X j, is not ab-
solutely summing for any i > i,. Then, by definition, there are elements xr, 1, ... Xi, n, such that ns
(7.3.13)
L, IIxr,lllro = 1 j =1
and
sup 11f sjx,,;
<1
(7.3.14)
2t e,=f1 =1 { Let us order xi,n in a sequence. Formula (7.3.14) implies that this series is unconditionally convergent. On the other hand, by (7.3.13) it is not absolutely convergent. El
COROLLARY 7.3.12 (Dvoretzky and Rogers, 1950). In each infinite dimensional Banach space there is an unconditionally convergent series which is not absolutely convergent.
Chapter 7
326
Proof. If each unconditionally convergent series were absolutely convergent, then the space would be nuclear, and thus a Montel space. Since each Banach space is locally bounded, it would be locally compact. Thus it would be finite dimensional. Rosenberger (1972, 1973) gave the following extension of the notion of
nuclear spaces. Let 0 be a class of continuous, subadditive, strictly increasing functions defined on [0,+oo], such that all functions (P e 0 vanish at 0. An F*-space X is called 1-nuclear if, for any balanced neighbourhood of zero U, there is a balanced neighbourhood of zero V such that 00
I cp(bn(V, U)) < +oo
for all ry e 0.
n=0
When 0 contains only one function rp, we shall mark 0-nuclear spaces as (p-nuclear spaces. By Theorem 7.3.2 if X is locally convex, then it is nuclear if and only if it is t-nuclear. Moscatelli (1978) gave the conditions for (P ensuring the
existence of a universal p-nuclear space for locally convex gyp-nuclear spaces.
7.4. BASES IN NUCLEAR SPACES
Let X be a locally convex space, i.e. a space in which the topology could be given by a sequence of homogeneous pseudonorms. Let {en} be a basis in X. We say that a homogeneous continuous pseudonorm jlxii is admissible if r
r+8
If tnen
for r, s > 0.
(7.4.1)
n=1
Theorem 3.2.14 implies that in each locally convex space there is a sequence of homogeneous admissible pseudonorms determining a topology equivalent to the original one. PROPOSITION 7.4.1. Let X be a locally convex space with a basis {en}. If for each sequence of homogeneous pseudonorms {Ilxilm} determining a topology
Nuclear Spaces. Theory
327
equivalent to the original one for every i there is a j such that Ci,1 =
' IIenIIs < +00,
(7.4.2)
n=1 IIenII1
0
where we assume 0 = 0, then the space X is nuclear. Let {IIxlIj} be an increasing sequence of homogeneous pseudonorm :. r n;;i:'ng the topology. Without loss of generality we can assume that p.': udonorm IIxII{ are admissible. Let us take an arbitrary i. Then, by the hypothesis, there is an index j such that (7.4.2) holds. Let us denote by {f,,} the sequence of basis functionals, Let IIxII1 < 1. Since the pseudonorm IHxIl1 is admissible n
I1fn(x)enII1
i=1
n-1
,Zf(x)eilll <2IIxJI1 < 2. (7.4.3) i=1
The canonical embedding T1, j of X1 into XX is nuclear, provided (7.4.2)
holds. Indeed. Co
P,(x),
Ti, {(x)
where PP(x) = fn(x) en
n=1
are one-dimensional operators. Moreover, by (7.4.3) IIPnII = sup IIPi(x)IIi = sup Ilfn(x)enllg Ilxllj<-l
= sup Ilfn(x)enlll I lxl b-
I1xib<1
IIenIIt I IenI I1
<2
IIenIIi I lent l5
Hence, by (7.4.2) 00
2, lip-11 < +oo.
(7.4.4)
n=1
Therefore, X is a locally convex nuclear space. PROPOSITION 7.4.2. Let X be a locally convex space with a basis {en}. If there is in X an increasing sequence {IIxIIs} of pseudonorms determining a topology such that (7.4.2) holds, then each increasing sequence of pseudonorms {Ilxllz} determining the topology has the same property.
Chapter 7
328
Proof. Let us take an arbitrary is Since the pseudonorms {Ilxlli} yield the topology, there are an index ii and a constant C such that Ilxlli < Cllxlli, Let ji be such an index Ci,,,, < -boo. The pseudonorms 11x11,, yield the
topology, therefore, there are an index j and a constant K such that Ilxll,, < Kllxll'. Thus Ilenlli
< KC
I lenH
Ilenllf.
Ilenll; Therefore, co
V Ilenlli < KCCil,1.. n=1
Ilenll
We shall now show a theorem converse in a certain sense to Proposition 7.4.1. THEOREM 7.4.3 (Dynin and Mityagin, 1960; Mityagin, 1961). Let X be a locally convex space with a basis {en}. Let the topology in X be given by an increasing sequence of admissible pseudonorms {llxlls}. If the space X is nuclear, then for every i there is a j such that (7.4.2) holds. Proof. Since the space X is nuclear, then for each i there is an index j such
that limn4b,,(B,, Bi) = 0,
(7.4.5)
where
B, = {x e X: IIxIIr < 1}. Let z'
II
en a ll
for
Ilenlli
0.
Since (7.4.5), lim Ilz7 Ili = 0. Let us reorder the positive integers in a se-
quence {kn} in such a way that Ilekmlli
Ilek.,ll;
>
Ilekm+illi Ilekm+lllf
(m = 1, 2, ...).
(7.4.6)
Nuclear Spaces. Theory
329
Let
n-1 let
an
I1en11t ,
bn
(we admit 1/0 = oo), QnzllenF?
and let A
= {x E X: I ft(x)I < at(i = 1, 2, ...)}, I ft(x)I < b{(i = 1, 2, ...)},
B = {x e X:
where {fn} are the basis functionals with respect to the basis {en}. If x e X and IIxik{ < 2, then jIfn(x)enjJ{ < 1 (i = 1, 2,...) because the pseudonorm IjxI!i is admissible. Therefore, 2 Bi C A. On the other hand, B C B5, because for x e B M
Go
00
j xII9 < f II.fn(x)enJI9 = I Ifn(x)I JIenjI1 < I b,I je, jj n=1
n=1
n=1
= 1. n=1 Therefore,
6n(Bj,(2)&) > 6.(B, A).
(7.4.7)
By (7.4.6)
sn-1(B A) = bk' akn
=
1 (-on-'Q'
Ilek 1k IIek.11j
Thus (7.4.5) imply limns IIekAII{ = 0
n-ao
Ileknl!J
Therefore, (7.4.2) holds.
0
We say that a basis {en} of an F-space X is unconditional if the series of
expansions with respect to this basis are unconditionally convergent. A basis {en} of an F-space X is called absolute if the series of expansions
Chapter 7
330
are absolutely convergent. Proposition 7.3.1 implies that in locally con-
vex spaces each absolute basis is also unconditional. Theorem 7.4.3 implies
THEOREM 7.4.4 (Dynin and Mityagin, 1960). If X is a nuclear B0-space, then each basis {en} in X is absolute. Proof. Let {IIxIIs} be an increasing sequence of admissible pseudonorms determining the topology. By Theorem 7.4.3, for each i there is a j such that (7.4.2) holds. Therefore W
00
Ilfn(x)enJI n=1
Ifn(x) I IIenII1 n=1
11e.11{
IIenII1
w
sup jIfn(x)enII1
IIenII{
n=1
< 2IIxIIjCC,t
CJ
11e.11?
COROLLARY 7.4.5 (Dynin and Mityagin, 1960). In a nuclear Bo-space X all bases are unconditional.
It is not known what situation there is in non-locally convex spaces. There is an interesting question : does Theorem 7.4.4 and Corollary 7.4.5 characterize nuclear B,-spaces ? Pelczyliski and Singer (1964) have proved that in each Banach space with a basis there is a non-unconditional basis. Wojtynski (1969) has proved that if, in a Bo space X, the topology is given by a sequence of Hilbertian pseudonorms and each basis in X is unconditional, then the space X is nuclear. In the same paper it is shown that if in a Bo-space X all bases are absolute, then the space X is nuclear. Let X be a nuclear Bo-space with a basis {en}. The basis functionals corresponding to {en} are denoted by {fn}. Let {IIxJIm} be a sequence of pseudonorms determining the topology in X. Let us assign to each element x e X a sequence {fn(x)}. Since {en} is a basis, we have, by Theorem 2.6.1. lim I.fn(x) I IIenJ Im = 0,
m = 1, 2, ... ,
Nuclear Spaces. Theory
331
i.e. {fn(x)} e M(am,n), where am,n = IIenIlm (compare Example 1.3.9). Moreover, the operator T(x) = {fn(x)} is a continuous operator mapping X into M(am,n) (see Theorem 2.6.1). Now we shall show that T(X) = M(am, n). To do this it is sufficient to prove that if {tn} e M(am, n) then 0
the series f tnen is convergent in X. The space X is nuclear, therefore, n=1
for each index m there is an index r such that a
W
=
Cm,r =
Ilenlim
G n-1
< b oo
IIenIIr
(see Theorem 7.4.3). Hence Iltnenllm = n=1
Iltnll IIenIIr n=1
II enII m
IIenIIr
< +00,
where, according to Example 1.3.9 Il{tn}IIr = Supltnlar,n = supltnl Ilenllr. n
Therefore, the series
n
tnen is absolutely convergent, and thus conver-
gent. Therefore, T(X) = M(am,n). Since both spaces are complete, by the Banach theorem (Theorem 2.3.2) the operator T-1 is continuous. Hence the following proposition holds : PROPOSITION 7.4.6 (Rolewicz, 1959c). Let X be a nuclear Bo space with a basis {en} and the topology determined by a sequence of pseudonorms {I IxI Im}. Let am, n = Il enll m The set of expansions with respect to the basis
{en} constitutes the space M(am,n) The corresponyence between x e X and {fn(x)} a M(am,,,) is a homeomorphism.
In many applications n and m are vectors consisting of non-negative integers (see Example 1.3.9). There are nuclear Bo-spaces without bases. The first such an example
was constructed by Mityagin and Zobin (1974) (see also Djakov and Mityagin, 1976; Bessaga, 1976). Even more, Dubinsky (Dubinsky, 1979,
Chapter 7
332
1981 ; see also Vogt, 1982) showed that there are Bo-spaces without the bounded approximation property'.
Proposition 7.4.6 play an important role in the theory of nuclear spaces, because the spaces M(am,n) have a very simple structure. Sometimes, however, it is more convenient to consider in M(am, n) other families of pesudonorms determining the topology. PROPOSITION 7.4.7. Let [am, n] be a matrix of non-negative reals. Let am,n n
(7.4.8)
am+1 n
Let LP(am, .n) (1 < p < + oo) (see Example 1.3.9) be the space of all such sequences {tn} that (7.4.9)
tnam,nlp)1/n < -boo
II{tn}Ilm,p = (.Y I n
with the topology determined by the pseudonorms IIxlIm Then the space LP(am,n) is identical with the space M(am,n). Proof. Since ll {tn}II'+n,p
(7.4.10)
II{tn}IIm,
where II {tn}II m = sup I to l am, n, we have LP(am,n) C M(am, n)
On the other hand, H{tn}Il m,p
_ (2
/p I
I tnam+l,nl p(
tnam,aV')1
am,n
)PY/P
am+1,n
where by (7.4.8) Cm,p
atn,n )PY/P (am+1,n n
<
+ CO.
As an application of Proposition 7.4.7, we obtain 1 We say that an F-space X has the bounded approximation property, if there is a uniformly continuous sequence of continuous finite dimensional operators T. mapping X into itself such that lim T. x = x for all x E X. n
Nuclear Spaces. Theory
333
PROPOSITION 7.4.8. Each continuous linear functional defined on a space M(am, n), where a.,. satisfies (7.4.8), is of the form
f(x) _ Z fnxn, n
where {xn} a M(am, n) and { fn} is such a sequence of scalars that
sup Ifnl n
am,n
+.
(7.4.11)
for a certain m. Proof. By Proposition 7.4.7 the space M(am,n) is identical with the space L1(am, n). The general form of continuous linear functionals on B0-spaces implies that the functional f is continuous with respect to a certain norm
-Ixllm,l. Hence the general form of continuous linear functionals in 11 implies (7.4.11).
We say that a basic sequence (see Section 2.6) of elements {x.} of a nuclear Bo-space is represented by a matrix [am,n] (we shall denote it briefly by {Xn} [am,n]) if there is a sequence of pseudonorms {I1x44m} determining the topology in X such that IIxnI1m = a.,..
7.5. SPACES WITH REGULAR BASES
Let X be a Bo-space with a basis {en}. A basis {en} is called regular if there is a sequence of pseudonorms {11 11m} determining the topology such that
the sequence
1jenil m 11en11m+i
I
is non-increasing for all m.
Let {an} be a non-decreasing sequence of positive numbers. The standard bases in the Kothe power spaces M(an) and M(an1Im) are regular. Let X be an F-space. Two bases {en} and {fn} in Xare called semi-equiv-
alent if there is sequence of scalars {rn}, rn > 0 such that the bases {en} and {rn fn} are equivalent.
THEOREM 7.5.1 (Djakov, 1975; Kondakov, 1974). Let X be a nuclear Bo-space with a regular basis {en}. Then each regular basis {fn} in X is semi-equivalent to {en}.
Chapter 7
334
Proof. By the definition of regular bases there are sequences of pseudonorms {II IIm, o} and {11 II;n, o} both determining a topology equivalent to
the original one and such that the sequences
IIenllm,o {II enIIm+1,o
land
IIfnll m,o , IIIIIAII m+1,0
}
are non-increasing. By Proposition 7.4.3 we may assume without loss of generality that 00
IIXIIm,o = f le' (x)l Ilenllm,o
(7.5.1)
n=1
and
IxII,,o =
(7.5.2)
if-'(X)I Ilfnlim,o, n=1
where {en} denote the basis functionals with respect to the basis {en} (resp. to the basis {fn}). Let Um, o = {x: I IxI Im, o < 1} and Vm, o = {x : I IxI Im,o < 1}. Clearly it is possible to find a subsequence {mr} and two sequence of positive scalars {ar} and {br} such that
a1Um,,o D biVm,,o) a2Um,,o D b2Vm,,o) ...
(7.5.3)
Let IxIIr = ar 1IIxjImr,O
and IxIIr = br 1IIxIIm,,o
Of course
Ur = {x: IIxIIr < 1} = arUm,,o and
Vr = {x: IIxIIr < 1} = Using the fact that the bases {en} and {fn} are regular, we can easily calculate the approximative diameters, bn-1(Ur, U8) =
IIenIIr
IIenIIr '
bn-1(Vr, V8) =
IIfniIr
(7.5.4)
IIfnII:
Take t < s ; then, by (7.5.3), Ut ) Vt > V8 ) U8}1 and an(U8+1, Ut) < bn(V8j Vt).
(7.5.5)
Nuclear Spaces. Theory
335
Thus, by (7.5.4), (7.5.6) Ilenll8+1
I Jnll8
Take t > s, then by (7.5.3) V8 ] U8+1) Ut) V, and an(Vt, VS) < Sn(Ut, Us+1).
7.5.7)
Thus, by (7.5.4), II.fnIls
IIenIls+1
JI.fnIJg
JlenIIt
(7.5.8)
Therefore, by (7.5.6) and (7.5.8), IIen t
_
IIenhI8+1
Ilfn11t C ins for all t, s, n = 1, 2, ... This implies that rn = sup Illnllt < -f-so.
(7 5 9)
(7.5.10)
Thus, by (7.5.9), IlenIIt < IIrnfnIIi < IIenIkt+i.
(7.5.11)
Therefore the bases {en} and {r,, fn}e ar equivalent. We say that two bases {en} and {fn} are quasi-equivalent (see Dragilev, 1960) if there is a permutation o of positive integers such that the bases {en} and f f.(,,)} are semi-equivalent. THEOREM 7.5.2 (Crone and Robinson, 1975). Let X be a nuclear Bo-space
with a regular basis {en}. Then each basis {fn} in X is quasi-equivalent to {en}.
The proof of the theorem is based on several notions and lemmas and propositions. LEMMA 7.5.3 (Kondakov, 1983), Let X = L'(am,n) Assume that IIXIIm < 2m
(7.5.12)
Chapter 7
336
Suppose that for an element f e X there is a functional f' c- X* such that
f'(f) = 1 and (7.5.13)
supllf'II,'n+2IlfIlm+1 = C < +00, m
where II IIm denotes the norm of the functionals induced by the pseudonorm II
IIm
Then there are 2 > 0 and an index i such that le{Ilm < 2Ilf llm+1 < Cile{IIm+2,
in = 1, 2, ...
(7.5.14)
where {en} denotes the standard basis in L1(am,n)
Proof To begin with, we shall show the first inequality. Let {en} denote the basis functionals corresponding to the basis {en}. Then co
00
Ien(f)ISUP
len(f)Illenllm
11
n=1 m=1
n=1
00
00 f lIen(f)enllm
IIfIIm
n=1
IIfIIm+1
m=1
W
Ilenllm
00
m=1 IIfIIm+1
_ <
1 2'n 1
__ 1
00
=f' (,Yen(f) en) < n=1
IIfIIm+1
00
(7.5.15)
le'(f)I If'(en) I . n=1
Comparing the series on the left and on the right, we find that there is an index i such that (7.5.16)
< If'(es)I llfllll+l
Putting A = f'(ei), we obtain the first part of the inequality. By (7.5.13) we have IIf'I1m+2lle{IIm+2IlfIlm+1 < C IIfIIm+1 < f'(e{)I and we obtain the second part of the inequality.
IIetllm+2
PROPOSITION 7.5.4 (Kondakov, 1983; cf. Dragilev, 1965). Let {f8} be a basis in a space L1(am, n). Then there are a sequence of constants {as}, as > 0, a sequence of indices {n8} and a subsequence {II Ilp} of the sequence
of standard norms such that Ilen.llp < asllfsllp+l
Ilen.IIP+2,
where {e} denotes the standard basis in L1(am, n).
(7.5.17)
Nuclear Spaces. Theory
337
Proof. Let
am,n = 22" sup ai,n 1
(7.5.18)
It is easy to verify that the space L1(an, n) is isomorphic to the space L}(am,n) and, for the space Ll(arnn), (7.5.12) holds. Let X = Ll(am,n) Let f f,} be a basis in X. Let { f 8} denote the basis functionals corresponding to the basis (f}. By (7.5.12) for each pseudonorm II Il,n and each a > 0, there is a pseudonorm II Ilm1 such that allxllm < Ilxll,ni. Thus for each index r there is an index m(r) such that IIfB (x)fsllr < Ilxllm(r)
(7.5.19)
IIfs Ilm(r)Ilfsllr < 1.
(7.5.20)
and
Then there is a sequence of pseudonorms determining a topology such that (7.5.13) holds for f = f8 and C = 1. PROPOSITION 7.5.5 (Dragilev, 1965). Let X be a nuclear Bo space with a regular basis {en}. Let {fn} be an arbitrary basis in X. Then {fn} can be reorderd in such a way that it becomes a regular basis.
Proof. By Proposition 7.4.7 and 7.4.8 the space X is isomorphic to the space L}(Ilenll m. Thus, by Proposition 7.5.4, for each s there are as and n8 such that (7.5.17) holds. Now we shall show that in the sequence {ns} each index can be repeated
only a finite number of times. Indeed, suppose that e,, = en,, = en,s = ... Then the space X0 spanned by { fs,, fs,, fs,, ... } is isomorphic to the
space 1. It leads to a contradiction, since X is nuclear and it does not contain any infinite-dimensional Banach space. Hence we can find a permutation o of positive integers such that the sequence n0(8). is non-decreasing. Now we shall introduce new pseudonorms in X W
Ilxllo,p = Yas lllf8 (x)Ilpllen.IIp 8=1
{II
Since the space X is nuclear, by (7.5.17) the sequence of pseudonorms 110,p} determines a topology equivalent to the original one.
Chapter 7
338
Observe that Ilfa(s)Ilo,m Ilfa(s)I lo,m+1
__
Ilen,(,)Ilo,m
(7.5.21)
I len,(.)I lo,m+1
Since u(s) is non-decreasing and the basis {en} is regular, by (7.5.21) the basis {fa(,)} is also regular. Proof of Theorem 7.5.2. Let X be a nuclear Bo-space with a regular basis {en}. Let {fn} be an arbitrary basis in X. By Proposition 7.5.5 there is a permutation u(n) such that the basis {fa(n)} is regular. Then, by Proposition 7.5.4 {fa(n)} is semi-equivalent to {en}, and this completes the proof. Theorem 7.5.2 was first proved by Dragilev (1960) for the space of ana-
lytic functions in the unit disc. He then extended it to nuclear spaces of the types d1 and d2 with regular bases (see Dragilev, 1965). Let d1, i = 1, 2 denote the class of spaces with regular bases belonging
todi,i=1,2. PROPOSITION 7.5.6 (Zahariuta, 1973). Let X, Y be nuclear Bo spaces such that X e d1 and Y c d2. Then all bases in the product X X Yare quasi-equivalent.
Proof. Let {ef }, {en } be regular bases in X and Y. Let e2._1 = (en,0), e2n = (0, ey). In this way we obtain a basis in X X Y. Let {f..} be another basis in X x Y. By Proposition 7.5.4 there are a sequence of indices {nk} and a sequence of positive numbers {ak} and a sequence of pseudonorms {II IIv} determining a topology equivalent to the original one such that IlentIIp < Ilakfkll,+1- IlenkIIP+2.
(7.5.22)
Let
X1 = : lin{fk: en. E X} and
Y1=lin{fk: en.eY}. Let {f1,3} be a subbasis of the basis {fk} consisting of those elements which belong to X1. Let {f2,8} be a subbasis of the basis {fk} consisting of
Nuclear Spaces. Theory
339
those elements which belong to Yl. By (7.5.17), in the same way as in the proof of Proposition 7.5.5, we infer that {f1,8} and {f2,s} are regular bases in X, and in Y1. Moreover, X, E d2, Y, c- d1. By Theorem 6.7.10, this implies that there is an integer s such that X1 is isomorphic to X(') and Y, is isomorphic to Y(-8). Since we can shift s elements of the basis form X
into Y if s > 0, and conversely from Y into X if s < 0, we can assume without loss of generality that s = 0. The bases {en}, {e'} are regular; by Theorem 7.5.2 there are permutations a(s), a'(s) and sequences of positive scalars {a8}, {a} such that fl, s = a8 ex, and f2, 8 = a eY (8), which trivially implies that the bases {en} and {fn} are quasi-equivalent.
Theorem 7.5.6 for X = M(expm(n1+...+nk)) and Y=M(exp-
1
M
(n1+...+nk) was proved by Dragilev (1970) and Zahariuta (1970). Zahariuta (1975) proved Theorem 7.5.2 for another important class of spaces. Let {a,,n}, {b1,n}, {a2,n}, {b2in} be four sequences of real numbers tending to infinity. We shall assume that there is a positive number c such that c-la{,n < ai,2n < ca{,n,
i = 1, 2,
c-lbt,n < bi,2n < cbi,n,
i = 1, 2.
Let n = (n1,n2). We shall consider two spaces, X =
and Y = M(azn, b2-1n'). Zahariuta (1975) showed that if the diametral dimensions of the spaces X and Yare equal, 6(X) = 6(Y), then the spaces X and Y are isomorphic and, what is more, that there is a permutation of indices o such that {fn} is equivalent to {e,()}, where {en} and {fn} are standard bases in X and respectively in Y. As a consequence of this fact, it is possible to obtain the following results, arrived at independently by Djakov (1974) and Zahariuta (1974).
Let X = M (exp (- m nl +mn2)) and Y = M (exp(- m nl"+mn2'.)) Then the spaces X and Y are isomorphic if and only if 1
1
1
1
pq -p
f
q
Chapter 7
340
Moreover, if X and Y are isomorphic, there is a permutation o of indices such that the bases {fn}, {e0( )} are equivalent, where {en} and {fn} are standard bases in X and Y respectively.
There are also other classes of spaces in which all absolute bases are quasi-similar. Let H be a Hilbert space and let A be a self-adjoint positively defined
operator acting in H. Let (x, x)a = (Aa(x), Aa(x)),
(-oo < a < oo)
Let Ha be the completion of the domain DA of the operator A with respect to the norm IIxila = (x,x)a. The system of spaces {Ha} (-co < a < -boo) is called a Hilbert scale (Krein, 1960; Mityagin, 1961). Let Hb = n Ha with the topology induced by the pseudonorms Ilxlia a
a < b. The space Hb is a Bo-space and it is called the centre of the Hilbert scale {Ha}. If b is finite, we say that the centre Hb is finite, if b = +oo, the centre Hb is called infinite.
If the operator A-' is a nuclear operator, then the space Hb is isomorphic to a Kothe power space, and by Theorem 7.5.2 all bases in Hb are quasi-equivalent. If the operator A-' is not nuclear, then by the previously mentioned result of Wojtytiski (1969) there are non-unconditional bases in Hb. Zahariuta (1968) showed that if A-' is compact, then all unconditional bases are quasi-equivalent. This result was generalized by Mityagin (1969) to positively defined operators A.
7.6. UNIVERSAL SPACE FOR NUCLEAR SPACES
In this section we shall prove that there is a nuclear Bo-space U universal for all nuclear Bo spaces with respect to linear dimension. Let X = L2(nm). Let o be the space of all sequences x = {xn}, xn E X, with the topology given by a sequence of pseudonorms IIXIIp,m = IIXPIIm,
where {IIxIIm} is a sequence of pseudonorms defining the topology in X.
Nuclear Spaces. Theory
341
THEOREM 7.6.1 (T. Komura and Y. Komura, 1966). The space o is a nuclear Bo-space universal for all Bo-spaces.
The fact that o is a nuclear Bo space is trivial. The proof that it is universal is based on the following notions and lemmas. Let X be an arbitrary nuclear Bo-space. By Theorem 7.2.7 we can assume without loss of generality that the topology in X is given by an increasing sequence of Hilbertian pseudonorms {IIxIIa}. We denote the respective inner products by (x,y)a. Let X, be the completion in the norm IIxIIa of the space X/{x a X: IIxIIa = 0}, where IIxIIa is the norm induced by IIxIIa. Since no confusion will result, we shall denote both norms by the same symbol IIxIIa Of course, Xa is a Hilbert space. Let Xa denote the space conjugate to X. Since the sequence of pseudonorms is increasing, Xa C X; for a < 9. LEMMA 7.6.2. For all positive integers a, k there are an index fl(ak) and an orthonorrnal basis {gk'"} in Xa such that, sup I n
Ink+lgn'aH
(7.6.1)
Ip (a,k) < +-0.
Proof Since the space X is nuclear, for a, k there is an index such that C1 fin(BB Ba) < nk+1
fi(a, k)
(7.6.2)
where, as usual, BY = {x a X: IIxIIY < 1}. Since Ba and Bfi are two ellipsoids, we can find a sequence {en} orthogonal with respect to the both inner products (x,y)a and (x,y)p and such that it is a basis in X. Let us assume, moreover, that {en} is normalized with respect to the norm IIxIIa i.e. IIenIIa = 1, and that the sequence {en} is ordered in such a way that the sequence {IIenIIa/llenII5} is non-increasing. Then (7.6.2) implies 1 = I Ienl Ia < n C
Let fn(x) = (x, en)a. Then, by (7.6.3), I1ffII,6 <
satisfy the lemma.
(7.6.3)
IIenI Ip c/nk+1;
therefore
fn
Chapter 7
342
LEMMA 7.6.3. If a sequence {gn'a} satisfies (7.6.1), then CO
CO
tnn
sup {
gn,a
Ifl: f ItnI2 <
1}
< +oo.
(7.6.4)
n=1
n=1
Proof. a,
CO
tnnkgn k,a
Y to k+1gnk,al nn n=1
B
n=1
Then (7.6.1) and the convergence of the series , 00 I tni imply the lemma. n=1 n LEMMA 7.6.4. For each a there is an orthonormal basis {fa.,n} in Xa such that, for all k, there is a # = fJ (x, k) such that
Ck = supllnkfa,nIIp < +oo.
(7.6.5)
s
Proof. Let us fix a and let us write gn'aby gk,n. Let us order gk,n as indicated : 91 ,1
91A
91,3
91,2
92,1---)"92,2 I g2,3 1 02,4
91,57 91,6-1
T
T 93,1__*93, 2_*g3 , 3 1 93,4
T g3,5
o°3,6
T
T
'T
1
T 92,6
g2,5
1
94,1__>94,2_>94,3__>94,4
g4,6
g4,5 T
4,
T g5,6
95,1--->95,2--->95,3-->95,4->95,6
g6,1-*g6,2-->g0,3-*g6,4-g0,8->g6,6 .
.
.
.
.
.
.
.
.
.
.
.
.
.
Let {gn} be an orthonormal sequence obtained from the one written above by the standard Schmidt orthogonalization procedure. This implies that, if m > n2, k2, then gm is orthogonal to gk, n. Since for each k, {gk,n} is an orthonormal basis, we can represent gm(m > k2) in
the form Qnm'kgk,n+
gm n>{/m
m,k 2 (an )-1
">{/m
Nuclear Spaces. Theory
343
Now Lemma 7.6.3 implies that there is a P = 9 (a, k) such that
n2ktngk,n ": f ItnI2 < 1} <+oo for m > k2+i,
sup{'
n>1m
>
and so sup{
an,gk,nllp: m=k2+i} <+oo.
n
n>lli Thus
sup m
mk I an m,k gk,n
= supllmkgmlla < +oo. M
n> I/m
Proof of Theorem 7.6.1. By Lemma 7.6.4 there are functionals Q' c, ,,} sat-
isfying (7.6.5). Let T be an operator mapping X into a defined in the following way :
T(x) = {{fi,n(x)}, {f2,n(x)}, {fs,n(x)}, ...}, where {fa,n(x)} E L2(nm).
Then II7'(x)lla,k =
LU n=1
[,.y (nklfa,n(x)I)2,1/2 = n=1
1/2
n (nk+11 fa,n(x)I )211/2 1/2
[( n2)sUp(nk+llfa,n(x) )2] n-1
/
< ]/6 Ck+1IIXIIR,
where # = i9 (a, k+ 1) and Ck+1 is given by formula (7.6.5).
Hence T(x) e o and T is a continuous linear operator mapping X into o. On the other hand, 00
IIT(x)Ila,o = [f Ifa,n(x)I2j1' = IIxIIa. n=1
Therefore, the operator T-1 is also continuous. Problem 7.6.5. Suppose that a nuclear space X does not contain the space (s). Do we have dimjX < dimjL2(nm) ?
Chapter 8
Nuclear Spaces. Examples and Applications
8.1. SPACES OF INFINITELY DIFFERENTIABLE FUNCTIONS
Let Ek be a k-dimensional real space. By Co (Ek) we denote the space of infinitely differentiable functions which are periodic with respect to each variable. For simplicity we shall assume that all those periods are equal to 27r.
We determine the topology in Co (Ek) by the sequence of the pseudonorms (8.1.1)
lIxIIn = sup Ix(n)(t)I, teEk
where n = (n1, ..., nk), nj are non-negative integers, t = (tl, ..., tk) and an,+...+nk
on,...ask
x(t)
Let us consider in Co (Ek) a sequence of inner products n
n
n
(x, y)n = f f ... f x(n)(t)y(n)(t)dtj ... dtk.
(8.1.2)
The Hilbertian pseudonorms Ilxlln = I/ (X-' x)n
define a topology equivalent to the original one. Indeed, jjxIIn < (2t)kjjxjjn.
(8.1.3)
On the other hand, there is a point to = (ti, ..., tk) such that Ix(n)(to)I <
1
7r)k
(8.1.4) (2Ilxll:
344
Nuclear Spaces. Examples and Applications
345
Moreover, tk
tl
x(n)(t)= x(n)(to)+ f ... f x(n+1)(t)dt, ... dtk, tl
(8.1.5)
tk
where n+l = (n,+1, ..., nk+1). Without loss of generality we can assume that
i= 1,2,...,k.
ti-til <21t,
Therefore, by the Schwartz inequality and (8.1.4) we obtain (8.1.6)
IxIIn < (2n)k
It is easy to verify that the elements
en = exp i (n t) = exp i (n, t,+... + nk tk) are orthogonal with respect to all the inner products and that their linear combinations are dense in Co (Ek). Therefore, the sequence {en} constitutes a basis in Co (Ek). Hence Proposition 7.4.6 implies PROPOSITION 8.1.1. The space Co (Ek) is isomorphic to the space M(am, n) where
am,n=n1 ...nkk Proof. aml+...+mk
IIen1Im = sup teEk a" til
...
a tkmk
nn l ...
nk
D
In Proposition 8.1.1 we have considered the space of complex-valued functions. But this Proposition also holds for real-valued functions. For the proof it is enough to replace exp i (n t) by sin (n t) and cos (n t). Let us now consider the space C°°[-1, 1] (see Example 1.3.7). We determine the topology in C°°[-1, 1] by a sequence of pseudonorms xIIn = sup Ix(n)(t)I 1
ItI
The system of pseudonorms n
Ilxlln = f Ilxlli i=1
(8.1.7)
Chapter 8
346
yields a topology equivalent to the original one. We shall show that the Tschebyscheff polynomials Tn(t) = cos(narccost)
constitute a basis in the space C°°[-1, 1]. The proof is based on the wellknown Markov inequality, which we give here without proof. LEMMA 8.1.2 (Markov inequality). Let P(t) be a polynominal of degree n defined on the interval [a, b] and let jP(t)j < M. Then
dPtIC2Mn2 dt
() < b-a
As an obvious consequence, we obtain by induction LEMMA 8.1.3 (Markov inequality). If P satisfies the assumption of Lemma 8.1.2, then (bmamn2m .n2(n-1)2... (n-m+1)2 < IP(m)(t)l < (8.1.8)
Now we shall prove
[-1,1], then there are constant Ck (k = 0,1, ...
LEMMA 8.1.4. If f c- C
such that sup
I
dkk f(cos 9)I < Ck4I f ft k
o-,e_,,,d0
(8.1.9)
Proof. The lemma follows from the fact that k
dd kk
ti
f(CoSO) _ :=o
d tt
f(t) I t=cosewk,d(t),
(8.1.10)
where wk, {(6) are trigonometrical polynomials depending only on k and i. PROPOSITION 8.1.5. Let f e C °°[-1,1] and let us write 1
an =
1
7r
-
(f(t)T"(t)dt. -1
1/1-t2
Nuclear Spaces. Examples and Applications
347
Then
an <
Ck !IflIk, nk
k = 0, 1, ...
(8.1.11)
Proof. Integrating by parts, we obtain 1 1
an =
f(t)T (t) dt = 1
Jvl_12 _1
a
J
f(cose)cosnOd&
0
n
n
f 0
f
for even k,
d
0
dk 17Lnk .J l
dOk
f(cosO)sinnOdO
for odd k.
o
Therefore, by (8.1.9) we obtain (8.1.11). M
THEOREM 8.1.6. For each f e C °°[-1,1] the series
a2 Tj(t) is absolutely i=0
convergent in C`°[-1, 1].
Proof. Let r be an arbitrary non-negative integer and let k > 2(r+1). Then, by (8.1.11) and (8.1.8), 00
Cfk i2r
i2
<+00.
THEOREM 8.1.7. The Tschebyscheff polynomials T. constitute a basis in C°°[-1, 1].
Proof. By Proposition 8.1.6 it is sufficient to show that the series00I a. T. n=1 co
bnTT = f then b = an. We know (Proposi-
converges to f and that if n=1
Chapter 8
348 w
tion 8.1.6) that the series I an Tn is convergent to g e C w[-1,1]. We n=1
shall show that f = g. For this purpose we shall introduce an inner product in Cw[-1, 1] 1
dt. (x,Y)f x(t)v(t) 1/1-t2 The Hilbertian pseudonorm IIxII = j/(x,x) is continuous in C'[- 1, 1], because 1
Ilo[_1
1-t2
11/2.
anTn-g = 0.
lim
n-w
d t
(8.1.12)
i=1
Let X be the completion of the space C w[-1, 1] with respect to the norm IIxII Tschebyscheff polynomials {Tn} constitute an orthogonal basis in X. Moreover, the coefficients of the expansion of the function f with respect to this basis are just an. This implies that
lim I f an Tn f = 0.
(8.1.13)
nr-sw i=1
Hence, by (8.1.12) f = g. Moreover, the uniqueness of expansions in X implies the uniqueness of expansions in Cw[-1, 1]. PROPOSITION 8.1.8. The coefficients of expansions with respect to the basis {Tn} cosntitute the space M(nm).
Proof. Formula (8.1.11) implies that if {an} is a sequence of coefficients of a function f e C'[- 1, 1], then {an} e M(nm). Conversely, if {an} a M(nm), then by a similar argument to that used in Proposition 8.1.6 we find that
{an} are coefficients of the expansion of a function f e C w[-1, 1] with respect to the basis {Tn}.
Let K = {(x1, ..., xk): Ixil < 1 (i = 1, 2, ..., k)}. By a similar argument to that used for one variable we obtain
Nuclear Spaces. Examples and Applications
349
THEOREM 8.1.8'. Let n = (n,, ..., nk). The sequence en(t) = Tn1(t,)... Tnt(tk) constitutes a basis in the space C°3(K). The basis coefficients constitute the space M(nm... nk ). COROLLARY 8.1.9. The space C°°(K) is isomorphic to the space M(nm), where n, m are positive integers. Proof. Let us order n = (n,, ..., nk), n{ being non-negative integers, in
a sequence {ap} in such a way that {(n,,p+1) ... (nk,P+1)}, where ap _ (n,,p, ..., nk,p), is a non-decreasing sequence. Then pl'k < (n,,p+1) ... (nk,p+1)
Ogrodzka (1967) showed that the spaces C°°(M) of all infinitely differentiable functions on a finite-dimensional compact manifold M are represented by the matrix [nm], and therefore all those spaces are isomorphic to one another. Now let us consider the space cS (E) (see Example 1.3.8) of all differentiable functions defined on the whole real line E and such that (8.1.14)
IIXIIn,nt = sup I tmx(n)(t)I e
with the topology defined by the pseudonorms IIxIIn, m. For convenience we shall introduce in cS (E) Hilbertian pseudonorms. W
IIXIIn,m =
[ f Itmx(n)(t)I2dt]1'2.
(8.1.15)
The system of pseudonorms {IIxII,,m} yields a topology equivalent to the original one. In fact 00
(II4n,m)2 = f I tmx(n)(t)I2dt
= f I t mx(")(t )I2dt+ Itl>1
=f
f
I tmx(n)(t)I2d t
191 <3
t2tm+lx(n)(t)I2dt+ f Itmx(n)(t)I2dt
191 >1 IIxIIn,m+1
fr
Itl>1
ItI
1
dt+IIXlln,m <
2(IIXIIn,m+l+Ilxlln,m)
Chapter 8
350
On the other hand, t
t
f tmx(n)(t)dt < f Itmx(n)(t)Idt -00
-00
< fItmx((t)Idt+ ItI>1
fitmx(n)(t)!dt= ItI_<1
r
Itm+lx(n)(t)ldt+
f Itmx(n)(t)ldt ItI41
ItI>1
00
00
<( ItI>1 J < 2 (I
dt)2(
Itm+1x(n)(t)I2dt)1/2+2I .1
fltmxn)(t)12d t1
\-ao
-00
IXIIn,m+1+IIXI In,m)
Moreover, integrating by parts we obtain t
t
f tmx(n+1)(t)dt
= tmx(n)(t)It ..- f
mtm-lx(n)(t)dt.
Thus t
e
IIXIIn.,m<supl f tmx(n+1)(t)dt +supml f tm-1X(n)(t)dt t
-0
2
-00 [IIXIIn+1,m+1+IIXI n+l,m+m (I
IXI In,m+IIXIIn,m-1))
Therefore, the systems of pseudonorms {IIxlln,,n} and {IIXIIn,m} yield the equivalent topologies. The system of pseudonorms {llxltn,m} is Hilbertian,
but it is still not convenient enough, because it is ordered into a double sequence. We shall now consider in the space cS (E) an operator D called the Hermitian derivative (see Antosik, Mikusiriski, 1968) and defined as follows :
D(x) = tx(t)- dt x(t). The operator D is continuous as a sum of two continuous operators. Let IXIt = IIDt(x)Ilo,o, i = 1, 2, ...,). The sequence of the pseudonorms {Ixlt} yields a topology not stronger that the original. We shall show later that the two topologies are/ equivalent. Let us consider the Hermite functions
Hn(t) = expl 2)
dt exp(-t2),
(8.1.16)
Nuclear Spaces. Examples and Applications
351
and let
hn(t) _
(-1)n (2nn 1)1/2 X1/4
Hn(t).
(8.1.17)
By a simple calculation (which can also be found in many text-books of the calculus) we find that {hn(t)} is an orthonormal sequence with respect to the inner product induced by the pseudonorm IIxIIo,o By the differentiation of (8.1.16), using (8.1.17), we obtain dthn(t) =
-/2(n+1)
hn+1(t) +thn(t).
(8.1.18)
Thus
hn+1(t) =
1
D (hn) .
This implies that {hn} are orthogonal with respect to all inner products induced by the pseudonorm Ixls and that, if a series E anhn is convergent n=1
with respect to the pseudonorms I xI i, i = 1, 2, ..., then {an} a M(nm). On the other hand, we have the following classical formula :
thn(t) =
11 n-+1 hn+:(t)+ 1/2
n hn-1(t).
(8.1.19)
Therefore, {an} a M(nm) implies that the series Y anhn is convergent with n=1
respect to the pseudonorm Ixllo,l, and we find by induction that it is convergent with respect to the pseudonorm I IxII m = 1, 2, ... Next, formula (8.1.18) implies that it is convergent with respect to the pseudonorms IIXIIn,m, n, m = 1, 2, ... Hence the following theorem holds : THEOREM 8.1.10. In the space cS(E) the Hermite functions {hn(t)} consti-
tute a basis. The coefficients of expansions contitute the space M(nm). The pseudonorms {IIxIIi} yield a topology equivalent to the original one.
Using this same arguments for several variables, we obtain
Chapter 8
352
THEOREM 8.1.10'. Let n = (n,,
...,
nk).The sequence
en(t) = hn,(t1)... hnm(tk) constitute a basis in the space 0 (Ek). The basis coefficients constitute the space M(nm ... nk
In the same way as in Corollary 8.1.9 we find that the space c5(Ek) is isomorphic to the space M(nm), where n, m are positive integers. Let Co [- 1, 1] denote a subspace of the space C[- 1, 1] formed by such
functions x(t) that x(n)(-1) = x(n)(1) = 0,
n = 0, 1, ...
PROPOSITION 8.1.11. The space Co [-1, 1] is isomorphic to the space 6 (E).
Proof. Let us consider a map T on the space Co [-1, 1] defined in the following way :
T(x) =
x2 arc tan t)
.
We shall show that the operator T maps Co [- 1, 1] into cS(E) in a continous way. Indeed, if x e Co [-1,1] then
Jim (t+1)mx(n)(t) = 0, n, m = 0, 1, ...
(8.1.20)
tf1->o
Let us remember that lim t±1->o
(-arctant±l) t= 2it
(8.1.21)
7G
and that n
dt x(narctant) _ I
dti x(t) i=1
Wi(t),
(8.1.22)
arctant
where wi(t), i = 1, 2, ..., n, are rational functions of t. Then by (8.1.21) and (8.1.20) we find from (8.1.20) that T is a continuous operator mapping Co [-1, 1] into cS(E). It is easy to prove by a similar argument that an operator T' defined as 79
T'(x) = x tan 2 t)
Nuclear Spaces. Examples and Applications
353
maps c5(E) into Co [- 1,1] in a continuous way and that it is the inverse operator to the operator T.
Proposition 8.1.11 can easily be extented to the case of several variables.
Let C°°(R) denote the space of all infinitely differentiable functions defined on the real line R with the topology given by the sequence of pseudonorms
IIXIIm = sup(lx(t)I+Ix'(t)I+...+Ix(n)(t)D tI<_m
In Section 7.6 we have defined the space o. Now we shall show. PROPOSITION 8.1.12 (Bessaga and Pelczynski, 1960). dims o = dims C °°(R). Proof. Let (C°°[-1,1])(8 be the space of all sequence x = {xn} (n
= 0,± 1,+2, ...) such that Xn E C °°[-1, 1] with the topology given by the pseudonorms Ixlk =
Ilxollk+IIXillk+Ilx-illk+...+IlXkllk+Ilx-kllk,
where Ix1l' are given by formula (8.1.7). By Propositions 8.1.8 and 7.4.8 the space (C°°[-1,1])(8 is isomorphic to the space o.
Let x e C '(R) and let
xk(t) = x(t+2k),
k=0,±1,±2,..., Itl <1.
It is easy to verify that the mapping U1(x) = {xo(t), xi(t), x_i(t), x2(t) x_2(t), ...} is a linear homeomorphism of the space C°°(R) into the space (C°°[-1,1])(8. Therefore
dimlC(R) < diml(C°°[-1, 1])(s) = dimlo. Let Vp(t), p = 1, 2, ... be a sequence of functions belonging to the space C °°(R) and such that (pp(t) = 1
for I t-4pnl < n,
VP(t) = 0
for It-4pnl > 2n.
and
Chapter 8
354
Let (Co (E))(8) denote the space of all sequences x = {xp} such that xp e Co (E) with the topology given by the pseudonorms IXIk = IIXOIIk+...+IIXkJIk,
where the pseudonorms IIxII{ are given by formula (8.1.1). Propositions
8.1.1 and 7.4.7 imply that (Co [-1,1])(8) is isomorphic to Q. Let x {x0(t), xl(t), ...} be an element of (Co Then the operator oo
U2(x) _
Vp(t)xp(t) is a linear homeomorphism of (Co [-l, l])(8) into p=O
C -(R).
As an obvious consequence of Theorem 7.6.1 we obtain COROLLARY 8.1.13. The space C °°(R) is universal for all nuclear B0-spaces.
Mityagin (1961) has shown that Proposition 8.1.12 can be formulated in a stronger way. Namely, the spaces o and C°°(R) are isomorphic. Since his proof is more difficult, we omit it in this book.
8.2. SPACES OF HOLOMORPHIC FUNCTIONS
Let D be a domain in a k-dimensional Euclidean space. Let p (e, z) be a non-
negative function defined for z e D and for 0 < s < 1. Let It (--,z) be non-increasing for all z with respect toe and, moreover, let
limp(e, z) > 0.
(8.2.1)
By 9(µ(D) we shall denote the space of all holomorphic functions x = x(z) defined on D such that IIxlle = suplx(z)Ip(e, z) < +oo
(8.2.2)
ZED
for all e, 0 < s < 1, with the topology determined by the pseudonorms II
Ill.
Since ,u (s, z) is a function, non-increasing with respect toe, the topology in the space 9(µ (D) may be determined by the sequence of pseudonorms {IIxII,1,j. Hence 9(µ (D) is a B0*-space.
Nuclear Spaces. Examples and Applications
355
Let AE be a non-increasing family of open sets such that
D=UAe. O<e<1 Let
µ(e, z)
{1
0
for z e Ae, for z Ae.
Then the space Q{, (D) is the space of all holomorphic functions defined
on D with the topology of uniform convergence on compact sets. We shall denote this space briefly by T(D). Since the almost uniform limit of a sequence of holomorphic functions
is a holomorphic function, the space Q((D) is complete. The spaces CIC,,(D) are also complete. In fact, if a sequence {xn} C cff,.(D) is fundamental, then it is fundamental in T(D). Therefore, it tends in CC (D) to a x (z) a R((D). Using (8.2.2), we can easily prove by a standard technique and that xn tends to x in the space' (D). that x(z) E If D is the whole k-dimensional space, then instead of T,,(D) we shall write CIY,.. The symbols C and Co will be reserved in the sequel for the whole complex plane and for the interior of the unit disc. C will denote the extended complex plane, i.e. Cu {oo}. Cr (resp. Co') will denote the Cartesian prod-
uct of r copies of C (resp. Q. In this and the next sections we shall give examples of spaces and the respective Kbthe spaces isomorphic to them. These results are taken from Rolewicz (1962). PROPOSITION 8.2.1.
The space ge(CrxCa-') is isomorphic to the space
M(am, n), where m is a positive integer, n = (n1, ..., nk) is a k-dimensional vector such that n{, i = 1, ..., k, are non-negative integers, and
am,n =
exp(m(nl+...+nr)-
m (nr+1+...+nk)
The isomorphism is given by the formula
T(f Cnzn) = {cn}, n
Chapter 8
356
where zn = zi' ... ztik Proof. As follows from the theory of analytic functions, the sequence {zn} constitutes a basis in QC(Crx Co-r). Let A,Im = {z: Izil, ..., Izrl < em, Izr+11, ..., Izkl < e-1/'n} and let jxljm = sup lx(z)l. By a simple calculation we obtain that IHznilm = am,n zeA,I.
Hence IIZnll ll+1
<
+°°.
Thus, by Propositions 7.4.1 and 7.4.6 the proposition holds.
11
Proposition 8.2.1 implies that 9C (Ck) is of type d1 and the space '3C (Co) is of type d2 , and if 0 < r < k then the space Q3C (Cr X Co -') con-
tains in the standard basis subbases of type (d1) and subspaces of type (d2). Therefore, by Theorem 7.5.2 the spaces (Ck), (Co), 9C (Cr x x Co-') are not isomorphic to each other. PROPOSITION 8.2.2. Let
µ(e, z)
= exp(t1+E)IziIP'), i=1
where Pj>0, j = 1, 2, ..., k,
t1=...=tr=0, tr+i...... k>0. Then
the space T,,, i.e. the space of all entire functions of the order p = (p1, ... ..., pk) and of the type t = (x1i ..., tk), is isomorphic to the space M(am,n), where
am,n = exp( m(n1+...+nr)- m (nr+i+...+nk)). The isomorphism T is given by the formula
T(' cnzn) = {dncn}, n
where k
d=
k
Y
j-1 P1
k
f
l /`T+1LtjJ1/n
Nuclear Spaces. Examples and Applications
357
Proof. By definition k exp(-.
lznjjt = sup jznj
(Tj+E)IzjIp')
ZECk
k
= supexp(f njlogtj-(Tj+e)t;'). r»o
j=1
Let
fj(t) = njlogtj-(Tj+E)ti',
j = 1, 2, ..., k.
We are looking for the maximum of the functions fj(t). For this purpose we calculate the derivatives of fj(t ) d
n1
p
atfj(t) = t. - (Tj+e)Pjt)Hence
f ,(t) is equal to zero only at the point to
=(
llpJ
nj
Pj(T+E))
Therefore k 1
nj/p1
dJj=1(tj +E) II
and this trivially implies the proposition. PROPOSITION 8.2.3. Let k
µ(E, z) = exp(-f(Tj+e)IloglzJHP), j=1
where pj > 1, j = 1,2, ..., k, T1 = ... = -Cr = 0, r,+,, ..., Tk > 0. Then the space 9C, of functions of the logarithmic order p = (pr, ..., pk) and of the type T = (r1, ..., Tk) is isomorphic to the space M(am, n), where
am,n = exp(m(n?+...+n:')- m (n°+i +...-l--nkk), and 4j =
1-Pj
Chapter 8
358
The isomorphism T is given by the formula
T(' cnzn) _ {dncn}, n
where k
do=
l9f
exp('1
(
n'' Pi l
(J)
Proof. By a method similar to the one used in the preceding proposition, we can calculate that k
IIznIIe = maxexp f (ntlogtj-(rj+e)logtplP, ). t»o f=1 Let us put u1 = logtj and let
fo = ntu5-(tj+e)uP . Calculating the derivative, we can prove that this function reaches the maximum at the point 1
u) _
n1
PJ(.ri+e)
Hence k
IIznlle = expl
1
n4,
l4
1
it+e)Pt l
` y=1
l (rJ+e) )
and this implies the proposition.
Let us remark that the space 9i considered in Proposition 8.2.3 is isomorphic to the space of all entire functions x(z) which are periodic with the period 2n with respect to each variable and are such that k Ilxlle =
suplx(z)IeXP(,.Y, (rJ+e)IlmzjlPs) z
j=1
with the topology determined by the pseudonorms IIx1Ie The isomorphism is given by the formula
U(x) = x(e'z...... eiz-).
Nuclear Spaces. Examples and Applications
359
PROPOSITION 8.2.4. Let k
µ(e,z)=exp(-
_ 81
1
sj>0,j=1,...,k.
log Iz1l/
5=1
Then the space ck (Co) is isomorphic to the space M(am,n), where k
am, n - exp I
m'Y n,'
and
1
qj
s1+ 1
j=1
The isomorphism T is given by the formula
T(Y, cnzn) _ {Cn}. n
Proof. k
IPz=
p Iznl sE
8,
(log
exp(_e
Iz_i
j=1
0
= maxexp (-e y(njuj+u7 )) , j=1
u,>0 1
where uj = log Izjl
Let fj(u) = nju+ ue, . By calculating the derivative we find that the 1
maximum of the function fl(u) is reached at the point
sj 8,+1 n1
1
value is nq, s
and its
/
1
(sj 8,+1 +s8,-°'). 1
k
nf (e
Therefore, 11zn9Le = exp
8,+1
1
[sj8,+1
+sf Q'])), and this implies
j=1
the proposition. ``k
Let µ1(e,z) = exp Then the space
'., (Ca) is isomorphic to the space clP.,,(C) described in
Chapter 8
360
Proposition 8.2.4. The above statement follows directly from the fact that 1
1-t
lim logl/t = j,_0
1.
PROPOSITION 8.2.5. Let k
exp(-Y, IzjIP1+e),
u(e, z) =
j=1
where pl = ... = Pr = 0 and Pr+1,
, Pk > 0. Then the space %C,. of
entire funct ions of order p = (p1, ... , pk) is isomorphic to the space M(am, n), where n.+i n, am,,n = (n1 ... nr nr)m(nr+l ...
k
-1/m
The isomorphism T is given by the formula
T(' cnzn) = n
where k
n1
do =
(nj) P1
j =r+1
Proof. As a consequence of the calculations given in the proof of Proposition 8.2.2, we obtain k Ilznlle =
J7
nj
n1
np1+8[e(Pj+e)] P1+e.
j=1
Hence for arbitrary positive 77 for sufficiently large n k
k
n1
H7
,r_P1+e+',
IIZnlle
j=1
C
n1 njP1+e+h
j=1
This trivially implies the proposition.
O
PROPOSITION 8.2.6. Let k
it(e, z) = exp(-f Ilogizjl j=1
1P1+e),
Nuclear Spaces. Examples and Applications
361
where pi = ... = Pr = 1 and pr+1, , Pk > 1. Then the space C)C of all holomorphic functions of logarithmic order p = (pl, ..., pk) is isomorphic to the space M(am,.), where m a2,,, = exP (nt - ...
m
E nr
q.+1-1/m nr+1
...
nkqm-'IM)
P!
1), j = r+1, ..., k. The isomorphism is given by the formula
(qj denotes the number
P9
T(2 Cnzn) = {C.}. n
Proof. As a consequence of the calculations given in the proof of Proposition 8.2.3 we obtain p1+e
k
jjznIle
np
= 7=1
exp ( P9+E
P3+E )p,lP5-l+)
Hence for each positive q for sufficiently large n
H
k
p1+e-h
k
exp(njP1+e-?j-1
P1+8+*
< 11zn11e <]I exp(ny1+e+7-1
3=1
9=1
This trivially implies the proposition.
8.3. SPACES OF HOLOMORPHIC FUNCTIONS. CONTINUATION
For further considerations the following lemma about Kdthe spaces will be useful. LEMMA 8.3.1. Let
am,n = exp[m(n1+r...+nk)]7 ll1\
(resp. am,n = exp[ - m (ni+...+nk)11 Then the space M(am,n) is isomorphic to the space M(am,r,), where n are
Chapter 8
362
non-negative integers, m are positive integers and k
am,n = exp (m j/n
(resp. a' n n= exp
/ I-
k
n 11 I I. In
Proof. Let [k] be the number of all systems of non-negative integers (n1,
..., nk) such that n1+ ... +nk = j. It is easy to verify that i [k]
[kp 1]
P=O
(8.3.1)
Let us write
fk(t) =
[Et],
where, as usual, Et denotes the greatest integer not greater than t. Formula (8.3.1) implies that n+1
f fk-1(t) d t = fk(n)
(8.3.2)
0
Therefore e
f fk-1(t)dt C fk(t). 0
Hence by induction we obtain k
fk+1(t) > k l
(8.3.3)
.
In particular, jkx!
[k+1] >
(8.3.4)
On the other hand, we have a trivial estimation
[](J+1)k2ik,
j = 1, 2, ...
(8.3.5)
Formulae (8.3.4) and (8.3.5) imply that we can find a one-to-one func-
tion p(n) = (pi(n), ..., pk(n)) mapping non-negative integers n onto a set of k-dimensional vectors consisting of non-negative integers such
Nuclear Spaces. Examples and Applications
363
that there are two positive constants A and B such that
AVn
am,n = expmj/n
resp. am,n = expl
-n
PROPOSITION 8.3.3. Let D be a bounded domain in the k-dimensional com-
plex space such that
aDCD for all jal < 1, where, as usual, D denotes the closure of the domain D. Then the space 9e (D) is isomorphic to the space M(am, n) where I/
am,n=expl-
n
1
I.
Proof. Let D. _ (1-e)D, where 0 < e < 1. We introduce inner products (f, g) = f f(zi, ..., Zk)g(z1, ..., zk)dx1 ... dXkdyl ... dyk, D.
where zp = xf+iyt, j = 1, 2,..., k, and the integral is taken over the domain De as a domain in the 2k-dimensional real Euclidean space. The family of the Hilbertian pseudonorms IIxIIe = 1/(x, x)e yields a to-
pology equivalent to the original topology. Indeed,
IIxIIe < IDCI11x118,
where IDES denotes the volume of De. On the other hand, by the Cauchy formula for several variables, we obtain IIxIIe < 71-1 reel Ixjle., where rEe, de-
notes the distance between the set D, and the complement of the set De,. Let f and g be monomials of the degree a and # respectively. The domains D and the volume elements are invariant with respect to the transformation z = e{tz' (i.e. z1 = ettz,, j = 1, ..., k). Therefore
(f, g) = (f(ettz')g(eUZ )) = ett(a 1)(f, g). This implies that, for or
8, (f,g)E = 0.
Chapter 8
364
Using the standard Schmidt orthogonalization, procedure, we obtain a basis of homogeneous polynomials in the space H2(D) of square intergrable analytic functions orthogonal with respect to the inner product (fig)o. We shall show that this basis is orthogonal with respect to all inner
products (f,g)E, 0 < E < 1. Let f be a homogeneous polynomial of degree a and let g be a homogeneous polynomial of degree 9. Then
(f, g) = f fgdxl ... dyk D,
= f f((1-E)z)g(1-E)z)(1-E)-2kdx1,... dyk D,
_ (1-E)a
2k(f, g)0.
{
Therefore, if (f g)o = 0, then (fig), = 0, 0 < e < 1. Obviously the number of linearly independent homogeneous polynomials of degree j is equal to
[1 This trivially implies that the space 9t (D) k
is isomorphic to the space M(am, n), where am, n = exp
(-
n
m
El
PROPOSITION 8.3.4. Let D be a bounded domain in a k-dimensional complex
space and let there be positive integers pl, ..., pk such that, for every real t, z = (z1, ..., zk) a D, then
zlexp(ipit), ..., zkexp(ipkt) e D,
(8.3.6)
and, moreover, for all reals r, 0 < r < 1,
rD C D.
(8.3.7)
Hence the space ck (D) is isomorphic to the space M(am, n), where am,n = exp (-
k1/n
Proof. Let 0(z) = z(zi', ..., zkk). Conditions (8.3.6) and (8.3.7) imply that the domain D* = 0-1(D) satisfies (8.3.4). Let U(x) = x(P(z)). It is easy to verify that U is an isomorphic mapping of the space 9 (D) onto a subspace X of the space 9e (D*). In the same way as in the proof of Proposition 8.3.3 we can construct
in X a basis formed by homogeneous polynomials. The elements of the
Nuclear Spaces. Examples and Applications
365
basis are sums of monomials zit ... zkk such that nj is divisible by pt, i = 1, 2, ..., k. Let Sy be the number of linearly independent polynomials of this type of degree not greater than j. In the same way as in Lemma 8.3.1 we can prove that there are two positive constants A and B such that Ajk+l < Sj < Bjk-Fl
and this implies the proposition. Zahariuta (1967) has proved that if D is arbitrary k-dimensional convex bounded domain, then the space BC(D) is isomorphic to the space M(am,n),
where am,n = exp
\ m /
The following stronger result has been proved by Mityagin and Henkin (1970). Let D0 be a Cartesian product of a finite numbers of pseudoconvexes bounded domains and let M be an n-dimensional analytic manifold contained in D. Then 9((M) is isomorphic to c3C (Co) (see also Zahariuta, 1970b).
PROPOSITION 8.3.5. Let D be a finite connected domain of dimension 1. Let
Z1, ..., Z. be the components of the set C\D. Then 1 ° if all Zi are points, then the space Q((D) is isomorphic to the space Ck (C),
2° if all Zt are continua, then the space ck (D) is isomorphic to the space W (CO),
3° if among Z there are points and continua, then the space CC (D) is isomorphic to the space 9((C) x W (CO).
Proof. According to the Riemann theorem on conformal mapping, we may suppose that the component Z. is either (a) the point {oo}, or
(b) the exterior of the unit disc, Z. = {z: Izi > 1}. In both cases there is a real number r greater than 1 such that every x(z) e c3C (D) can be expressed by the Laurent series X(Z) n=0
`1 bn anz"+G !! Zn n=1
Chapter 8
366
for Izi > r in case (a), for (1-1/r) < IzI < 1 in case (b). It is easy to verify that the correspondence 00
x.e(xl, x2),
where x1(z) = f anzn n=o
and n-1
in case (a) and )n-1 x2(z) = N bn ( r z
LJ n=1
1- -
in case (b) is an isomorphism between BC(D) and 1WC (C) x Q! (D u Zm) in case (a) and between 1W (D) and C)C (CO) x cJC (D u Zm) in case (b).
The domain D u Zm is (m-1)-connected. Hence, repeating the preceding argumentation, we find after m steps that the space Qt (D) is isomorphic to the space QC (C) x ... x T (C) x cC (CO) x ... x T (CO), where r fold
(m-r) fold
r denotes the number of those components of C\D which are points. This trivially implies the proposition.
Zahariuta (1970) gave a full characterization of the case where the space ck(D) (D being a one-dimensional domain) is isomorphic to the space cY(C0) (or respectively to the space 9C(C)). Namely, let K be a compact set such that the set C\Kis connected. The space 9C (C\K) is isomorphic to the space 9C(Co) (resp. ck(C)) if and only if there are a disc CR with radius R containing K and a harmonic function u(x,y) defined on CR\K such that lim
u(x, y) = 0
and
Izl'+Ivl'-+R'
(resp.
lim
u(x, y) = 1
(x,v)-+(zo,vo)EK
lira
u(x, y) _ --boo).
(Z,v)-(Za,vo)EK
Zahariuta (1970) has shown also that T (D) (D being a plane domain) is isomorphic to the space 9C (C) x 9C (C0) if and only if the compact set.
Nuclear Spaces. Examples and Applications
367
K = C\D can be represented as a union of two disjoint compact sets K1, K2 such that cC (C\Kl) (resp.'3C (C\K2)) is isomorphic to W(C) (resp Cly (CO))
This implies that there are plane domains D such that oaf (D) is not isomorphic to any of the spaces 9((C), 9C (C0), 9C (C) X 9C (CO). PROPOSITION 8.3.6. For an arbitrary one-dimensional domain D
dima9C(D) < dimffl((C0).
Proof. To begin with, let us consider the case where the set C\D contains at least three points. Then the Poincare theorem implies that there is an
analytic function f(z) defined on Co such that f(C0) = D. Let U(x) = x(f(z)). It is easy to verify that the operator U is an isomorphism between H(D) and a subspace of H(C0). In the particular case where C\D = {O,1,oo} the space 9C (C) is isomorphic to the space 9C(D). Then dime 9C(C) < dime 9C(C0). Let us observe that, if C\D contains either one or two points, then, by
Proposition 8.3.5, 9C(D) is isomorphic to H(C). This completes the proof. Since 9C(C) E dl and 9C(C0) E. d2, we obtain an example of a subspace of type d, of a space of type d2 (cf. Theorem 6.7.12). By similar arguments to those used in the proofs of Propositions 8.3.5
and 8.3.6 we obtain PROPOSITION 8.3.7. For an arbitrary one-dimensional domain D
dim19C(C) < dima9C(D).
Proof. Let us suppose that a component Z of the set C\D is a point (or a continuum). Then, by a similar argument to that used in the proof of Proposition 8.3.5, we find that the space QC (D) is isomorphic to the space Rat,
9C (C) x
(D u Z) (resp. 9C (Co) x 9C (D u Z)). Therefore, dime 9((C)
< dim, `BC (D) (resp. dim, 9((C) < dim, 9C(CO) < dims 9C (D)).
In a natural way we can extend the results of Propositions 8.3.5, 8.3.6 and 8.3.7 to domains D of type
D=D,xD2X...xDk,
Chapter 8
368
where Di, i = 1, 2, ..., k are one-dimensional domains. Then we can formulate the following PROPOSITION 8.3.8. Let D1, ..., Dk be one-dimensional finite connected domains. Suppose that : 1 ° all components of the set C\Dj are points for i = 1, 2, ..., r, 2° all components of the set C\DA are continua for i = r+ 1, ..., r+p,
3° among the components of C\Dj there are points and continua for
i=r+p+1, ..,k.
LetD=D1x ...xDk. Then the space T (D) is isomorphic to the space re
7((C'x Cp-')x
(C'
X Cp-t-1)x... x
-7L(Ck-Px C'P')
Zahariuta (1974, 1975) proved that the spaces c3C(Cr x Ck-r) 0 < r < k are isomorphic to c?C (C1 x Co -1)
Thus, basing ourselves on his result, we can formulate Proposition 8.3.8 in a stronger way. Namely PROPOSITION 3.3.8'. Under the assumption of Proposition 8.3.8, if 0 < r+
+p < k then 9E (D) is isomorphic to g (C X Cr'). PROPOSITION 8.3.9. Let D = Dl x one-dimensional domains. Then
... x Dk, where Dz (i = 1, 2, ..., k) are
dimicY(D) < dimz9Y(Co). PROPOSITION 8.3.10. Let D = D1 X ... X Dk, where D¢, i = 1, one-dimensional domains. Then
..., k are
dimlA((Ck) < dima9P(D).
Let us remark that from the proof of Propositions 8.3.5 and 8.3.8 follows
PROPOSITION 8.3.11. Let D = D1 x ... x Dk, where D{ (i = 1, bounded one-dimensional domains. Then dim1Qt'(D) = dimzQ((Co).
..., k) are
Nuclear Spaces. Examples and Applications
369
Proof. Let Z' be the component of the set C\Dj which contains the point oo. Then 9l (D) is isomorphic to the space C3C (Co x 9C (Di x ... x D'), where Da = D{ v ZI, i = 1, 2, ..., k. Therefore dimj9e(D) >, diml`)f(Co). Hence Proposition 8.3.9 implies the proposition in question. PROPOSITION 8.3.12. Let D = Dl x ... x Dk and D' = Di X ... X Dk+p, where p is a positive integer and D¢ (i = 1, ..., k), D'(j = 1, ..., k+p) are one-dimensional domains. Then the space QC (D) and Q Y (D') are not isomorphic.
Proof. To begin with, let us calculate the diametral approximative dimensions of the spaces QC (Co) and QC (Ck+P). By Corollary 8.3.2, {tn} E 6(QC(Ck+')) if and only if lim tnexp(mk++y'n) = 0 (m = 1, 2, ...) and 11X00
{tn} e 6 (T (Co)) if and only if for certain m' /
limtnexpl
k= 0.
+.j/ ,
Since, for arbitrary m, j//m' tends to infinity faster than m
n
S(W(Co)) I S(C C(Ck+P)) Thus, by Propositions 8.3.9 and 8.3.10, 6
(CM
(D)) C 6 (W
(C. k))
6 (CM (Ck+P)) c g (T (D'))
Hence, by Proposition 6.5.1, the spaces QC(D) and 9C(D') are not isomorphic. Let X be a Schwartz space. Let
r(X) = supinflimsup UV
Z- o
loglogM(V, U, e) 1
loglog 8-
where U, V run over all balanced neighbourhoods of zero. The number r(X) is called a functional dimension (see Gelfand and Vilenkin, 1961, p. 127).
Chapter 8
370
Of course, if { Ut} is a countable basis of neighbourhoods of zero, then
r(X) = supinflimsup
loglogM(U{+p, Ui, E)
E-->o
loglog
1
E
Let X = 9t (Ck) (or T(Ca)). Then, by Proposition 6.5.19 and Corollary 8.3.2,
M(Us+', Us, E) _ L1 {
2
at+E
ai,n 1=1[1+2 ex-1i/n,
Ei+ji n/J)
/I
Since 2
2
exp(-a l/n) > 1 if and only if n <
exp (-a yin) >
1
if and only if n <
/
1
(--
log
)k, e
and
2/k log E , we get
E 2
112(Alog E)k
where a = j (resp. a = limsuP C- o
1
1
k
alog ED
1
i+j ). Hence
loglogM(Ui+', Us, E)
loglog 1
= k+ 1,
and we obtain PROPOSITION 8.3.13. The functional dimension of the spaces CY(Ck) and QC (Co) are equal to k+ 1.
Komura (1966) has investigated the following problem. Let P be a differential operator with constant coefficients defined on a real k-dimensional space Rk. Let Ep be the space of all continuous solutions of the
equation P(u) = 0 defined on the whole space Rk with the topology of uniform convergence in compact sets.
Nuclear Spaces. Examples and Applications
371
Komura (1966), has proved that the following three conditions are equivalent : (1) The operator P is hypoelliptic, i.e., Ep C C (2) The space Ep is nuclear. (3) The functional dimension of the space Ep is finite
r(EE) < +oo. Moreover, if the operator P is elliptic, then r(Ep) = k. If the operator P is only hypoelliptic, but not elliptic, then this equality does not necessarily hold.
8.4. SPACES OF DIRICHLET SERIES
In this section we shall consider subspaces of the space 9t(D) of a special type, called spaces of Dirichlet series. Let An = (An, ..., Ak), Ati > 0. We shall assume that lim
logn = C{ < +00,
i = 1, ..., k
(8.4.1)
Pi
and that all An are different from one another. Let z = (z,, , .. , zk) be a point of a domain D contained in a k-dimensional Euclidean complex space. We shall write exp (A"z) = exp (Aiz1+... +Ak zk) .
By a Dirichlet series we shall mean a series of the following type : Go
Z_j
a"exp(Anz).
n=1
A Dirichlet series is called an entire Dirichlet series if it is convergent for all z e Ck. The space of all entire Dirichlet series determined by the sequence {An} will be denoted by S().
Let us remark that, if a Dirichlet series is convergent at a point z° _ (zi, ..., zj, then there is an M > 0 such that l ani lexp(Anz°)l < M. Let
Chapter 8
372
z = (zj, ..., zk) be such a point that
i= 1,2,...,k.
Rezi
(8.4.2)
Then, for sufficiently large, n k
1
IaneXp()nz)I <Mexp(--,Y-, 3.%'Cj) k
<M i=1
1
nl ... nkz 2
00
Therefore, the series Y anexp(Anz) is convergent. n=1
This implies that an entire Dirichlet series is uniformly convergent on all compact sets (i.e., in the topology of the space 9C(Ck)) if and only if it is uniformly convergent on the sets
A. = {z = (zi, ..., zk) : Re z{ < m). Let us introduce the topology in S(An) by a sequence of pseudonorms IIxIIm = sup 1x(z)1 zeAm
As follows from the preceding considerations, the pseudonorms
I1x11+n
yield in S(an) a topology equivalent to the topology in 9C(Ck). We shall show that the sequence en = exp(2nz)
constitutes a basis in Scan. In order to prove this fact we introduce in Scan> a sequence of inner products T
T
1
(x , y)m = lim sup (2T) f ... f X(m+itl, ..., m+itk) X Tc. T
-T
xy(m+itj,..., m+itk)dtl ... dtk. The topology determined in S() by the Hilbertian pseudonorms IIxIIm =1/(x, x)m is equivalent to the original one. Indeed, IIxIIm < IIxIIm.
Nuclear Spaces. Examples and Applications
373
On the other hand, if m' > m+3C{, i = 1, 2, ..., k, then 00
Y Ilanexp(An)zllm < supllanexp(AnZ)Ilm'
Ilxllm <
n=1
_ asupllanexp(t' Z)IIm' < W
where or =
'
1
2 n=1 n
.
Let us observe that en(z) = exp(Anz) are orthogonal with respect to a]t inner products (x, y).. This implies that {en} is a basis in S(2 ).
Since IIenIIm = expmlA"I, where IA" I = i+...+Ak we obtain by Proposition 7.4.6 PROPOSITION 8.4.1. The spaces S(a.) is isomorphic to the space M(am,n) where am,n = expm IAnI.
Suppose now that in condition (8.4.1) all C{ are equal to 0. In the same way as before, we can prove that if a Dirichlet series W
an exp (Anz)
(8.4.3)
n=1
is convergent at a point z° _ (z?, ..., zk), then it is convergent at each point z = (z,, ..., zk) such that
Rezi < Rez°,
i = 1, 2, ..., k
(8.4.4)
Hence, for each Dirichlet series (8.4.3), there is a system of real num-
bers R = (R,, ..., Rk) such that the series (8.4.3) is convergent for all z = (z,, ..., zk) such that Rezi < R¢, i = 1, 2, ..., k and it is divergent for all z = (z,, ..., za) such that Re zI > R{, i = 1, 2, ..., k. The vector R is called the abscissa of convergence. Obviously some R{
may be infinite. Let us assume that R{ = -boo for i = 1, 2, ..., r and Ri < +oo for i = r+ 1, ..., k. By S( ,t.) (R) we shall denote the space of all Dirichlet series with the sequence of exponents {An} and the abscissa
Chapter 8
374
of convergence R, with the topology induced by the space (D), where
D={z=(zl,...,zk): Rezi
i=1,2,...,k}.
In the same way as in the case of the space Sung we can show that this topology is equivalent to the topology determined by the pseudonorms
114. = zcA. sup WO, where
Am={z=(zl,...,zk): Rez{<mfori=l,2,...,rand Rezi
(x, y)m = lim sup T- -co
T
fxrn + it1, ..., m+itr, Rr+l-1/m
2T -T j... -T
+itr+i, ..., Rk-1/m+itk)Y(m+itl, ..., m+itr, Rr+1 -1/m+itr+r...... k-1/m+itk)dti ... dtk. In the same way as before, we can show that the Hilbertian pseudonorms IIxIIm = 11(3c,x)m yield a topology equivalent to the original one. Let us observe that en = exp(Anz) are orthogonal to one another with
respect to all inner products (x, y)m. Thus {en} is a basis in the space S() (R) and the following proposition holds : PROPOSITION 8.4.2. The space S(2.) (R), where R, = ... = Rr == +oo and Rr+,, , Rk < +oo is isomorphic to the space M(am, n), where
am,n = exp m(Ai+...+Ar)-
m
The isomorphism T is given by the formula 1 C n exp (Anz)) = {d n cn} ,
T( n
where do = exp(.1,+1Rr+1+ ... +AkRk). Proof. IIenIIm = expm(2 + ... +.1;+2+i(Rr+1-1/m)+2k(Rk-1/k)) = dnam,n. Thus Proposition 8.4.7 implies the proposition.
Nuclear Spaces. Examples and Applications
375
w
Let µ(s,z) = exp(- E (rj+r) Iz,I "), where pj > 1, j = 1, 2,..., k, j=1
T1 = ... = Tr = 0, -r,+1, rk > 0. The space T. is the space of the type r = (r1, ..., rk). By Six j we shall denote the subspace of the space sequence {en} = {exp()nz)}.
%
spanned by the
PROPOSITION 8.4.3. The space S(' j!) is isomorphic to the space M(am,n), where
m
am ,n
L(M+1)q,+1+... r"(Ak)9k!)
and
j= 1,2,...,k.
-
q5- pp1
The isomorphic T is given by the formula
T (2: en exp (t.-z)) = {dn cn}, n 4l 1
where do = exp( '
1
1
n\4f (
pj / qs ( Tj )4f) . Proof. The sequence {en} is a basis in S('j . Moreover, replacing log IzjI by Rezj in the calculation in the proof of Proposition 8.2.3, we obtain j=r+1
k
11en11e = expl
1j=1
1
(4.)
q,(
1
1
Tj+E
q
(1Pj
)Q1).
Thus, by Proposition 7.4.7, the proposition holds.
Let k
exp(-f ,u(--, z) = j=1 Pr+1, ... , Pk > 1.
1Z51PJ+E),
p1=...=Pr= 1,
The space Q1 is the space of all entire functions of the order p = (pi, ..., pk). By S(P,, we denote the subspace of the space %C spanned by the elements en = exp(Anz).
Chapter 8
376
PROPOSITION 8.4.4. The space S(x is isomorphic to the space M(am,n), where
+...+(Ak4k-1
am,n = exl)[ Pi
and qj =
,
.1 = 1, 2, ..., k.
The isomorphism T is given by the formula
T(E cnexp(t11z)) = {cn} . n
Proof. The sequence {en} is a basis in the space S( ). By a similar calculation to that used in the proof of Proposition 8.2.6 we obtain
- j-1 exp k
II
en ll
p3+8
n
pf-1+e pj-1+E Pj+E Pj+E 2j
and this implies the proposition. The spaces of Dirichlet series of one variable have been investigated by Srinivasan (1966).
8.5. CAUCHY-HADAMARD FORMULA FOR KOTHE POWER SPACES
Let us recall (see Section 7.5) that a space M(a), where an->oo, is called a Kothe power space of infinite type, and that a space M(an 1/m) is called a Kothe power space of finite type.
THEOREM 8.5.1 (Cauchy-Hadamard formula ; Rolewicz, 1962b). Let m
am.n = dnan
(or am.n =
dnan-1/m),
where m is a positive integer, n = (n1i ..., nk), nj being non-negative integers, lim an = +oo. Then a sequence x = {xn} belongs to the space M(am,n) if and only if 1
lim I dnxnl loea = 0 n-a oo
1
(resp.limsupIdnxnl'Oea" <+1). n-oo
Nuclear Spaces. Examples and Applications
377
Proof. Necessity. Let x = {xn} e M(am,n). Then for each in there is a constant M. such that (resp. I dnxnan l/mI < Mm)
Idnxnan I <Mm
Hence d o xn I
toga" em < M m g
(resp.
I d o xn I
logo" e _ 1/m
< M 'gO'
Since 1
lim
1,
n--a ao
we have 1
m = 1, 2, .. .
lim sup I do xn l l°ga" < el--, co
1
(resp. lim sup IdnxnI t°ga" < el/'n,
in
= 1, 2, ...,)
n-+co
and this trivially implies the conclusion. Sufficiency. Let 1
1
limsupldnxnl logo" = 0
(resp. limsupldnxnll°ga" < 1). n-iao
n--ioo
Then for any integer in there is an integer N. such that if Inil+ ...
... + I nkI > Nm, then i
IdnxnI logo" < e-m
(8.5.1)
1
(resp. Idnxnllogan
(8.5.2)
<ell-).
Let Mm = sup I dnxnl an,n+e-m (resp. Mn = sup I dnxnl am,n+ellm), InI
lnl-
where, as usual, InI = Inil+ ... +Inkl. Then by (8.5.1) (resp. (8.5.2))
0
I doxn am,nI < Mn, and this implies the conclusion. COROLLARY 8.5.2. A function 00
n
nk
xoZ
X(Z) n
Zk
n1....,nk=0
Chapter 8
378
is an entire function if and only if 1
where Ini _ Jnli+...+lnkl.
lim IxnI In, = 0, Inj- co
This is an immediate consequence of Proposition 8.2.1 and Theorem 8.5.1. COROLLARY 8.5.3. A series 00
fXnZn=
X(Z) = n
ni ..., nk ZI ... Zk I Xn, ni,...,nk=0
is convergent in a polycylinder
D={z=(z1i...,zk): Izil
i=1,2,...,k}
if and only if 1
limsup Ixni
1 R1n,...
n ao
nk
Rk
II <
This is also an immediate consequence of Proposition 8.2.1 and Theorem 8.5.1. Corollaries 8.5.2 and 8.5.3 give the classical Cauchy-Hamadard formulae. As a consequence of Proposition 8.2.2 and Theorem 8.5.1 we obtain trivially the following two corollaries : COROLLARY 8.5.4. Let k
X--, Z) _ 11 exp(-eizjIP'). j=1
Then the power series
X(Z)=
-
xnZn
=
n
n,
nk
Xf...... nkZ1 ... Zk ni.....nk=0
represents a function x (z) e W,,, i.e. x (z) is a function of the order p _ (P1, , pk) and of the minimal type if and only if limldnxnIIn] = 0.
(8.5.3)
Nuclear Spaces. Examples and Applications
379
where 11k
n,/Pi
do
= 1_i
(epj)
For k = 1 (8.5.3) gives the classical formula lim
(n) P
n
I
P
= 0.
(8.5.3')
COROLLARY 8.5.5. Let k
p(e, z) _ fj expl-(i1+E) IzjIP'], j=1
where pi > 0, Ti > 0 (i = 1, 2, ..., k). Then a power series x(Z)
x,aZ
n
,akZ1nt
Xn......
...
Zknk
n
represents a function x(z) e 9e,,, i.e. a function of the order p = (P1, ....'Pk) and of the type T = (r1i ..., xk) if and only if 1
limsupldnxnllnl < 1,
(8.5.4)
n- co
where k
j=1
dnj
)fli/Pi
epjTj
In the particular case of k = 1 we obtain the classical formula limsup i/Ix,yl n11P < (Tpe)11P.
(8.5.4')
Formulae (8.5.3) and (8.5.4) have been obtained in a different way by Goldberg (1959, 1961).
As a consequence of Proposition 8.2.3 and Theorem 8.5.1 we obtain the following two corollaries :
Chapter 8
380
COROLLARY 8.5.6. Let y
jC(e, z) = exp
Ilog Izj! 1P1)
.
1=1
Then a function 00
n,
n XnZ =
Z= X()
xni,...,neZl ...
Zkn
n
belongs to 19N if and only if lim Rej/Fxni = 0, where
91= pPi 1,
.1 = 1,2,...,k
(8.5.5)
and nQ = nlQ.+...+Qnk= .
(8.5.6)
COROLLARY 8.5.7. Let k
p(e,z) = exp(t1+E)IloglzljlP'). 1=1
Then an entire function X(Z)
=
XnZ
n
ni
n,,...,ns=o
n
belongs to %3C if and only if 1
limsupIdnxnl n° < 1, where nQ is determined by formula (8.5.6) and
dn=11 exp(n;'g1(P1)Q'\
i)P'1).
1
J-1
In the particular case of k = 1
limsupj lxnl
<expH r)P-1
q
p)
nu
Nuclear Spaces. Examples and Applications
381
COROLLARY 8.5.8. Let k
p(e, z) = exp (-81 9=1
II
\ 1 1IztI
)d'),
where st > 0. Then a function 00
n
(Z)
xn z =
X(Z) = n
xn....., f ni,...,nk=0
belongs to the space
if and only if.
limi4xnI = 0. where
9j =
Si
s1+
l
1
ql
= 1, 2, ..., k and n 4rQ
-{-
This is an immediate consequence of Proposition 8.2.4 and Theorem 8.5.1.
As a consequence of Proposition 8.2.5 and Theorem 8.5.1, we obtain the following two corollaries : COROLLARY 8.5.9. Let k
p(e, a) = exp(f=1
Then a function 00
x(z) =
X.
zn
=
Xn.....,nk
z1ni
nk
... Zk
m,....nk=o
n
belongs to the space QC,,, i.e. x(z) is an entire function of the minimal type,
if and only if lim
nlogn IxnI
= 0,
(8.5.7)
where
nlogn = nllognl+...+nklognk with the convention 0 1og0 = 0.
(8.5.8)
Chapter 8
382
COROLLARY 8.5.10. Let k
p(e, z) = exp(-' IzjIP,+e) i=1
/
where all pj > 0, j = 1, 2, ..., k. Then a function 00
X
(Z)
x, zn
=
nL
nk
n
belongs to the space c C,,, i.e. x(z) is a junction of the order p = (pi, ..., pk),
if and only if limsup-nlognj/jdnxnj
n
< 1,
(8.5.9)
n is defined by formula (8.5.8) and k
dn=
njnfh1f j=1
In the particular case where all pj are equal to a number p, pi >_ ... = pk = p, we obtain the classical formula nlogn'/
limsupj/Jxnj <
(8.5.9)
e-1"P.
n-co
The formulae given in Propositions 8.5.8 and 8.5.9 were obtained in another way by Goldberg (1959, 1961). COROLLARY 8.5.11. A Dirichlet series co
x,,eXp(A'z)=
xn.,...,nkexp() z1
...+,kkzk)
nk = 0
is convergent for all z, i.e. it is an entire Dirichlet series, if and only if Ix"I ,
lim-1 IxnI = 0, co
where
IA"I = ii.1+...+/lk.
(8.5.10)
Nuclear Spaces. Examples and Applications
383
This is a trivial consequence of Proposition 8.4.1 and Theorem 8.5.1. COROLLARY 8.5.12. A Dirichlet series
f 00xnl....,nkexP(, Z1+...+ilkzk)
xnexp(Anz)
n,,...,nk=0
n
has the abscissa of convergence R = (R1, ..., Rk), Ri < +oo (i = 1, 2, ... ..., k) if and only if n-.w
where IA*I is defined by formula (8.5.10) and
dn = exp(rR) = exp(% R1+...+rkR,).
This is an obvious consequence of Proposition 8.4.2 and Theorem 8.5.1.
As an immediate consequence of Proposition 8.4.3 and Theorem 8.5.1 we obtain the following two corollaries : COROLLARY 8.5.13. An entire Dirichlet series 00
xtexp(A'z) = n
xn,,...,nkexp(nlzl+...+, kzk)
n,,...,nk=0
belongs to the space S(o j if and only if lim (x )° /Iznl = 0, n-.oo
where qg =
pi
1, 2, ..., k and
p9-1
(/Zn)q =
(f)4i+... + (Ak)qk
COROLLARY 8.5.14. An entire Dirichlet series 00
'xneXp(tnz)
,
n
_n,,...,nk=0xn,,...,nkexp(2lzl+...+) Zk)
Chapter 8
384
belongs to the space Stx j where pt > 1, r5 > O, j = 1, 2, ... , k if and only if
limsup-)' IxnI < 1, where k
1
PI-1
(An)q,(
bn =
.
j=1
For k = 1 Corollaries 8.5.13 and 8.5.14 were obtained in another way by Ritt (1928) COROLLARY 8.5.15. A Fourier series oo
xnexpi(n, t) =
f,
xn.,,...,nkexpi(nhtl+...+nktk)
M..... nk=-00
ri
is uniformly convergent together with all derivatives if and only if
lim
1ognj
j/TXnI
=0
where
Ilogni
logn{,
(8.5.11)
and we take the sum (8.5.12) over all i such that In{I > 2.
This is an obvious consequence of Proposition 8 1.1 and Theorem 8.5.1
Chapter 9
F-Norms and Isometries in F-Spaces
9.1. PROPERTIES OF F-NORMS
Let X be a real F*-space with norm IIxII Let fx(t) = IItxII. The properties of F-norms imply that : (1) fx(t) is a continuous function,
(2)fx(t) =f.(-t), (3)fx(0)=0,and if x# 0,fx(t)=0,then t=0, (4) fx(t1+tz)
fz t'
2
t2J > a Ux(tl)+fz(t2)),
(7) the function fz(t) for positive t is infinitely differentiable.
The proof of Theorem 9.1.1 is based on the following lemmas :
Chapter 9
386
LEMMA 9.1.2. Let X be a real F-space with norm Ilxll Then there is in X an equivalent norm Ilxll** such that thefunction t ** (t) = lltxll**is concave for positive t. Proof. Let llxll* = sup lltxll, Ilxll* be an equivalent norm and let the funco
tion fz(t) = lltxll* be non-decreasing for positive t (see Theorem 1.2.2).
Let Ilxll** = sup
b-1
Ilaxll* + 1-a llbxll*) . In other words, llxll** is
b-a
b> i>a>o (b_-_a
equal to f * (1), where f z *(t) is the smallest concave function not smaller than fx(t ).x llxll** satisfies the triangle inequality. Indeed,
+ 1-aa b
b-a
llaxll*+b1-- a
sup>a>o(bb--1
IIb(x+Y)ll* )
+ b-1 +
b-1 lla(x+Y)ll*
b>1>a>0(
IIx+YIl** = sup
b>1
llbxll*
a
a
1-a
(b-1 Ilaxll*
sup b-a b-a Ilaxll*+ b-a IlbYll*) < b>1>a>o 1-a llbxll*)
b-a
(b-1
+sup b>1>a>o b-a Ilaxll*+
1-a
b-a
lbYll*1=JIxII**+IFYII**
Since lltxll** ' f **(t), lltxll** is a concave function for positive t.
Of course, Ilxll* < llxll**. On the other hand, lltxll* is non-decreasing for positive t. Hence by the
triangle inequality, for a > 0 llaxll* = ll(Ea+(a-Ea))xll* < II(Ea)xll*+II(a-Ea)xll* Eallxll*+II(a-Ea)xll* < (a+1)llxll* 1 Then
(b-1 Ilaxll*+ 1-a llbxll*)
llxll** = sup
b>1>a>o b-a sup
b-a
(b-1(a+l)llxll*+ 1-a (b+l)IIxll*)
b>1>a>o
b-a
b-a
= 211xll*.
Therefore Ilxll** is an equivalent norm satisfying the required condition.
Let Ck(0,1) (k may also be equal to infinity) be the space of all functions x = x(t) defined in the interval (0,1) having continuous derivatives 1 Ea denotes the greatest integer not greater than a.
F-Norms and Isometrics in F-Spaces
387
up to the order k. The space Ck(0,1) is a Bo space with the topology determined by the following sequence of pseudonorms : IIxIId = sup Ix(')(t)I a
0 < a < 1,
(i = 0, 1, ..., k),
dit x(t) and x(°)(t) = x(t).
where, as usual,
Let {Ei} be a sequence of positive numbers such that co
i=1 Co
(9 1.1) trivially implies that the infinite product [J (1 -Et) is converi=1
gent. Let us denote its limit by S. Of course S < 1.
Let us define the operation UP,, acting on the spaces Ck(0,1] in the following way :
f
1
1
... U,.r(x) = Ep.. 1Ep+r 1-e9
f x(tsp ... sp+r)dsp ... dsp+r.
1-ep+r
LEMMA 9.1.3. The sequences of operators {Up,,} have the followingproperties :
(a) For every x a Ck(0,1] there exists a limit
p = 1, 2, ... ;
limUp,r(x(k)) = Up,.(x(k)),
(b) If x a C(0,1] = C°(0,1], then Up,,(x) e C'+1(0,1] ; (c) The operation U1,.,, is a continuous linear operator mapping the space
C(0,1] into the space C'(0, 1]. Proof. Let x e Ck(0,1]. By a simple estimation we obtain (9.1.2)
IIUp.r(x)IIa' < IIxIlaa.
Then, if r2 > r1i we have I I UP.r.(x)- UP,r.(x)I Ia = II UP ,.(x- UP+n+1,r.-r.(x))I I,
C IIx- Up+r.+l,r.-r.(x)IIi 3a = Sup
f 1'
X
I
1
a6
X EP+r.
1'
...
f/
Ix()(t)-x(
(tsP+r,+l ... Sp+r.)I dSp+r.+1 ...
... dSp+r,.
Chapter 9
388
p+r'
Since lim
[1
(1-ej) = 1, the continuity of the functions x(t)(t)
r1 r,-sao j=p+r1+1
implies that x({) (t)-x($) (tsp+,,+1 ... sp+,) tends uniformly to 0 for
aS < t < 1, 1-ej < sj < 1, where r1, r2->OTherefore the completeness of the space Ck(0,1] implies (a). We shall show (b) by induction with respect to r. For r = 0 it follows from the identity
tf
t
fx(ts)ds=±
x(u)du
(9.1.3)
(1-e)t
e
by the theorems on the differentiation of the integral of a continuous function. Suppose that (b) holds for r = r0, i.e., dr,+1 dtr1+1
(p = 1, 2, ...).
[Up+,,p+r,+1(x)] = F(t) e C(0, 1]
Then 1
d:,+1dtr°+1 [Up,p+ra+1(x)]
=
1
f F(ts,)dsp a C1(0, 1).
1 Elp
1-e>
Hence Up,p+,o+1(x) e C"-+'(0, 11. Then (b) holds.
Now we shall prove (c). By (a) U1,co(x) = lim U1,q(x) = lim Up,q_p(U1,p_1(x)) q-soo
q- oo
= Up,-(Ui,p-1(x)) e C2'(0, 1]
(p = 1, 2, ...).
Hence Ul,,(x) e C`°(0,1]. The continuity of Ul,,, follows from the inequality IIUi,oo(x)IIa
and from the fact that the operator Ul.p maps C(0,1] into Cp(0,1] in a continuous way. Proof of Theorem 9.1.1. Let IIxII** be a norm defined by Lemma 9.1.2 and
let f,,(t) = U1,.(fx (t)) and let IIxII = fz(1). By Lemma 9.1.3 the function fz(t) is infinitely differentiable. It is easy to verify that fz(t) is concave and
F-Norms and Isometries in F-Spaces
389
that IIxII' is an F-norm. We shall show that it is equivalent to IIxII**. Indeed, for q = 1, 2, ... inf IItxII** < Ul,q(IIxll**)II < sup IItxII**,
6<9<1
d_
where U1,q(IUxll**) = U1,q(iltxII**)It=i Passing to the limit, we get inf IItxII** < IIxII' < sup Iltxll**. Hence Ilbxll** < IIxII' < IIxII*. This im-
a,t<1
6_
plies that the norms IIxII** and IIxII' are equivalent.
Problem 9.1.4. Is it possible to replace in Theorem 9.1.1 property (7) by property (7') : the function fz(t) is analytic for positive t?
9.2. SPACES WITH BOUNDED NORMS
We say that an F*-space X is a space with bounded norms if for each norm IIxII equivalent to the original one,
sup IIxII < +oo.
xc-X
The following theorem gives a characterization of F*-spaces with bounded norms. THEOREM 9.2.1 (Bessaga, Pelczyfiski and Rolewicz, 1957). An F*-space X is a space with bounded norms if and only if for each neighbourhood of zero U there is a positive integer n(U) such that
Un =U+ ... +U = X. n(U)-times
Proof. Sufficiency. Suppose that there is in X an F-norm IIxII equivalent to the original one and such that supllxll = +oo Let U = {x: IIxII < 1}. ZEX
The triangle inequality implies that for each n Un C {x: IIxII < n} :t- X.
Necessity. Let U be such a neighbourhood of zero that Un # X, n = 1, 2, ... Without loss of generality we can assume that U is balanced.
Chapter 9
390
Let us repeat the construction of a norm described in the proof of Theorem 1.1.1. It is easy to verify that for the constructed norm IIxII
sup IIxII = +o. zeX
PROPOSITION 9.2.2. The space S[O, 1] (see Example 1.3.1.a') is a space with bounded norms.
Proof. Let U be an arbitrary neighbourhood of zero. Then there is a positive integer n such that Kl,n = {x: IIxII < 1/n} C U, where IIxII denotes the norm in the space S[O, 1] described in Example 1.2.1a'. Let x(t) be an arbitrary element of S[O, 1]. Let
xi (t) =
Inl
x(t)
for
0
elsewhere.
Obviously, x(t) = xl(t)+ ... +xn(t). On the other hand, IIx{II < 11n; hence x{ e U. This implies that Un = X.
9.3. ISOMETRIES AND ROTATIONS Let (X, II IIx) and (Y, I I I Iy) be two F-spaces over reals. We say that a trans-
formation U (not necessarily linear) mapping X into Y is called an isometry if
II U(x)-U(y)IIy = IIx-ylix. An isometry U such that U(X) = Y and U(0) = 0 is called a rotation Mazur and Ulam (1932) proved that if X and Y are Banach spaces and the norms II IIx and II IIy are homogeneous, then each rotation Uis a linear
operator. Charzyfiski (1953) proved that in the case of finite-dimensional spaces the above is true without the assumption that the norms are homogeneous. The original proof of Charzyfiski was difficult. Wobst (1973) gave a simple proof of Charzyliski's theorem. In this section we shall present extensions of the Mazur-Ulam theorem to two cases : to strictly galbed F-spaces with the strong Krein-Milman property (this result contains as a particular case the result of Charzyliski) and to locally bounded spaces with concave norms.
F.Norms and Isometries in F-Spaces
391
We say that an F-space X has the strong Krein-Milman property if each closed bounded set A C Xhas an extreme point. Example 9.3.1 Every Montel Bo space X has the strong Krein-Milman property. Indeed,
each closed bounded set A C X is compact and by the Krein-Milman theorem (Proposition 5.5.1) A has an extreme point. Example 9.3.2
Every reflexive Banach space has the strong Krein-Milman property. Indeed, every closed bounded set is compact in the weak topology by Corollary 5.2.8. Thus by the Krein-Milman theorem (Proposition 5.5.1) it has an extreme point. Example 9.3.3
The spaces LP[O, 1], 0 < p < 1, do not have the strong Krein-Milman property. Indeed, the unit ball is closed and bounded but it does not have extreme points. THEOREM 9.3.4 (Mankiewicz, 1979). Let X be a real strictly galbed F*space with the strong Krein-Milman property. Then every rotation mapping X into X is linear.
The proof is based on several notions, proposition and lemmas. Let X be an F*-space. By H(X) we shall denote the group of all homeomorphisms mapping X onto itself. By T(X) we shall denote the subgroup
of H(X) consisting of a mapping g of the form g(x) = ex+y, e = ±1, y e X. A subgroup G C H(X) is called fat if T(X) C G. By G(f) we denote the subgroup of H(X) generated by a homeomorphism.f and T(X). Let G C H(X). By Go we denote Go = {f e G: f(0) = 0}. We shall write Go(.f) = (G (.f ))0.
A subgroup G C H(X) is called equicontinuous if for every e > 0 there is a S > 0 such that sup IIf(x)II < e (cf. the equicontinuity of families flE1l
of linear operators, Section 2.2).
Chapter 9
392
If G is a fat subgroup, then
Go={geG: g(x)=f(x)f(0)for some feG}, G={feH(X): f(x)=g(x)--y for some gEGo, yc--X}. Thus G is equicontinuous if and only if Go is equicontinuous. It is clear that T(X) is equicontinuous.
By Lin (X) we denote the group of all linear homeomorphisms of X onto itself and by Aff(X) we denote the group of all affine homeomorphisms, i.e. f e Aff(X) if and only if f(x) = g(x)+y, where g e Lin(X)
andyeX. We say that an F*-space X has the affine group property if each equicontinuous fat subgroup G of H(X) is contained in Aff(X). PROPOSITION 9.3.5 (Mankiewicz, 1979). Let X be a real F*-space. Then the following conditions are equivalent : (9.3.1.i) X has the affine group property, for every fat equicontinuous subgroup G C H(X),
G C Lin(X),
(9.3.1.ii)
for ev ery f such that G(f) is equicontinuous, (9.3.1.iii)
G(f) C Aff(X), for every f e H(X) such that G(f) is equicontinuous, Go(f) C Lin (X) ,
(9.3.1.iv)
for every f e H(X) such that G(f) is equicontinuous, the mapping g(x) = f(x) -f(O) is linear and
G0(f) = G0(g) = {egn: e =
1, n = 0, ±1, +2, ...}, (9.3.1.v)
for every f e H(X) such that G(f) is equicontinuous, Go(f) is abelian, for every f e H(X) such that G(f) is equicontinuous,
we have g(x) = -g(-x) for all g e G0(f), x e X,
(9.3. IM)
(9.3.1.vii)
for every equicontinuous fat subgroup G C H(X), we have
g(x) = -g(-x) for g e Go, x e X.
(9.3.l.viii)
F-Norms and Isometrics in F-Spaces
(9.3.1 .iii) c . (9.3.1.iv). These implications are
Proof. (9.3.1.1).
trivial. (9.3.l.iv)
393
(9.3.l.v). Under (9.3.1.iv) each gE G0(f) is linear. Obviously
G(f) = G(g) and Go(f) = Go(g). By the linearity of g, G(g) consists of
mappings of the form h(x) = Egl(x)+y, E _ ±1, n = 0, ±1, ±2, ..., y c- X. Thus (9.3.1.v) holds. (9.3.l.v) (9.3.1.vi) is obvious.
(9.3. L vi) = (9.3. L vii). Observe that the operator -I, where his an identity belongs to Go(f ). Since Go(f) is abelian g = (-I)g(-I). (9.3.1 .vii) . (9.3.1 .viii). Let G be an arbitrary fat equicontinuous subgroup of H(X). Fix g E G. Then G(g) C H(X) is also equicontinuous.
Thus, by (9.3.l.vii), g(x) = -g(-x). (9.3.1 .Viii)
(9.3.1.i). By (9.3.1 .viii),
g(2 (x+(-x))) = 0 = z (g(x)+g(-x)) for each g c- Go, x c- X. Since G is a fat subgroup, for any f e G
f (2 (x+Y)) =' (f(x)+f(y))
(9.3.2)
for all x, y c- X. Since X is a space over reals and f is continuous, by (9.3.2) f E Afi (x).
Let G be a subgroup of H(X). Let A be a subset of X. By GA we shall denote the smallest G-invariant set containing A
GA=Ug(A) pEG
The class of all G-invariant subsets will be written
Inv (G) _ {A : GA = A}. LEMMA 9.3.6 (Mankiewicz, 1979). Let X be an F*-space and let G be a fat subgroup of H(X). Then for every A E Inv(Go), x c- X, g E Go we have
g(x+A) = g(x)+A = g(x)+g(A), if A E Inv(G0), then A = -A,
(9.3.3.i) (9.3.3.ii)
Chapter 9
394
if At e Inv(G0), t e T, then n At e Inv(GO) for any set of indices T, tET
(9.3.3. iii)
if A e Inv (GO), then its closure A e Inv (GO), if Al, ..., An e Inv(G0), then A,+ ... +An Inv a (G0),
(9.3.3. iv)
(9.3.3.v)
if G is equicontinuous, then there is a basis of neighbourhoods of zero which are Go invariant. (9.3.3.vi) Proof. (9.3.3.i). Let x e X, A e Inv(GO), g e Go. Letf(u) = g(u+x)-g(x). Of course f e Go. Then f(A) = A and by the definition off
g(x+A)-g(x) =f(A) = A = g(A). (9.3.3.ii), (9.3.3.iii), (9.3.3.iv) are obvious. (9.3.3.v). Take A, B e Inv(G0). Then
g(A+B) = 2EA U g(x+B) = U g(x)+B = g(A)+B = A+B, xEA i.e. A+B e Inv(G0). Thus, by induction, we obtain (9.3.3.v),
(9.3.3.vi). Let {U.) be a basis of neighbourhoods of zero. The sets {Go Uj are Go invariant. Since Go is equicontinuous, they form a basis of neighbourhoods of zero. LEMMA 9.3.7. Let X be a stricly galbed space. Let G be an equicontinuous fat subgroup of H(X). Let A be a bounded set contained in X. Then G. A is also bounded. Proof. Let V be an arbitrary neighbourhood of zero in X. Since the space
X is strictly galbed, there is a neighbourhood of zero U such that, for each positive integer k, there is a positive number ak such that
U+... +U C akV.
(9.3.4)
k-times
Basing ourselves on (9.3.3.vi), we can assume without loss of generality that U is Go invariant. Let A be a bounded set in X. Then there is a constant k such that
ACkUC U+... +U. k-times
(9.3.5)
F-Norms and Isometries in F-Spaces
395
By (9.3.3.v) the set U+ ... +U is Goinvariant. Thus, by (9.3.4) and k-times (9.3.5).
G0(A) C G0(U+
... + U) = U+ ... + U C ak V.
k-times
(9.3.6)
k-times
The arbitrariness of V implies that G0(A) is bounded.
Problem 9.3.8. Is Lemma 9.3.7 true without the assumption that X is strictly galbed ?
LEMMA 9.3.9 (Mankiewicz, 1979). Let X be an F*-space. Let G be a fiat subgroup of H(X). Assume that A C X is a Go invariant subset of X and that x0 is an extreme point of A. Then for every g e Go we have
g(xo) _ -g(-xo).
(9.3.7)
Proof. By (9.3.3.ii) A = -A, hence -xo e A and -xo is also an extreme point. Let B = (xo+A) rte, (-xo+A). Observe that B = -B. We shall show that B = {0}. Indeed, suppose that x 0, x B. Since 0 = (x+ z +(-x)), 0 is not an extreme point of -xo+A and x0 is not an extreme point of A and we obtain a contradiction. Thus B = {0}. By (9.3.3.i)
{0} = g({0}) = g((xo+A) n (-xo+A)) T g(xo)+g(A)) n (g(-xo)+g(A)).
(9.3.8)
Since g(x0) +g(-xo) E (g(xo)+g(A)) n (g(-xo)+g(A)), g(xo)+g(-xo)
=0.
THEOREM 9.3.10 (cf. Mankiewicz, 1979). Let X be a strictly galbed real F*-space with the strong Krein-Milman property. Then X has the affine group property. Proof. Fix f e H(X) such that G(f) is equicontinuous. Let
B = {xe X: g(x) = -g(-x), for ge G0(f)}. The set B is Go(f )-invariant. Indeed, let h e G0(f) and y e B. Then
g(h(y)) = (gh)(y) _ -(gh)(-y) = -g(h(-y))
= -g(-h(y)) and h(y) a B.
Chapter 9
396
We shall show that
B = X.
(9.3.9)
Suppose that (9.3.9) does not hold. Let x 0 B. It is easy to see that the set B is closed. Hence by (9.3.3.vi) there is a G0(f)-invariant neighbourhood of zero V such that x+ V r) B = 0. Thus x 0 B+ V. By (9.3.3.v) B+ V is Golf )-invariant. Thus by (9.3.3.i)
g(x) 0 g(B+ V) = B+ V = g (B)+g(V)
(9.3.10)
for each g e Golf ). Hence the set A = {g(x) : g e Go(f )} is disjoint with the
set B+ V. The set B+ V is open, thus the closure A of the set A is also disjoint with B+ V. On the other hand, A is the smallest G0(f)-invariant set containing x. By Lemma 9.3.7 the set A is bounded. Since X has the strong Krein-Milman property, the closure A of the set A contains an extreme point x0. Then, by Lemma 9.3.9, xo e B, and we obtain a contradiction. Thus (9.3.9) holds and (9.3.l.vii) also holds. Therefore, by Proposition 9.3.5, X has the affine group property. Proof of Theorem 9.3.4. The theorem is an immediate consequence of Theorem 9.3.10.
The original versions of the Mankiewicz theorems (Theorem 9.3.4 and Theorem 9.3.10) was formulated for locally convex spaces. The following example shows that there are non-locally convex spaces with the strong Krein-Milman property. Example 9.3.11
Let 0
IIXIIm =
.
(n-11mlxnJ)P,
n=1
with the topology defined by the sequence of p-homogeneous F-pseudonorms {1I 11m} (see Example 1.3.9). It is easy to verify that LP(n-11m) are Schwartz spaces, and thus Montel spaces. This means that every closed
F-Norms and Isometries in F-Spaces
397
bounded set is compact. Moreover, the space LP(n-1l'") has a total family of linear continuous functionals. Thus, by Remark 5.5.6, every compact
set has an extreme point. Hence the spaces LP(n-1/m) have the strong Krein-Milman property. The spaces LP(n-1/m) are not locally convex. Indeed, let en = {0, ... 0, 1, 0, ...}. The sequence {en} tends to 0, because n-th place
IIeni I,n =
(n-IIm)P ->0
form= 1,2,... On the other hand, n
e1+ ... +en n
m
P
i=1
n
P -n-1/m) = n1-P-P/m ->. 00, 1
n
provided p(1+l/m) < 1. Therefore the sequence I
e1+ ... +en } is not
bounded. This implies that the spaces LP(n-11n') are not locally convex. The important class of spaces, namely LP[O, 1], 0 < p < 1, do not have the strong Krein-Milman property. For this reason we shall prove THEOREM 9.3.12 (Rolewicz, 1968). Let (X, II IIx) and (Y, II IIy) be two real locally bounded spaces. Suppose that the norms II IIx and II IIY are concave, i.e., for all x e X, y e Y, the functions IItxl Ix and IIty I IY are concave for pos-
itive t. Then every rotation mapping X onto Y is a linear operator.
Proof. Let r be a positive number such that the set K2, = {x e X: IIxii 2r} is bounded. Such an r obviously exists, since the space X is locally bounded. Using the concavity of the norm, we shall show that sup IIxiix < r.
(9.3.11)
112zIIx<_ r
Suppose that (9.3.11) does not hold. Then there is a sequence {xn} of elements of X such that IIxniix = r and an =
r
21r-I- Xn 2
.I
II
2
I
x
-+r. Therefore
Chapter 9
398
The concavity of the function ItxIIx implies that I Ian xnIJx < 2r. This leads to a contradiction, because the set K2,. is bounded.
If the set K2r is bounded, then the set K2, is also bounded for all s, xIIx. The function n (r) is continuous and
0 < s < r. Let n (r) = sup
I I2zI Ix?
it is strictly increasing, provided r < ro, where K2,, is bounded as follows from (9.3.11). Let us define by induction
n= 1,2,...
rn=n(rn-1)
Obviously, by (9.3.1), ro > r1 > r2 > ... > rn > ... We shall show that lim rn = 0.
(9.3.12)
indeed, suppose that (9.3.12) does not hold, i.e. that
r'= lim rn>0.
(9.3.13)
Since n(r) is strictly increasing, n(r') < r'. The continuity of the function n (r) implies that there is an r > r' such that n (r) < r'. By the definition of r' there is a positive integer n such that rn < r. Hence n(rn) < n(r) < r'. This leads to a contradiction, because n(rn) = rn+l > r'. Let x' and y' be two arbitrary elements of X such that IIx'-y'. < r0/2. Let Ho = fx e X : IIxx'IIx
and IIx-y'Ix
Obviously
S (Ho) = sup IIx-YIIx < 2n (2) < ro. x,yEHo
We define by induction H. = {x e H._1: IIx-YIIx < rn for all y e Htt_1} (n = 1, 2, ...). We shall prove by induction that the sets H. are not void and are such that
x'+Y' e Hn, 2
n = 0, 1, 2, ...,
if xeH,,,then x=x'+y'-xeH,,,n=0, 1,2,..., S(Hn)
n = 0, 2, 2, ...
(9.4.14.i) (9.3.14.ii) (9.3.14.iii)
F-Norms and Isometries in F-Spaces
399
For n = 0 this is trivial, since x-x' = y'-x and x-y' = x'-x. Suppose that the above holds for a certain k-1. Then (9.3.14.iii) implies that (9.3.14.i) holds for n = k. Hence the get Hk is not empty. By the definition of Hk
8(Hk)= sup IIx-YIIx< sup IIx-YIIx
xEHk
yEHk -1
i.e. (9.3.14.iii) holds for n = k. Let x e Hk and y E Hk_1. Then, by definition of Hk, (9.3.15) 11x-YIIx = IIY-xllx
the point (x'+y')/2 in the metric language, this implies that there is a number a > 0 such that if IIx-yMI,< < a then, for each isometry U,
U( x+Y 1 2 /
=
U(x)+ U(Y) 2
This implies that if jxllx < a/2 then
2U(kx) = U((k+1)x)+U((k-1)x).
(9.3.16)
Basing ourselves on formula (9.3.16), we shall show by induction that
U(nx) = nU(x),
n = 1, 2, ...
(9.3.17)
Putting k = 1 in formula (9.3.16), we find that (9.3.17) holds for n = 2.
Let us suppose that (9.3.17) holds for n = m and let us put k = m in (9.3.10). Then the induction hypothesis implies
2mU(x) = 2U(mx) = U((m + 1) x) + U((m - 1) x)
= U((m+1)x)+(m-1)U(x). Hence
U((m+1)x) = (m+1) U(x), and we have proved (9.3.17).
Chapter 9
400
Let x and y be arbitrary elements of X. Obviously, there is a positive integer n such that IIx/nllx < a/2 and IIy/njJx < a/2. Therefore
U(x+Y) = U( 2
2n(x+y) 2n
)
= 2nU
x-{-y 2n )
Hn)+U(n ))
U(x)+U(Y)
Hence the operator Uis additive. Since it is continuous and the spaces are real, it is a continuous linear operator.
We do not know whether Theorem 9.3.12 is true for arbitrary norms in a locally bounded space. We can only prove THEOREM 9.3.13. Let X and Y be two locally bounded real spaces. Let U be
a rotation such that II tU(x)=t U(y)IIy = II tx-tyllx for all positive t. Then the operator U is linear. Proof. Since no confusion will result, we shall denote the two norms II IIx and II IIy by the same symbol II J. Let IIxII* = sup IItxII and IIxII** o
sup 6>1>a>
(b-
a IIbxI**
(cf. Lemma 9.1.2). IIxII** is a norm
equivalent to the original one and the function IItxII** is concave for positive t. It is easy to verify that II U(x)- U(y)I I ** = 11x-YII** Therefore by Theo-
rem 9.3.12, the operator U is linear.
9.4. ISOMETRICAL EMBEDDINGS INBANACH SPACES Let (X, II IIx), (Y, II IIy) be two real Banach spaces. Let U be an isometry mapping X into Y. If U is not a surjection (i.e. U does not map X onto Y), then U need not be linear, as can be seen from the following
Example 9.4.1
Let X be the space of real numbers R with norm IIxII = IxI and let Y be a two-dimensional real space RxR with norm II(x,Y)II = max(IxI, IYI).
F-Norms and Isometrics in F-Spaces
401
Let U be a mapping X into Y given by formula U(x) _ (x, sin x). It is easy to verify that Uis a non-linear isometry. However, for Banach spaces the following substitute of Theorem 9.3.12 holds : THEOREM 9.4.2 (Figiel, 1968). Let X and Y be two real Banach spaces. Let U be an isometry (non-necessarily linear) mapping X into Y and such that U(O) = 0. Let the linear hull of the set U(X) be dense in the space Y. Then there is a continuous linear operator F mapping Y into X and such that the
superposition FU is an identity on X. The operator F is uniquely determined and it has norm one.
The proof of Theorem 9.4.2 is based on the following notions and lemmas. Let X be a Banach space with a homogeneous norm I IxI I. Let
Sr = {x: IIxil = r}. We say that a point x e Sr is a smooth point if there is only one continuous linear functional f-- of norm one such thatfx(x) = r. PROPOSITION 9.4.3 (Mazur ; see Phelps, 1966). If (X, II II) is a separable Banach space, then the set Ss of all smooth points is a dense G,5-set in X. Proof. In the case of a complex Banach space X, we consider it as a linear
space over reals. The set of all smooth points S's remains unchanged. Therefore we can restrict ourselves to the real Banach spaces. As can easily, be verified, it is sufficient to show that Ss n Sl = S,, is a dense G. set in S. Let {xn} be a dense subset of S. For each positive integer m, let D,n,n _ {xc S1: IIf(xn)-g(xn)II < 1/m for all continuous linear functionals f ,g such that IIfil = f(x) = 1 = g(x) = IIgIJ}. Of course, if x is a smooth point, then there is only one functional f such that IIfJI fore,
IIf(xn)-g(xn)II = 0 < m,
n, m
= f(x) =
1. There-
Chapter 9
402
This implies that oo
Ss c n Dm,n.
(9.4.1)
m,n =1
On the other hand, if x is not a smooth point, then there are at least two different continuous linear functionals f, g of norm one such that f(x) = g(x) = 1. Since the set {xn} is dense in S1,
supIf(xn)-g(xn)I = Ilf-gI n Hence, there are such m and n that x
Dn,n. This implies
00
is ) nn,mDm,n. =1
(9.4.2)
(9.4.1) and (9.4.2) imply M
Ss = n Dm,n. n,m =1
We shall show now that for every n, m the set Dm,n is open. Let yk e Sl\Dm,n and yk --may. Since yk e S\Dm,n, there are functionals fk and gk such that IIfkuI = I fk(yk)I = 1 = Igk(yk)I =
IIgkII,
k = 1, 2,...
(9.4.3)
and
I fk(xn)-gk(xn)I > 1/m,
k = 1, 2, ...
(9.4.4)
By the Alaoglu theorem (Theorem 5.2.4) the sequence { fk} and {gk} have cluster points f and g respectively. By (9.4.3)
IIfll=If(Y)I=1=Ig(Y)I=IlgHI, and by (9.4.4)
I f(Y)-g(Y)I > 1/m. This implies that y 0 Dm,n. Then the set Dm,n is open. For the completion of the proof it is enough to show that for each n and m the set Dm,n is dense in S1. Suppose that the above does not hold. Then
there are such n, m, y e Sl and 6 > 0 that, if x e S1 and IIx-YII < S, then
F-Norms and Isometries in F-Spaces
403
x 0 Dm,n. Let us write y1 = y. Then there are continuous linear functionals f1 and g1 that
and
ii =f](Y1) = 1 = g1(Y1)
.fi(x) > g1(xn)+1/m. Now we shall choose by induction a sequence {yk} C S1 and sequences of functionals { fk} and {gn} of norm one such that IIY1-YklI < (1-2-k)6,
(9.4.5.i)
fk(Yk) = 1 = gk(Yk),
(9.4.5.ii)
fk(xn) > m +g1(xn) .
(9.4.5.iii)
For k = 1 conditions (9.4.5.i)-(9.4.5.iii) are satisfied. Suppose that they hold for a certain k. Then Yk+1
Yk+axn IIYk+axnll '
where a is chosen so small that (9.4.5.i) holds. Then yk+l 0 Dm,n and there are continuous linear functionals fk+1, gk+j such that IIfk+1II = Ifk+1(Yk+1)I = 1 = Igk+1(Yk+l)I = IIgk+1II
(9.4.6)
and fk+l(xn) >
m +gk+1(xn)
We have 1=IIYk+11I > fk(yk+1)
_ =
1 +afk(xn)
(9.4.7)
IIYk+axnll
Since
gk+1(Y)k+1 = 1 > gk+1(yk),
(9.4.8)
we have IIYk+axnll = gk+] (Yk+axn)
1 +agk+1(xn)
(9.4.9)
By (9.4.7) and (9.4.9) fk(xn) -< gk+1(Xn).
(9.4.10)
Chapter 9
404
Then, by the induction hypothesis, fk+1(xn) >
YYt
r +9k+ 1(X.) > In +fk(X.)
k+1 m
+gj(xn).
(9.4.11)
Hence (9.4.5.iii) holds. Let us observe that this leads to a contradiction, because Ifk+i(xn)I < IIfk+]II IIXnII = I. LEMMA 9.4.4. Let U be an isometry of the space of reals R with the standard norm IxI into a Banach space (Y, II II). Let U(O) = 0. Then there is a continuous linear functional f e Y of norm one such that
f(U(x)) = X. Proof. Let n be an arbitrary positive integer. The Hahn-Banach theorem implies that there is a continuous linear functional fn of norm one such that
fn(U(n)-U(-n)) = II U(n)-U(-n)II = 2n. Thus, for every t, I t I < n, we have
2n = In-tI+It-(-n)I
= II U(n)- U(t)II+II U(t)- U(-n)II
fn(U(n)- U(t))+fn(U(t)- U(-n))
=fn(U(n)-U(-n)) = 2n.
(9.4.12)
Therefore, in formula (9.4.12) the equality holds, and this implies
fn(U(t)-U(-n)) = II U(t)-U(-n)II = t+n.
(9.4.13)
Putting t = 0 in (9.4.13) we obtain fn(- U(-n)) = n. Thus
fn(U(t)) = t.
(9.4.14)
The Alaoglu theorem (Theorem 5.2.4) implies that the sequence f f.} has a cluster point f. Formula (9.4.14) implies thatf(U(t)) = t. LEMMA 9.4.5. Let x be a point of the Banach space X. Let a be a smooth point of the set Siiaii = {x: IIxiI = IIaII}. Let fa be a functional of norm one such that fa(a) = IIaII. Let fa(x) 0. Then there is a real t such that Ila+txll < IIaII.
(9.4.15)
F-Norms and Isometries in F-Spaces
405
Proof. Suppose that (9.4.15) does not hold, i.e., for all real t, (9.4.16)
IIa+txll > Hall.
Since fa(x) # 0, x 0. Formula (9.4.16) implies that a and x are linearly independent. Let Xo denote the space spanned by a and x. The formula g(ax+fia) = ,9Hall
defines a continuous linear functional on X0. Formula (9.4.16) implies that IgII = 1. Since g(a) = Ilallg is a restriction of the functional f. into Xo, fa(x) = g(x) = 0 and we obtain a contradiction. LEMMA 9.4.6. Let U be an isometry of a Banach space X into a Banach space Y such that U(0) = 0. Let a be a smooth point of the sphere Siiaii. Let f e Y* be a continuous linear functional of norm one such that, for all real r,
f(U(ra)) = rllall
(9.4.17)
f(U(x)) = Mx)-
(9.4.18)
Then
Proof. Let x, y e X. We have
If(U(x))-f(U(Y))I = I.f(U(x)- U(y)) I II U(x)- U(y)II = Ily-xll
(9.4.19)
Suppose that for a certain p e X
fa(p) #f(U(p))
(9.4.20)
Let us write
a =f(U(p))
llail
Then (9.4.20) implies
fa(p-aa) # 0.
(9.4.21)
By Lemma 9.4.5 there is a real t such that
llaa+t(p-a)II < llaall It is clear that t 0. Let us put 9 = alt. Then (9.4.22) implies
llp-(a-i)all < llflall
(9.4.22)
(9.4.23)
Chapter 9
406
By (9.4.17), (9.4.19) and (9.4.23) we obtain
= If(U(P))-f(U((a-f3)a))I
IIRall =
jaIIIaII-(a-fl)IIaII
which leads to a contradiction. LEMMA 9.4.7. Let X be a finite-dimensional real Banach space. Let U be an isometric embedding of the space X into a Banach space Y such that U(O)
= 0. Then there is a continuous linear operator F mapping the linear hull of the set U(X), lin U(X), onto X and such that
F(U(x)) = x. The operator F is uniquely determined and it is of norm one.
Proof. Let us denote the dimensional of X by n. Let al, ..., an be linearly independent smooth points. Such points exist, as follows from Proposition 9.4.3. Let fa, be a continuous linear functional of norm one such that fas(ai) = IIaill,
i = 1, 2, ..., n.
Without loss of generality we may assume that ai are chosen in such a way
that the functionals {ft}, i = 1, 2, ..., n, are linearly independent. Let 01 be continuous linear functionals of norm one such that
fi(U(rai)) = r l aill,
i = 1, 2, ..., n
for all real r. The existence of such functionals follows from Lemma 9.4.4.
Lemma 9.4.6 implies that
fi(U(x)) =fai(x),
i = 1, 2, ..., n.
(9.4.24)
The mapping G of the space X into Rn given by the formula
G(x) = {fa,(x), ...,fa.,(x)}
(9.4.25)
is an isomorphism, because the linear functionals { fal, ... , fag} are linearly independent. Now we define an operator F mapping lin U(X) into X as follows :
F(y) = G-1({fi(y),... , fa(y)}) .
F-Norms and Isometries in F-Spaces
407
It is obvious that Fis a continuous linear operator. For x e Xwe have by (9.4.24) and (9.4.25)
F(U(x)) = G-1({f1(U(x)), ...,fn(U(x))}) = G-1(jfa,(x), ..., fa (x)}) = x.
(9.4.26)
Formula (9.4.26) implies that the operator F is uniquely determined on U(X) and thus on lin U(X). We shall now show that IIFII = 1. Let x e Ybe such a point that F(x) is a smooth point. By Lemma 9.4.4 there is a continuous linear functional f of norm one such that for all real r f(U(rF(x))) = rMMF(x)II.
(9.4.27)
By Lemma 9.4.6 f(U(y)) = fF(x)(y)
(9.4.28)
Therefore, the superpositions f(U(.)) and
are continuous
linear operator Moreover, we have by (9.4.26)
f(U(F(U(y)))) =f(U(y))
for y e X.
Hence, for x = U(y) such that F(x) is a smooth point, (9.4.29)
IIxII > If(x)I = If(U(F(x)))I = IIF(x)II.
Since F is a continuous linear operator, the set of all x such that F(x) is
a smooth point of dense in X (cf. Proposition 9.4.3). Therefore, by (9.4.29), IIFII < 1. On the other hand, by (9.4.26), IIFII > 1. Thus IIFII
= 1. Proof of Theorem 9.4.2. Let {Xn} be a sequence of finite-dimensional 00
spaces such that dim Xn = n, X. C
and the set Z = U Xn is dense n=1
in X. By Lemma 9.4.7 there are continuous linear operators Fn, IIFnlI = 1 mapping fin U(XX) into X and such that
Fn(U(x)) = x
for xe Xn, n = 1,2,...
In view of the uniqueness of the operators, F., the operator
F(y) = Fn(y)
for y e lin U(XX)
Chapter 9
408
is a continuous linear operator of norm one, well defined on lin U(Z). The extension of the operator F to the closure lin U(Z) = lin U(X) has the required properties. 9.5. GROUP OF ISOMETRIES IN FINITE-DIMENSIONAL SPACES Let (X, II II) be a finite-dimensional real F-space. Let G(II II) denote the set
of rotations mapping X into itself. By Theorem 9.3.4 all those rotations are linear. It is easy to verify that G(II II) is a group with the superposition of operators as the group operation. THEOREM 9.5.1 (Auerbach, 1933-1935). Let (X, II II) be an n-dimensional
real F-space. Then there is an inner product'(x,y) defined on X such that G(II II) C G(II
II1), where IIxII1 = (x,x)112 is the norm induced by the inner
product (x,y). Proof. Let K = {x: IIxII < I). Of course, for any isometry, U c- G(JI 1j), U(K) = K. Let E denote an ellipsoid with the smallest volume containing K. We shall show that this ellipsoid is uniquely determined. Indeed, let E, be another ellipsoid with the smallest volume containing K. Let E = {x : (x, x) < 1} and E1 = {x: (x, x)1 < 1}, where the inner products (x,y) and (x,y)1 are determined by E and E. Let a be an arbitrary real number contained between 0 and 1. Let
Ea={x: (x,x)a<1}, where (x, y)a = a(x, y)+(1-a) (x,y)1. Of course Ea is an ellipsoid. More-
over, K is contained in Ea. Indeed, since KC En E1 if x e K, then (x, x) < 1 and (x, x)1 < 1. Therefore
(x,x)a = a(x,x)+(1-a) (x, x), < 1
and x e Ea. Since the ratio of the volumes is an invariant of affine transformations, we may assume without loss of generality that E is a ball in the usual Euclidean sense and E1 is an ellipsoid with the equation
b1xi+... + b,,x,2,<1,
b;>0,
i=1,...,n.
Since vol (E) = vol (E1), b1
.. bn = 1.
(9.5.1)
F-Norms and Isometries in F-Spaces
409
The volume of the ellipsoid E. is expressed by the formula
vol (Ea) = Cn([a+(1-a)b1] ... [a+(1-a)bn])-1/2, where C. is a constant dependent on n. Let us observe that the function vol(Ea) reaches its minimum at those points at which the function
V(a) = (a+(1-a) b1) ... (a+(1-a)ba) reaches its maximum. The function V(a) is differentiable, and by a simple calculation we get
d
da
V(
= V(17)
i=1
1-bi a+(1-a)b{
We have assumed that E and E1 are ellipsoids with the minimal volume. d Therefore, the derivative da V(a) should be equal to 0 for a = 0 and a = 1. The second case implies
b1+ ... +bn = n. (9.5.2) Equations (9.5.1) and (9.5.2) have a unique solution b1 = b2 _ ... = bn = 1 within positive numbers. This follows from the fact that the function W(b) = b1 ... bn-(b,+ ... +bn) has only one extremal point b1 = ... = bn = 1 and it is positive for other positive b. Therefore the ellipsoid E is uniquely determined by K. Thus any isometry U e 0 maps E onto itself.
9.6. SPACES WITH TRANSITIVE AND ALMOST TRANSITIVE NORMS Let (X, II II) be an F-space. By G(I II) we denote the group of rotations
mapping X onto itself. A norm II II is called maximal if, for any equivalent norm II II1 such that G(II II) C G(II II1), we have equality G(II
II) = G(II II).
PROPOSITION 9.6.1. In finite-dimensional real F-spaces the maximal norms are induced by inner products.
Proof. The proposition is an immediate consequence of a theorem of Auerbach (Theorem 9.5.1).
Chapter 9
410
Let (X, II I) be an F-space. The norm IxII is called transitive if, for all
positive r and each x e X, IIxII = r,
{A(x) : A e G} _ {y: IIYII = r},
(9.6.1)
where G(II II) denotes the group of rotations with respect to the norm IIxII.
A norm IIxII is called almost transitive, if for each positive r and for all x e X of norm r, {A(x) : A e G(II ID} = {y : IIYII = r}.
(9.6.2)
If the space X is locally bounded and the norm IIxII is p-homogeneous,
then the conditions of transitivity and almost transitivity are simpler, namely, it is sufficient that (9.6.1) (respectively (9.6.2)) should hold for one
fixed r, for example r = 1. Strictly connected with transitive norms is the classical problem of Banach. Problem 9.6.2 (Banach, 1932). Let (X, II D be a separable Banach space with a transitive homogeneous norm IIxII. Is X a Hilbert space and is IIxII
a Hilbert norm? As we shall show later, without the assumption of separability the answer is negative. THEOREM 9.6.3 (Pelczynski and Rolewicz, 1962). Let (X, II II) be an Fspace and let IIxII be an almost transitive norm. Then the norm IIxII is maximal.
Proof. Let IIxII, be a norm equivalent to the norm IIxII and such that G(II
II) C G(II II1). Let ,
IIxII, = f IItxIIidt. 0
The norm IIxII, is equivalent to the norm IxII, and such that the function IItxii, is strictly increasing for positive t (cf. Theorem 9.1.1). It is easy to verify that a linear isometry U with respect to the norm IIxII, is also an isometry with respect to the norm IIxII, Hence, G(II II1) C G(II II,). Since
F-Norms and Isometries in F-Spaces
the norm IIxii is almost transitive for any fixed xo such that 11x011 IIxollI = r1, we have
411
= r and
S(r) = {x : Ixii = r} = {A(xo) : A e G(JI II)} C {A (xo) : A e G(JI I1'1)} C {y: IIyII1= rl} = S'(rl).
(9.6.3)
Since the function 11tx1I1 is growing for positive t, (9.6.3) implies that
S(r) = S'(rl).
(9.6.4)
Hence the isometries belonging to G(JI II) map S(r) onto itself. This implies that G(JI 111) C G(JI 11), i.e. G(JI II) is a maximal group of isometries.
THEOREM 9.6.3 (Pelczynski and Rolewicz, 1962). In the spaces LP[O, 1],
1 < p < oo, the standard norm i 11x11=
(1Ix(t)IPdt)1IP
0
is almost transitive.
Proof. Letf(t) e LP[O,1] be a function of norm one such that
a = inf f(t) > 0.
(9.6.5)
o
Let us consider the operator t
Tf(x) = x(F(t))f(t),
where F(t) = f I f(t)jPdt. 0
The operator Tf is an isometry acting in the space LP[O, 1]. Indeed, IITf(x)IIP = f I x(F(t))f(t)I Pdt 0
= f I x(F(t))I PdF(t)
= IIxllP
(9.6.6)
0
because F(t) is a strictly increasing function such that F(O) = 0 and F(1) = 1. Let us observe that by (9.6.5) the inverse isometry
Tf'(y) =
y(F-1(t))
I f(F-1(t))
(9.6.7)
Chapter 9
412
is well defined on the whole LP[O,1]. Therefore, the isometry Tf maps LP[O, 1] onto itself.
Let f and g be two arbitrary elements of norm one. Let a be an arbitrary
positive number. Obviously, there are two functions f and g' such that inf f'(t) > e/4, inf g'(t) > e/4, II f-f'II < e/2, Ilg-g'11 < E/2. o
o
Let U = Tp,Tj 1. The operator U is an isometry.
Since Tf,(l)
= f', T,,(1) = g', we get Uf ' = g'. Hence
IIU(f)-giJ < IIU(f)-U(f')II+IIU(f')-g')II+IIg'-gII < E. The arbitrariness of e implies that the norm x is almost transitive.
Another example of a separable Banach space (X, II II) such that the norm II II is almost transitive are Gurarij spaces.
Gurarij (1966) gave an example of a Banach space (X, II II) with the following extension property. For each finite-dimensional subspaces E, F,
E C F C X and for an operator T: E-*X and each e > 0, there is an extension TE of T, T.: F-->X such that (1-e)IIxII < IITE(x)II < (l+e)IIxII
Lusky (1976) showed that all Gurarij spaces are isometric and that the norms in a Gurarij space is almost transitive. Later Lusky (1979) showed that every Banach space (X, II ID is a subspace of a space (Y, II ID (its norm being an extension of II ID such that in Y the norm II II is almost transitive and there is a continuous projection of Y onto X. THEOREM 9.6.4 (Pelczyriski and Rolewicz, 1962). In the spaces LP[0,1],
0
IXI = f Ix(t)IPdt 0
is almost transitive.
The proof follows the same lines as the proof of Theorem 9.6.3; only in formula (9.6.6) it is necessary to replace IITf(x)IIP by IITf(x)II and IIxiiP by IIxII.
F-Norms and Isometries in F-Spaces
413
COROLLARY 9.6.5 (Pelczynski and Rolewicz, 1962). In the spaces LP[O,1],
0 < p < +oo, the standard norms are maximal. PROPOSITION 9.6.6 (Pelczynski and Rolewicz, 1962). There is a normed (non-complete) separable space with a transitive norm which is not a preHilbert space.
Proof. Let LP, 1
maps the sets Ea = {t: jx(t)j > a} onto the intervals [0, JEaJ) Let
Uxf =f(hx(t)) Obviously, Ux is an isometry and maps the function x(t) on a function x'(t) such that x'(t) is a non-increasing function and
>0 lx '(01
{
=0
for 0 < t < ax, for ax
where ax = J{t: x(t) 0}1. Since x e LP[O, 1], ax < 1. In the same way we construct an operator Uy mapping y on y' with the same properties as x'. Let us observe that the operator T71 defined by formula (9.6.7) is well defined also if we replace condition (9.6.6) by the condition f'(t) =,4 0. Indeed, the operator Tf 1 is well defined on the set Xa :
Xo = {x a LP[0,1] : there is a positive a such thatf(t) < a
implies x(t) = 0}. Since the set X0 is dense in LP[0,1], the isometry T71 is well defined on the whole LP[0,1]. Therefore, using the method described in the proof of Theorem 9.6.3 we can define the isometries Uz, and Uy mapping x' and y' on the func-
Chapter 9
414
tions x" and y", where allp x"(t) = lox
for
Iallp
for
0 < t < ay,
for
ay < t < 1.
Y"(t) _
Y
to
0 < t < ax,
for ax
Let U be an operator defined by the formula
U(f)
= (ay )i/ f \ ay tl
for 0 < t < ay,
(i'::)il f(-(1-ax) (1-t)+1)
for ay < t < 1.
It is easy to verify that the operator Uis an isometry and that U(x") = y". Hence the isometry A = Uy 1Uy,1UUx,Ux maps the element x on y. PROPOSITION 9.6.7 (Pelczyriski and Rolewicz, 1962). There is a non-separable Banach space with a transitive norm which is not a Hilbert space.
Proof. Let Q be the product of a non-countable set A by the closed interval [0, 1], S2 = A x [0, 1]. Let E be the algebra of all such subsets E of S2
that for each a e A the set E. = En ({a x [0,1]) is measurable in the Lebesgue sense. Let a measure p be defined by the formula
p(E) = ' I(Ea)I, aEd
is the Lebesgue measure. If x = x(a, t), a E A, t E [0,1] is an element of LP(92, E, p), then the support Sx = {a c- A: J (t: x(a, t) 0}I} of the element x is at most countable. We shall show that the standard norm in LP(Q, E, p) is transitive. Let x = x(a, t), y = y(a, t) be arbitrary elements of LP(12, E, p) of norm one. Let {an} be a sequence containing the supports of x and y. Let {A,,,} be as sequence disjoint with {an}. Let us take a sequence {fl.} such that N2n-1 = an and ,82n = An. Let U be an operator defined in the following where I
I
way :
g(/1,t)= U(f(a,t))),
g(fn,t) =
where g(a,t) = t (or, t) for ix
1211Pf(f'2n-1,2t)
for 0 < t <1/2,
21/Pf(P2n,2t)
for 1/2 < t < 1.
F-Norms and Isometries in F-Spaces
415
Let us observe that the operator U is an isometry and that it maps x on x' and y on y', so that x'(a, t), y'(a, t) belong, for any fixed a, to the space LP[0,1] defined in the preceding proposition. Thus, by Proposition
9.6.6, there is an isometry V mapping x' on y'. Hence, the isometry U-'VU maps x on y. Hence the standard norm in the space LP(Q, is transitive. If p 2, the space LP(Q, E, p) is not isomorphic to a Hilbert space.
9.7. CONVEX TRANSITIVE NORMS
In this section we shall consider only Banach spaces with homogeneous norms. Let X be a Banach space with a homogeneous norm 11x1 j. The norm IIxII is called convex transitive if, for each element x of norm one,
cony {A(x) : A E G(II ID} = (x: IIxII < THEOREM 9.7.1 (Pelezyriski and Rolewicz, 1962). Let X be a Banach space. Each convex transitive norm IIxII equivalent to the original one is maximal. Proof. Let IIxiii be a norm equivalent to IIxII Let G(II II) C G(II Il). Then IIA(x)II = IIxII
for all
A e G(II
II1).
(9.7.2)
Suppose that (9.7.2) does not hold. Then there are Ao e G(II Ilt) and x e X, Ilxlll = 1 such that IIAo(x)II # IIxII Without loss of generality we can assume that IIAo(x)II > IxII Indeed, otherwise we might consider the operator Ao 1. Let
r=
IIAo(x)II IIxII
Since G(II II) C G(II IIl) and the norm IIxII is convex transitive,
{y: Ilylli < 1} = cony {A(x): Ae G(II II1)} = conv {A(Ao(x)) : A e G(II Iit)} j conv {A(Ao(x)) : A E G(II ID} = {y: IIYII = rllxll}. Repeating this argument for the element rx we can prove that I1r2xlll < 1 and by induction llr"xll1 < I for n = 1, 2, ...
Chapter 9
416
Since r > 1, this leads to a contradiction. Therefore (9.7.2) holds and this implies that A e G(II II), i.e. G(II ID=G(II III). PROPOSITION 9.7.2 (Cowie, 1981). Let (X, II II) be a Banach space. Suppose that II II is not convex transitive. Then there is a norm II IIi equivalent to the norm II I and such that (X, II ID and (X, II IIi) are not isometric and G(II
II) C G(II D.
Proof. Take any x0 such that IIx0MI = 2. The set
U = cony ({x: IIxII < 1} u {Ax°: A e G(II ID}) is open convex and invariant under the group G(II ID. Let II III be the norm induced by the set U. Since Uis G(II ID invariant, G(II ID C G(II II). On the other hand Uis not a ball of any radius in the norm II J. Hence (X, II ID and (X, II IIi) are not isometric.
Let X = C(Q\l°) (see Example 1.3.4) be the space of all continuous function defined on a compact set Q and vanishing on a compact subset 0° with the standard norm IIxII = sup Ix(t)I
(9.7.3)
test
The set S = Sl\Q° is locally compact. Thus the spaceC(Q\Q°) can be considered as the space of continuous functions defined on a locally compact set S vanishing at infinity with the standard norm (9.7.3). For brevity, we shall denote this space C°(S).
In the space C°(S) each rotation U of the space onto itself is of the form
U(f) = a(t)f(b(t)),
(9.7.4)
where a(t) is a continuous function of modulus one and b(t) is a homeomorphism of S onto itself. Indeed, the conjugate space to the space C°(S) is the space M(S) of measures defined on Borel sets (see Example 4.3.3). The isometry U induces the isometry U* in the space M(S). Of course, U* maps the extreme points of the unit ball on extreme points. The extreme
points of the unit ball in M(S) are measures concentrated at one point multiplied by scalars of modulus one.
F-Norms and Isometries in F-Spaces
417
This implies that U is of the form (9.7.4). The continuity of a(t) follows from the fact that U maps continuous functions on continuous functions. The same fact implies the continuity of b(t). Since U maps C°(S) onto C°(S), b(t) is a homeomorphism. THEOREM 9.7.3 (Wood, 1981). Let C°(S) be the space of complex-valued functions defined on a locally compact set S vanishing at infinity with the standard norm IIxII = sup Ix(t)1 £ES
The norm I 11 is convex transitive if and only if the group of homeomorphism I
F of the set S is almost transitive on S, i.e., that for each t e S, rt = {b(t): b e I'} D S. Proof. Necessity. Suppose that T is not almost transitive on S. Then there is an s° e S such that I's° does not contain S. Let U C S be an open set disjoint with Fs°. Take any f Of e C°(S) such that the support of
f is contained in U. Thus f(Fs°) = 0. Hence, by the form of isometry (9.7.4), for each g e {Af: A e G(I )}, g(s°) = 0 and by definition the norm II II is not convex transitive. Sufficiency. Take any norm 11 111 such that G(II I) C G(II II1). We shall
show that there is a c > 0 such that Ix1I = cjIxlI1 Indeed, let II' denote the norm conjugate to the norm I I1 defined on M(S). Let 83 denote the unit measures concentrated at the point s. Since a homeomorphism b(t) induces isometrics in (C°(S), II 111), it also induces isometrics in the space (M(S), I1 II') Therefore I
Ilbsll' = I1bb(s)11'
for all s e S and all homeomorphisms b c- F. Fix s e S. Take any t E S. Since the orbit Ts is dense in S, SL belongs to the closure in the weak-*-topology of {SL : t e I'S} Hence 116t1l' < sup {I16b(3)II' : b e I'} = Iios11'.
(9.7.5)
Exchanging the roles of t and s, we find that 115311' are all equal. Let c (IISeI')-1 Then
11x111 = sup {I/"(X)I : p E M(S), hull' < 1}
< c-1 sup {Ix(t)1 : t e S} = 1 IIxIl
(9.7.6)
Chapter 9
418
Now we shall show the converse inequality. Let p be a measure of finite
support n
p = 11 ai8t
ti
for i
tj
j.
i=1
Take g(s) e C°(S) such that IIgII = 1 and g(ti) = ai/lail. By (9.7.6) IIgl11 < 1/c. Thus n
Ilpll' '> cp(g) = cf lail. i=1
On the other hand, n
n
Ilpll' <
jail Ilbt,jl'
ti=1
Therefore, for measures of finite support, (9.7.7)
I1PI1' = cllpll,
where II II denotes the norm in the space M(S) induced by the standard
norm. Since the measures of finite support are dense in the weak-*-topology in the space M(S), (9.7.7) implies IIpII' '> cllpll
(9.7.8)
llxll < cllxlll,
(9.7.9)
Hence
and finally Ixii =
cllxll1.
Thus, by the Cowie proposition (Proposition
9.7.2) the norm II II is convex transitive. COROLLARY 9.7.4 (Wood, 1981). The standard norm (9.7.3) is maximal in
a space C°(S), provided the group of homeomorphisms of S onto itself is almost transitive. Example 9.7.5 (Pelczytiski and Rolewicz, 1962)
Let Q = [0,1]. Let 92o = {0} u {1}. The standard norm (9.7.3) is convex transitive in the space QQ\S20).
F-Norms and Isometries in F-Spaces
419
Example 9.7.6 (Pelczyliski and Rolewicz, 1962) Let Q be a unit circle and let 90 be empty. Then the standard norm is convex transitive.
It may happen that in certain spaces C(Q\Q0) the standard norm is not convex transitive, and yet it is maximal. Kalton and Wood (1976) gave conditions ensuring that the standard norm is maximal in the space C°(S). There are two such conditions, namely the set S contains a dense subset such that each point of the subset has a neighbourhood isomorphic to an open set in an Euclidean space,
(9.7.10.i)
the set S is infinite and has a dense set of isolated points (9.7.10.ii)
If either (9.7.10.i) or (9.7.10.ii) holds, then the standard norm in the space C°(S) is maximal. In particular, the interval [0,1] satisfies (9.7.10.i) and by the result of
Kalton and Wood (1976) the standard norm is maximal in the space C,[0,1]. This is an answer to the question formulated by Pelczynski and Rolewicz (1962).
At present the only known example of a space C°(S) of continuous complex valued functions vanishing at infinity in which the standard norm is not maximal are spaces C°(S) where S has a finite number of isolated points t1, ..., tn. In those spaces the norm Ix(td)I2)1,2
Ix111= sup Ix(t)1+(V tES
t#ti
i=1
obviously has a biger group of isometries than the standard norm. There are also spaces which do not satisfy conditions (9.7.10.i) and (9.7.10.ii) and yet their standard norm is maximal. Let D be a closed unit circle on a two-dimensional Euclidean plane. Let {Sn} be a dense sequence in D. Remove form D by induction the interior of an n-blade propellor centred at sn and the missing boundaries of all the previously removed propellors. The remaining set E is a compact, connected and locally connected metric space. It clearly has no non-trivial
Chapter 9
420
homeomorphism since, for each n, the neighbourhoods of s are unique to that point, and so any homeomorphism must map sn on sn. Obviously, E does not satisfy either (9.7.10.1) or (9.7.10.ii). However, Wood (1981) showed that the standard norm is maximal in C,(E). Now we shall pass to investigations of spaces of real valued continuous functions. THEOREM 9.7.7 (Wood, 1981). Let S be locally compact. Let C°,(S) denote the space of all continuous real-valued functions vanishing at infinity. The
standard norm is convex transitive if and only if S is totally disconnected and the group of homeomorphisms of S is almost transitive.
Proof. Necessity. Suppose that S is not totally disconnected. Then there are s, and s2i si - s2 belonging to the same component. By the form of isometry (9.7.4) the function equal to one on that component can only be transformed into a function equal either to + 1 or to -1 constant on that component. Thus the standard norm is not convex transitive. The proof of necessity of almost transitivity of the group of homeomorphisms is precisely the same as the proof of the necessity in the proof of Theorem 9.7.4. Sufficiency. Since S is totally disconnected, for each finite system of points {ti, ..., to}, tti tj and each system of numbers {ai, ..., an}, ai _ _ 1, there is a continuous function g(t) such that jg(t) I < 1 and
g(ti)=at,
i= 1,2,...,n.
The rest of the proof follows the same line as the proof of sufficiency in Theorem 9.7.3. Example 9.7.8 (Pelczyriski and Rolewicz, 1962) Let E be the Cantor set. The standard norm (9.7.3) is convex transitive in the space C,(E) of real continuous functions defined on E.
Kalton and Wood (1976) proved that, if S is a connected manifold without boundary of dimension greater than one, then the standard norm is maximal. Of course, by Theorem 9.7.7, the standard norm is not convex transitive. It is not clear what the situation in the case of manifolds
F-Norms and Isometrics in F-Spaces
421
with boundaries and of manifolds of dimension one is like. For example,
it is not known whether the standard norm is maximal in the space of real valued continuous functions defined on the interval [0, 1], Cr[0,1]. There are also spaces S such that the standard norm is maximal in the space Ce(S) but it is not maximal in the space C.(S). Indeed, let E be the
compact space, described above, with trivial homeomorphism only. Wood has shown that in Cr(E) the standard norm is maximal. By the form of the rotations in C°(S), the unique isometries in Cr(E) are I and -I. On the other hand we have PROPOSITION 9.7.9 (Wood, 1981). Let (X, II II) be a real Banach space with
dimension greater than 1. Then there is a norm II
II,
in X such that the
group G(II IIl) contains isometries different from land -I.
Proof Take any x° E X and a linear continuous funclional f such that IIxoII = I = I.III and f(xo) = 1. Let The a symmetry
Tx = x-2f(x)x0. Of course, T' = I. T. I, -land it is easy to verify that T is an isometry with respect to the norm IIxII1 = max {IIxII, ITxHI}.
9.8. THE MAXIMALITY OF SYMMETRIC NORMS
Let X be a real F-space with the F-norm IIxII and with an unconditional basis {en}. The norm IIxII is called symmetric (see Singer, 1961, 1962) if, for any permutation {pn} and for an arbitrary sequence {En} of numbers equal either to 1 or to -1, the following equality holds :
ltlel+ ... +tnenll = IIe1t1ep,+ ... As follows from the definition of symmetric norms, the operator U defined by the formula
U(tte1+ ... +tnen+ ...) = e1tlep, +...
(9.8.1)
is an isometry of the space X onto itself. We shall show that if X is not isomorphic to a Hilbert space, then each isometry is of type (9.8.1).
Chapter 9
422
We say that a subspace Z of the space X of codimension 1 is a plane of symmetry if there is an isometry U I such that U(x) = x for x e Z. Let Z be a plane of symmetry and let V be an isometry. Then V(Z) is also a plane of symmetry. Indeed, let W = VUV-1. The operator W is an isometry different from I. Let x = V(y), y a Z. Then
W(x) = VUV-1V(y) = VU(y) = V(y) = X. If a Banach space Xhas a symmetric basis {en}, then the planes
Ai = {x: xi = 0}, Ai,i+ _ (X: xi = Xi},
Ai,i- = {X: xi = -xi}, where
x = xiel I xze2+ ... +xnen+ ... are planes of symmetry. Let us suppose that P is an arbitrary plane of symmetry. Let n be an arbitrary positive integer and let it
Xo=Pn(nAs). i=1
Let us consider the quotient space Xi = X/Xo. The space X1 is (n-f-1)-dimensional. The symmetries which have planes Ai, Ai,i,+, Ai,i,- as planes of symmetry imply that there is a basis {ek}, k = 1, 2, ..., n+1), in Xi such that the group Sn of operators of the type U(t1e1+ ... +to+l en+i) = (e1 tie,i+entoer,,,+to+1en+1) (9.8.2)
is contained in the group of isometries G. In virtue of Theorem 9.5.1 there is an ellipsoid invariant with respect to G. Since S. C G, this ellipsoid is described by the equation
a(xi+ ... +xn)+bxn+i z( 1. where x = x1ez+ ... +xn+len+1'
(9.8.3)
F-Norms and Isometries in F-Spaces
423
Since the replacing of by a subgroup of the group of isometries, we can assume without loss of generality that the invariant ellipsoid has the equation
xi-f ... +X,2 +J = 1.
(9.8.4)
Now we shall prove LEMMA 9.8.1. Let Xl be an (n + 1)-dimensional real Banach space with norm IIxJ1. Let the group of isometries G contain the group Sn. If the group G is in-
finite, then the intersection of the sphere S = {x: IIxMM = 1} with the subspace X' spanned by elements ei, ... , e , is a sphere in the Euclidean sense. Proof. Since the space Xl is finite-dimensional, the group G is a compact Lee group. Thus G contains a one parameter group g(t). Obviously, there is an element xo, IIxoll = 1 such that g(t)xo defines a homeomorphism between an open interval (-e, E), e > 0, and a subset of points of S. Now we have two possibilities : (1) g(t)xo-xo 0 X (2) g(t)xo-xo e X'. Since Sn C G, we can find in the first case n locally linearly independent trajectories (in the second case (n-1)). This implies that there is a neighbourhood U of the point x0 such that for each x e U (in the second case
for x e Un X') there is an isometry A such that A(xo) = x. This implies that the group G (resp. the group G' of isometries of X' is) transitive. This implies the lemma (cf. Section 5).
0
Lemma 9.8.1 implies that the group G of isometries of the space Xl is
finite or that the quotient Xl n As are Hilbert spaces for n = 1, 2, ... The i=
second case trivially implies that the space Xis a Hilbert space. Let us now consider the first case, i.e. the case where the group of isometries G of the space Xl is finite. By (9.8.4) we can assume without loss of generality that the group G is contained in the group of orthogonal transformation of the space Xl. LEMMA 9.8.2. Let Xl be an (n+ 1)-dimensional real Banach space with the norm IIxH. Let Sn C G C G,,+1. Let P be a plane of symmetry determined by
Chapter 9
424
an isometry U c G. 1 hen P is either of type A' or of type Ai,3f (i, j = 1, 2, ..., n+1) provided n is greater than 71 Proof. To begin with, let us assume that the plane P does not contain the element en+,. Let PO be the plane of symmetry determined by an isometry belonging to G and such that en+, 0 PO and PO is nearest to en+, (nearest in the classical Euclidean sense). Let
PO = {X: alx,+ ... +an+lxn+ 1 =
O},
where
ai- ...
(9.8.5)
1.
0. The planes A' (i = 1, 2, ..., n) contain the Since en+1 0 P., element e,,+1. Therefore, the angle between P, and At ought to be of type it/n, because otherwise, composing symmetries with respect to PO and At, we could obtain a plane of symmetry P, nearer to en+1 than Pa. This implies that
at=cos 7rn-
i=1,2,...
Hence either at = 0 or jail > 1/2. Therefore, by (9.8.5) we have the following possibilities :
(1) aa+11 = 1,
(2) Ia.+1I = 1/j/2, there is an i, such that laill = z , (3) an+1 I = 2 , there is an i1 such that Iaii l = 3
2
(4) Ian+II = 1/1/f, there are i1 and i2 such that laill = lai2l = 2, (5) an+ll = 2, there are ii and i2 such that laill , s , jail = 2' (6) Ian+1i = 2 , there are i,, i2, i3 such that jail = ai,I = lai,l = 2, (7) an+II = 2 , there is an i, such that lai,I and at = 0 otherwise. We shall show that only cases (1) and (7) are possible. Let us take indices j1,j2,j3 such that I jkl < n, 3k - i,,, for all k and m. This is possible since n > 7. Let ai = (al, ..., an+1), where an+1 = an+1, a,k = aik and a, = 0 otherwise. The plane
P = {x: a x1+
... +an+1 xn+1 = 0}
1 Indeed, Lemma 9.8.2 holds also for n = 2,4,5,6,7. It does not hold for n = 3, but for our purpose it is sufficient to show this for sufficiently large n.
F-Norms and Isometries in F-Spaces
425
is also a plane of symmetry. It does not contain en+1, but its distance from that point is exactly the same as P0. Therefore, the angle between P° and P' ought be of the type 2ir/n, because otherwise, composing symmetries with respect to these two planes, we could find a plane P1 nearer to en+1, such that en+1o P1. The cosinus of the angle between P° and P' is equal to 3/4 in case (2), to 1/4 in cases (3), (5), (6), and to 1/2 in case (4); this eliminates cases (2), (3), (5), (6). Let us take jl = i1 and j2 zk i2 ; then the cosinus of the angle between the respective plane and P° is equal either to 3/4 or to 1/4. This eliminates case (4). Finally, only cases (1) and (7) are possible. So far we have assumed that the plane P does not contain en+1. Suppose now that en+1 E P. Let P° = P n X', where X' is the space spanned by the elements e1, ..., en. P° is a plane of symmetry in the space X', and restricting all considerations to the space X' we are able to prove our lemma. THEOREM 9.8.3. Let X be a real infinite-dimensional F-space with a basis {en} and with a symmetric norm IIxII Then either X is a Hilbert space or each isometry is of type (9.8.1. Proof. As it follows from the previous considerations, if X is not a Hilbert space, then the planes Ai, A'.2+, A'°'- are all possible planes of symmetry.
Let us denote the isometrics corresponding to At, Ai.i+, A''''- by Si, Si.i-, respectively. Let U be an arbitrary isometry. Then U(A1), U(A1''+), U(Ai''-) are planes of symmetry corresponding to the isometries USIU-1, USi.j+U-1, USi,i-U-1, respectively. Therefore, those isometrics are of type Si, Si.j+, Let us denote the class of all such isometrics by U Let A, B e 2C be such commutative isometrics that there
is one and only one isometry C e 1 such that AC = CB. Then A = Si, B = SJ, C = Si". -This implies that each isometry US' U-1 is of the type Si. Thus U is of the type (9.8.1).
We shall now consider the spaces over complexes. Let X be a complex F-space with basis {en} and norm IIxII The norm IIxII is called symmetric
if, for any permutation of positive integers pn and for any sequence of complex numbers {en}, IB"I = 1, the following holds :
Iltle1+ ... +tnen+ ...II = IIE1t1e,,+ ...
I1.
426
Chapter 9
Obviously, if the norm IIxII is symmetric, then each operator of the type (9.8.1) (where E. are complex numbers of moduli 1) is an isometry. In the same way as in the real case we define planes of symmetry. LEMMA 9.8.4. Let X. be an (n+1)-dimensional complex F-space with basis {e,, ..., en+,} and norm IIxII If the group of isometries G contains all operators of type (9.8.2) (where ej are complex numbers of moduli 1), then either G consists of operators of type (9.8.1) or G contains all orthogonal transformations which map the space generated by e,, ..., en onto itself.
Proof. Suppose that an isometry V maps an element e;, 1 < i < n, on an element x, which is not of the type e5. Without loss of generality we may assume that the first n coordinates of x, are reals. Let us now consider the real space spanned by the elements e,,..., en, x,. Applying Lemma 9.8.2, we find that the intersection of the set {x: IIxII = R} with the space spanned by e,, ..., en is a sphere. This implies the theorem. Lemma 9.8.4 implies in the same manner as in the real case the following :
THEOREM 9.8.5. Let X be an infinite-dimensional complex F-space with basis {en} and the symmetric norm IIxII If X is not a Hilbert space, then each isometry is of type (9.8.1). COROLLARY 9.8.6. The symmetric norms are maximal.
9.9 UNIVERSALITY WITH RESPECT TO ISOMETRY
We shall say that an F-space Xn with the F-norm IIxII is universal with respect to isometry for a class U of 'F-spaces if, for any F-space X E 2t, there is a subspace Y of the space X. and a linear isometry U mapping X onto Y. PROPOSITION 9.9.1. There is no F-space Xn universal with respect to isometry for all one-dimensional F-spaces.
F-Norms and Isometries in F-Spaces
427
Proof. Let Xn be the real line with the following F-norm :
ItI < 1, for 1
that II tell n
= 11t 1j,,, where IIXI I denotes the norm in X. Hence II2enII = 11211n
1
_ - n> 0. On the other hand, IIenII = 11 1 I In = I does not tend to 0. There-
fore, the multiplication by scalars is not continuous, and this leads to a contradiction since X. is an F-space. All one-dimensional Banach spaces are isometric ; hence, the real line
(the complex plane in the case of complex one dimensional Banach spaces) with the usual norm Itl is universal for all one-dimensional Banach spaces with respect to isometry. Banach and Mazur (1933) proved that the space C[0,1] is universal with respect to isometry for all separable Banach spaces. As a particular case we find that C[0,1] is universal with respect to isometry for all twodimensional Banach spaces. The space C[0,1], however, is infinite-dimensional. Hence, the following problem arises. Does there exist an n-dimensional Banach space universal with respect to isometry for all two-dimensional Banach spaces ? The answer is negative. It was given for n = 3 by Grunbaum (1958) and for all positive integers n by Bessaga (1958).
Of course, it is enough to restrict ourselves to real two-dimensional Banach spaces and in the rest of this section only real Banach spaces will be considered. LEMMA 9.9.2 (Bessaga, 1958). Let Z be a bounded set in the n-dimensional real Euclidean space. Let f,, ..., fm map Z into the (n+1)-dimensional real Euclidean space. If the set
A (Z) U A (Z) U ... U fm (Z) contains an open set, then at least one of the functions fl, ..., fm is not Lipschitzian.
Chapter 9
428
Proof. Let M(A, s) = max {p: there are xi, i = 1, 2,..., p,
xieA,Ilxs-xjII>E} (compare Section 6.1). If the set A is n-dimensional and bounded, then by a simple calculation we find that 71
M(A,E)
(9.9.1)
.
On the other hand, if A contains an open set of dimension (n+1), we find that there is a positive Kl such that 1
n+i
M(A, s) > K1 - I
(9.9.2)
.
Let f(z) be a Lipschitzian function defined on A and let L denote the Lipschitz coefficient of the function f Then
M(A,E)
M(f(A),e)
(9.9.3)
Suppose that the functions fl, ..., fm are Lipschitzian. Then by (9.9.3) there is a positive constant K2 such that
M(.fi(Z) V ... V fm(Z), e) < K2(E )n This leads to a contradiction of formula (9.9.2). THEOREM 9.9.3 (Bessaga, 1958). There exists no finite-dimensional Banach space Xo universal with respect to isometry for all two-dimensional real Banach spaces. Proof. Suppose that such a universal space (X0, II I) exists. Let dim Xo = n
and let
Q={xeX0:Ilxll<1} be a unit ball in the space X0. Q is a convex centrally symmetric body in the n-dimensional Euclidean space. Let la-bj denote the Euclidean dis-
F-Norms and Isometries in F-Spaces
429
tance between the points a and b. Let Jai = la-0l. Let r = inf jai and aEQ
R = sup Jai. aEQ
Let 8 denote the class of all convex centrally symmetric plane figures
A such that r < a < R for all a e A. Let d, d' e Let
p (d, d') = inf sup ja-a' l , U
where U is an isometry (in the Euclidean sense) a e d, a' e U(A') and a, a' are collinear. The quantity p(A, d') is called the distance between d and d'. We shall now introduce the distance between two subspaces in the following way :
d(M, L) = max (sup inf Ix-yl, IIz!I=1 yEM
xcL
sup inf Ix-yD). IIyII=1 zEL
yEM
The set of all two-dimensional subspaces with this metric is a compact
(n dimensional set.
2)_ By a simple calculation we can prove that there is a positive constant K such that
p(A(L),d(M) < Kd(M,L),
(9.9.4)
where
d(L)=L n {xeXo: IlxHI= l}
and A(M)=Mn {xeXo: JlxlI= 1}. Let p = (n) +1 and let S2 be the set of all 2(p+l)-gons contained in 2
Let d e Q. Let us number all vertices of d in the positive orientation a1, ..., a2n+2. Let p, =
area of the triangle Oa{ a,'+1
total area of d
i= 1,2,...,p.
The numbers pl, ..., pp are affine invariant. The function h(a1,
..., ap+1) = (pi, ..., pp)
Chapter 9
430
is a Lipschitzian function of vertices a,, ..., ap+,. In this way we can assign to each d functions h,(d), ..., hp(d) depen dent on the point from which we begin the numbering. If do a Q, then we can find such positive Eao and a positive M4, such
that if, for d, d' p(do,4) < ej,
and
P (d o,d') < Ed, ,
then we can establish a correspondence between the vertices of d and d' ai -,at so that
Iai-ail < M4,p(d,d'). Let U be an affine transformation of the plane such that, for d, d' e we have U(A), U(d') e S. Then p(U(d), U(A'))
that there is a positive constant K such that
i = 1, 2, ...,p.
lhi(d(L)-hj(d(M))j < Kd(M,L),
It is easy to verify that each of the functions hi maps Q into a set containing the interior. Thus the set h,(A) u ... u hp(d)
contains an open p- dimensional dimensional set. Since
=
)
2
-( +
-1, this leads to
a contradiction of Lemma 9.9.2. The following problem are strictly connected with Theorem 9.9.3. Problem 9.9.4. Let X be a separable Banach space universal with respect to isometry for all finite-dimensional Banach spaces. Is Xuniversal for all separable Banach spaces with respect to isometry ? Problem 9.9.5. Let X be a separable Banach spaces universal with respect to isometry for all two-dimensional Banach spaces. Is X universal with respect to isometry for all finite-dimensional spaces ?
F-Norms and Isometrics in F-Spaces
431
Problem 9.9.6. How can we show a minimal number of two-dimensional real Banach spaces X1, ..., Xk(n) such that there exists no n-dimensional real Banach space universal with respect to isometry ?
The existence of such numbers follows from Theorem 9.9.3. There is
a conjecture that k(n) _
(2) + 1 (in a particular case k(3) = 4).
Klee (1960) has investigated the following related problem : What is the minimal dimension of a body Kuniversal in the affine sense for all polyhedra of dimension n with r faces (r vertices) ? More precisely : We say that a finite-dimensional convex body Kis a-universal for a class of convex bodies CC provided each member of CC is affinely equivalent to some proper section of K; and Kis centrally a-universal for CC provided K is centrally symmetric and every centrally symmetric member of 9C is affinely equivalent to a central section of K. Replacing affine equivalence by similarity leads to the notion of s-universality and central s-universality. z,V (n, r) or f (n, r) is the smallest integer k such that there is a k-dimensional convex body K x-universal for all n-dimensional convex polyhedra having r+ 1 vertices, or r+ l maximal faces, respectively.
Yia.V (n, r), or 7a,f (n, r) is the smallest integer k such that, there is a k-di-
mensional convex centrally symmetric body, centrally a-universal for all centrally symmetric n-dimensional polyhedra having 2r vertices or 2r maximal faces, respectively. Klee (1960) has shown the following estimation : n
+ oo > a.v (n, r) > -;j+-1 (r+ 1)
<
a,f r) < r, (n,
+oo > 27a'9(n,r) > yia,f(n,r) = r, 11
+oo > s'°(n, r) > n-
+l
(r+2) < E 'f(n, r) < +oo.
The following proposition is strictly connected with Problem 9.9.6. PROPOSITION 9.9.7 (Rolewicz, 1966). There are three plane, centrally sym-
metric convex figures P1, P21 P3 such that there exists no centrally symmetric convex three-dimensional body K admitting a line L through its
Chapter 9
432
centre and sections Pi, P, P3 through L such that Pf is affinely equivalent
to P' , j= 19253. Proof. Let Pl be a square, P2 circle and P3 a square with side 2 with corners rounded off by circular arcs of radius E (see Figure 9.9.1). We shall show that for a sufficiently small E the required body K could not exist.
Fig. 9.9.1
Suppose that such a body K exists for all E. Let t be a point at which L intersects the boundary K. Since t e P2 n P3, there is only one supporting hyperplane of K at t. Thus t cannot be an extremal point (corner) of
Pi. The point t must belong to one of the curved arcs of P. Otherwise t would be a flat point of the body K, and this is impossible, because P. is strictly convex.
Since we are only interested in affine properties, we may assume without loss of generality that P2 is a unit disc in the plane z = 0, P3 is similar to P3 and situated in the plane y = 0, and Pi is a parallelogram situated in a certain plane z = ay. (Here we put the x-axis as L.) Now we shall investigate the relationship between a and E. Let pi denote the boundary of Pi, i = 1, 2, 3. The three curves pi, P2, P3 intersect at point t. Since P3 is a normal section of K, the normal curvature x3 of the boundary in the direction of p3 is equal to the total curvature of p3 at t, i.e. it is equal to 1/E. The total curvature x2 of p2 is equal to 1 ; hence, the normal curvature of the boundary of K in this direction is at most 1.
The set K is convex and pl is a straight line the neighbourhood of t, hence, the minimal curvature of the boundary of K is equal to 0. The maximal curvature is in the direction perpendicular to pl. Let us denote
F-Norms and Isometries in F-Spaces
433
the maximal curvature by k. Using Euler's formula, we can express the normal curvature of the boundary of K in the directions of p2 and p3 as K2 = k sin2j9, K3 = k cos2j9,
where fl is the angle between P2 and pl at point t. Therefore K3 K2
= Cot21
.
On the other hand
Hence P tends to 0 as e->0 ; this implies that a-0, as -->O. Since for all a the points (1,0,0) and (-1,0,0) belong to K, it follows that for a-0 the set Pi tends to P2. This leads to a contradiction, because Pz is a circle and Pi is a parallelogram.
References
Alaoglu, L.: (1940), 'Weak Topologies of Normed spaces', Ann. of Math. (2) 41, 252-267.
Albinus, G.: (1970), 'Uber lokal radialbeschrankte oder lokal pseudokonvexe metrisierbare Raume', Math. Nachr. 45, 363-372. Antosik, P. and Mikusinski J.: (1968), 'On Hermite Expansions', Bull. Acad. Pol. Sci. 16, 787-791. Aoki, T.: (1942), 'Locally Bounded Linear Topological Spaces', Proc. Imp. Acad. Tokyo 18, No. 10. Arnold, L.: (1966), Convergence in Probability of Random Power Series and a Related Problem in Linear Topological Spaces, Stat. Lab. Publ. Michigan State University, East Lansing, Michigan. Arnold, L.: (1966b), 'U7ber die Konvergenz einer zufalligen Potenzreihe', Jour. Reine Angew. Math. 222, 79-112. Aronszajn, N. and Szeptycki, P.: (1966),'On General Integral Transformation', Math. Ann. 163, 127-154. Atkinson, F. V.: (1951), 'Normal Solvability of Operators in Normed Spaces' (in Russian), Matem. Sb. 28, 3-14. Atkinson, F. V.: (1953), 'On Relatively Regular Operators', Acta, Sci. Math. Szeged 15, 28-56. Auerbach, H.: (1933-35), 'Sur les groupes lineaires bornes', Studia Math. 1-4, 113127; 11-4,158-166 ; 111-5, 43-49.
Banach, S.: (1922), 'Sur les operations dans les ensembles abstraits et leur application aux equations integrales (These de doctorat)', Fund. Math. 3, 133-181. Banach, S.: (1929), 'Sur les functionelles lineaires', Studia Math. I-1, 211-216 11-1, 223239.
Banach, S.: (1932), Theorie des operations lineaires, Monografie Matematyczne I. Warszawa. Banach, S. (1932b), 'Sur les transformations biunivoques' , Fund. Math. 19, 10-16. Ba nach, S. (1948), Lectures of Functional Analysis (in Ukrainian) (It is an extended version of Banach, 1932), Kiev. Banach, S. and Mazur, S.: (1933),'Sur la dimension lineaires des espaces fonctionnels', C. R. Acad. Sci. Paris 196, 86-88.
References
435
Banach, S. and Mazur, S.: (1933b),'Zur Theorie derlinearen Dimension', Studia Math. 4,100-112. Banach, S. and Steinhaus, H. : (1927), 'Sur le principe de la condensations de singularites', Fund. Math. 9, 50-61.
Bartle, R., Dunford, N. and Schwartz, J.: (1955) 'Weak Compactness and Vector Measures', Canad. Jour. Math. 7, 289-305. Beck, A.: (1962), 'A Convexity Condition in Banach Spaces and the Strong Law of Large Numbers', Proc. Amer. Math. Soc. 13, 329-334. Bessaga, C.: (1958),'A Note on Universal Banach Space of a Finite Dimension', Bull. Acad. Pol. Sci. 6, 97-101. Bessaga, C.: (1969),'Some Remarks on Dragilev's Theorem', Studia Math. 31,307-318. Bessaga, C.: (1976), 'A Nuclear Frechet Space Without Basis I. Variation on a Theme of Djakov and Mitiagin', Bull. Acad. Pol. Sci. 24,471-473. Bessaga, C.: (1980), 'A Lipschitz Invariant of Normed Linear Spaces Related to the Entropy Numbers', Rocky Mountains Math. Jour. 10, 81-84. Bessaga, C. and Pelczynski, A.: (1957), 'An Extension of the Krein-Milman-Rutman Theorem Concerning Bases to the Case of Be-spaces', Bull. Acad Pol. Sci. 5, 379383.
Bessaga, C. and Pelczynski, A.: (1958), 'On Bases and Unconditional Convergence', Studia Math. 17,151-164. Bessaga, C. and Pelczynski, A.: (1960), 'On Embedding of Nuclear Spaces into the Space of All Infinite Differentiable Functions on the Real Line' (in Russian), Dokl. A.N. S.S.S.R. 134,745-748. Bessaga, C. and Pelczynski, A.: (1960b), 'Banach Spaces Non-Isomorphic to Their Cartesian Square', I. Bull. Acad. Polon. Sci. 8, 77-80. Bessaga, C., Pelczynski, A. and Rolewicz, S. : (1957), 'Some Properties of the Norm in F-Spaces', Studia Math. 16, 183-192. Bessaga, C., Pelczynski, A. and Rolewicz S.: (1957b), 'Some Properties of the Space (s)', Coll. Math. 7, 45-51. Bessaga, C., Pelczynski, A. and Rolewicz, S.: (1962), 'On Diametral Approximative Dimension and Linear Homogeneity of F-Spaces', Bull. Acad. Pol. Sci. 9, 677-683. Bessaga, C., Pelczynski, A. and Rolewicz, S.: (1963), 'Approximative Dimension of Linear Topological Spaces and some Its Applications', Studia Math., Seria Specjalna (Special series) 1, 27-29. Bessaga, C. and Retherford, J. R.: (1967), Lectures on Nuclear Spaces and Related Topics, preprint, Louisiana State University, Baton Rouge. Bessaga, C. and Rolewicz, S.: (1962), 'On Bounded Sets in F-spaces', Coll. Math. 9, 89-91.
Bohnenblust, H. F. and Sobczyk, A.: (1938), 'Extensions of Functionals on Complex Linear Spaces', Bull. Amer. Math. Soc. 44, 91-93. Bourgin, D. G.: (1943) 'Linear Topological Spaces', Amert Journ. of Math. 65, 637659.
436
References
Burzyk, J.: (1983), 'On Convergence in the Mikusifiski Operational Calculus', Stud. Math. 75, 313-333. Charzyfiski, Z.: (1953), 'Sur les transformations isometrique des espaces du type F', Studia Math. 13, 94-121. Cowie, E. R.: (1981), A Note on Uniquely Maximal Banach Spaces, preprint. Crone, L. and Robinson W. B.: (1975) `Every Nuclear Frechet Space with a Regular Basis Has the Quasi-Equivalence Property', Studia Math. 52, 203-207. Day, M. M. (1940), `The Spaces LP with 0 < p < 1', Bull. Amer. Math. Soc. 46, 816823.
Day, M. M. (1958), Normed Linear Spaces, Springer-Verlag, Berlin. Dieudonne, J.: (1954), `Sur les espaces de Montel m6trisables', Comp. Rend. Acad. Paris 238, 194-195. Dieudonne, J.: (1955), `Bounded Sets in F-Spaces', Proc. Amer. Math. Soc. 6, 729-731. Djakov, P. V.: (1974), 'On Isomorphism of Certain Spaces of Holomorphic Functions (in Russian), Proceedings of the 6-th Winter School on Mathematical Programming and Related Problems, Drogobych.
Djakov, P. V.: (1975), 'A Short Proof of Crone and Robinson Theorem on QuasiEquivalence of Regular Bases', Studia Math. 53, 269-271. Djakov, P. V. and Mityagin, B. S.: (1976) 'Modified Construction of a Nuclear Frechet Space Without a Basis', Jour. Func. Anal. 23, 415-423. Djakov, P. V. and Mityagin, B. S.: (1977) The Structure of Polynomial Ideals in the Algebra of Entire Functions, Preprint of Institute of Mathematics of the Polish Academy of Sciences, 123.
Douady, A.: :(1965), 'Uespace de Banach dont le group lineaires n'est pas connexe', Indag. Math. 25,787-789. Dragilev, M. M. : (1960),'Canonical Form of a Basis in the Space of Analytic Functions' (in Russian), Usp. Matem. Nauk, vyp. 2,92,181-188. Dragilev M. M.: (1964),'On Bases in Nuclear Spaces', Abstract of the VII-th All-Union Conference on Complex Analysis, Rostov. Dragilev, M. M.: (1965),'0.1 Regular Bases in Nuclear Spaces' (in Russian), Matem. Sb., vyp. 2, 68 (110), 153-173. Dragilev, M. M.: (1969),'On Special Dimensions Defined on Certain Classes of Kothe Spaces', (in Russian), Matem. Sb. 80 (122), 225-240.
Dragilev, M. M.: (1970), 'On Multi-Regular Bases in Kothe Spaces' (in Russian), Dokl. A.M. S.S.S.R 193, 752-755. Dragilev, M. M.: (1970b),'On Kothe Spaces Distinguishable by Diametral Dimension', (in Russian), Sib. Matein. Zhur. 11, 512-525. Drewnowski, L.: (1972), 'Topological Rings of Sets, Continuous Set Functions, In-
tegration', Bull. Acad. Pol. Sci. 20,1- 269-276,11- 277-286,111- 439-445. Drewnowski, L.: (1972b), 'Equivalence of Brooks-Jewett, Vitali-Hahn-Saks and Nikodym Theorems', Bull. Acad. Pol. Sci. 20, 725-731. Drewnowski, L.: (1973), 'Uniformly Boundness Principle for Finitely Additive Vector Measures', Bull. Acad. Pol. Sci. 21,115-118.
References
437
Drewnowski, L. : (1973h), 'On the Orlicz-Pettis Type Theorem of Kalton', Bull. Acad. Pol. Sci. 21, 515-518. Drewnowski, L.: (1974), 'On Control Submeasures and Measures', Studia Math. 50, 203-224. Drewnowski, L. and Labuda, I. : (1973), 'Sur quelques th6or6mes du type Orl icz-Petti s II' Bull. Acad. Pot. Sci. 21, 119-226. Drewnowski, L., Labuda, I. and Lipecki, Z.: (1981), `Existence of Quasi-Bases for Separable Topological Linear Spaces', Archiv der Mathematik, 37,454-456. Dubinsky, E.: (1971), Every Separable Fr6chet Space Contains a Non-Stable Dense Subspace', Studia Math. 40, 77-79. Dubinsky, E.: (1973), `Infinite Type Power Series Subspaces of Finite Type Power Series Spaces', Israel Jour. of Math. 15, 257-281. Dubinsky, E. (1975), Concrete Subspaces of Fr6chet Nuclear Spaces', Studia Math. 52, 209-219. Dubinsky, E. (1979), The Structure of Nuclear Frechet Spaces, Springer-Verlag Lecture Notes 720, Berlin. Dubinsky, E. (1980), `Basic Sequences in a Stable Finite Type of Power Series', Studia Math. 68, 117-130. Dubinsky, E.: (1981), `Nuclear Fr6chet Spaces without the Bounded Approximation Property', Studia Math. 71, 85-105. Dubinsky, E. and Ramanujan, M. S.: (1972), 'On 1-Nuclearity,' Mem. Amer. Math. Soc. 128. Dubinsky, E. and Robinson, W.: (1978), `Quotient Spaces of (s) with Basis', Studio Math. 63, 267-281. Dunford, N. and Schwartz, J.: (1958), Linear Operators, Part I: General Theory, Interscience Publishers, New York, London.
Duren, P. L., Romberg, R. G. and Shields, A. L.: (1969), `Linear Functionals in `?l"-Spaces with 0 < p < 1', Jour. Reine Angew. Math. 238, 32-60. Dvoretzky, A.: (1963), 'On Series in Linear Topological Spaces', Israel Jour. of Math. 1, 37-57. Dvoretzky, A. and Rogers, C. A.: (1950), `Absolute and Unconditional Convergence in Normed Linear Spaces', Proc. Nat. Acad. Sci. USA 36, 162-166. Dynin, A. and Mitiagin, B.: (1960), `Criterion for Nuclearity in Terms of Approximative Dimension', Bull. Acad. Pol. Sci. 8, 535-540. Eberlein, W. F.: (1947), `Weak Compactness in Banach Spaces I', Proc. Nat. Acad. Sci. USA 33, 51-53. Eidelheit, M.: (1936), 'Zur Theorie der Systeme linearer Gleichungen', Studia Math. 6, 139-148.
Eidelheit, M. and Mazur, S.: (1938),'Eine Bemerkung fiber die Raume vom Typus F', Studia Math. 7, 159-161. Fenske, Ch. and Schock, E.: (1970), 'Nuklearitat and lokale Konvexitat von Folgenraumen', Math. Nochr. 45, 327-335.
438
References
Figiel, T.: (1969), 'On Non-Linear Isometric Embeddings of Normed Linear Spaces', Bull. Acad. Pol. Sci. 16, 185-188. Figiel, T.: (1972), `An Example of Infinite-Dimensional Reflexive Banach Space NonIsomorphic to Its Cartesian Square', Studio Math. 42, 295-305. Frbchet, M.: (1926), Les espaces abstrait topologiquement affine, Acta Math, 47, 25-52. Gawurin, M. K. : (1936), `(7ber Stieltjessche Integration abstrakten Funktionen', Fund. Math. 27,254-268. Gelfand, I. M.: (1941), `Normierte Ringe', Matem. Sb. 9,41-49. Gelfand, I. M. and Vilenkin, N. Ya.: (1961), Generalized Functions 4. Some Applications of Harmonic Analysis. Equipped Hilbert Spaces, (in Russian), Gosud. Izd. Fiz.-Mat. Lit., Moscow. Gohberg, I. C. and Krein, M. G.: (1957), `Fundamental Theorems on Defect Numbers, Root Numbers and Indices of Linear Operators' (in Russian), Usp. Matem. Nauk, vyp. 2, 12, 43-118. Goldberg, A. A.: (1959), `Elementary Remarks on Formulas for the Determining of
Orders and Types of Entire Functions of Several Variables' (in Russian), Dokl. A.N. Arm. S. S. S. R 29,145-151.
Goldberg, A. A.: (1961), 'On Formulas for the Determining of Orders and Types of Entire Functions of Several Variables' (in Russian), Dokl. i Soobshcheniya uzhgorodskovo Un-ta, Seriya Fiz-Matem. 4, 101-103. Goldstine H. H.: (1931), `Weakly Compact Banach Spaces', Duke Math. Jour. 4, 125131.
Gramsch, B.: (1965), `Integration and holomorphe Funktionen in lokalbeschrankten Raumen', Math. Ann. 162,190-210.
Gramsch, B.: (1966), 'a-Transforrnationen in lokalbeschrankten Vektorraumen', Math. Ann. 165, 135-151. Gramsch, B.: (1967), `Funktionalkalkiil mehrerer Veranderlichen in lokalbeschrankten Algebren', Math. Ann. 174, 311-344. Gramsch, B.: (1967b), 'Tensorprodukte and Integration vektorwertiger Funktionen', Math. Zeitschr. 100, 106-122. Grothendieck, A.: (1951), 'Sur une notion de produit tensoriel topologique d'espaces vectoriels topologiques, et une classe remarquable d'espaces lilies a cette notion', Comp. Rend. Acad. Paris 233, 1556-1558. Grothendieck A.: (1955), 'R6sum6 des r6sultats essentiels dans le theorib des produits tensoriels topologiques et des espaces nucl8aires', Ann. Inst. Fourier 4, 73-112. Grothendieck, A. : (1955b), `Sur les espaces (97) et (Cl 9), Summa Brasil Math. 3. Grothendieck, A.: (1955), Produit tensoriels topologiques et espaces nuclearies, Mem. Amer. Math. Soc. 16.
Grothendieck, A.: (1956), 'La th6orie de Fredholm', Bull. Soc. Math. de France 84 319-384.
Grunbaum, B.: (1958), On a Problem of Mazur, Bull. Res. Cauncil of Israel 7F; 133-135.
References
439
Gurarij, W. I.: (1966),'Spaces of Universal Decomposition, Isotope Spaces and Mazur Problem on Rotation of Banach Spaces' (in Russian), Sib. Matem. Zhur. 7, 10021013.
Hahn, H.: (1927), Ober lineare Gleichungen Systeme in linearen Raumen, Jour. Reine. Agnew. Math. 157, 214-229 Holsztynski, W.: (1968), 'Linearization of Isometric Embeddings of Banach Spaces. Metric Envelopes', Bull. Acad. Pol. Sci.16,189-193. Hyers, D. M.: (1939),'Locally Bounded Linear Topological Spaces', Rev. Ci. Lima 41, 558-574. lyachen, S. O.: (1968),'Semi-Convex Spaces', Glasgow Math. Jour. 9 (2),111-118. James, R. C. : (1950), 'Bases and Reflexivity of Banach Spaces', Ann. of Math. 52, 518527.
James, R. C.: (1951),'A Non-ReflexiveBanachSpace Isometric with Its Second Conjugate', Proc. Nat. Acad. Sci. USA 37, 174-177. Kakutani, S.: (1936), 'Ober die Metrisation der topologischen Gruppen', Proc. Imp. Acad. Tokyo 12, 82-84. Kalton, N. J.: (1974), 'Basic Sequences in F-Spaces and Their Applications', Proc. Edinburgh Math. Soc. 19, 151-177. Kalton, N. J.: (1974b), 'Topologies on Riesz Groups and Applications to the Measure Theory', Proc. London Math. Soc. 3 (28), 253-273. Kalton, N. J. : (1977),'Universal Spaces and Universal Bases in Metric Linear Spaces', Studio Math. 61, 161-191. Kalton, N. J.: (1977b), 'Compact and Strictly Singular Operators on Orlicz Spaces', Israel Jour. of Math. 26, 126-137. Kalton, N. J. : (1978), 'Quotients of F-Spaces', Glasgow Math. Jour. 19,103-108. Kalton, N. J.: (1979),'A Note on Galbed Spaces', Comm. Math. 21,75-79. Kalton, N. J.: (198Q), 'An F-Space with Trivial Dual, where the Krein-Milman Theorem Holds', Israel Jour. of Math. 36, 41-50. Kalton, N. J.: (1981), 'Curves with Zero Derivatives in F-Spaces', Glasgow Math. Jour. 22, 19-30. Kalton, N. J.: (1981b),'Isomorphism between LP-Function Spaces whenp < 1', Jour. Func. Anal. 42, 299-337. Kalton, N. J. and Peck, N. T.: (1979), 'Quotients of LV(0,1) for 0 _< p < 1', Studia Math. 64, 65-75. Kalton, N. J. and Peck N. T.:, (1981), 'A Re-Examination of the Roberts Example of a Compact Convex Sets without Extreme Points', Math. Ann. 253,89-101. Kalton, N. J. and Roberts, J. W.: (1981), 'A Rigid Subspace of L°', Trans. Amer. Math. Soc. 266, 645-654. Kalton, N. J. and Shapiro, J. H.: (1975), 'An F-Space with Trivial Dual and Non-Trivial Compact Endomorphism' Israel Jour. of Math. 20,282-291. Kalton, N. J. and Shapiro, J. H.: (1976) 'Bases and Basic Sequences in F-Spaces', Studio Math. 56, 47-61.
440
References
Kalton, N. J. and Wood, G. V.: (1976), 'Orthonormal Systems in Banach Spaces and Their Applications', Math. Proc. Camb. Phil. Soc. 79,493-510. Klee, V. L.: (1952),'Invariant Metrics in Groups', Proc. Amer. Math. Soc. 3,484-487. Klee, V. L.: (1956), 'An Example in the Theory of Topological Linear Spaces', Arch. Math. 71, 362-366. Klee, V. L. : (1958),'On the Borelian and Projective Types of Linear Subspaces', Math. Scan. 6, 189-199. Klee, V. L.: (1960), 'Polyhedral Sections of Convex Bodies', Acta Math. 103,243-267. Klee, V. L.: (1961), Exotic Topologies for Linear Spaces, Symp. of General Topology and Its Relations to Modern Algebra, Prague.
Klein Ch. and Rolewicz, S.: (1984), On Riemann Integration of Functions with Values in Linear Topological Spaces, Studia Math. 80 Kolmogorov, A. N. (Kolmogoroff, A) : (1935), 'Zur Normierbarkeit eines allgemeinen topologischen Raumes', Studia Math. 5, 29-33. Kolmogorov, A. N.: (1958), 'On Linear Dimension of Topological Vector Spaces' (in Russian), Dokl. A.N. S.S.S.R. 120, 239-241.
Kolmogorov, A. N. and Tichomirov, W. M.: (1959), 'e-Entropy and c-Capacity of Sets in Functional Spaces' (in Russian), Usp. Matem. Nauk, vyp. 2, 16 (89),3-86.
Komura, T. and Komura, Y.: (1966), 'Ober die Einbettung der nuklearen Raume in (s)A', Math. Ann. 162, 284-288. Komura, Y.: (1966), 'Die Nuklearitat der Losungraume der hypoelliptischen Gleichungen', Funkcialaj Ekvacioj 9, 313-324. Kondakov, V. P.: (1974), 'On Quasi-Equivalence of Regular Bases in Kothe Spaces' (in Russian), Matem. Analiz i ego pril. RG U 5, 210-213.
Kondakov, V. P.: (1979), 'On Orderable Absolute Bases in F-Spaces' (in Russian), Dokl. A.N. S.S.S.R. 247, 543-546. Kondakov, V. P.: (1980),'On Properties of Bases in Certain Kothe Spaces' (in Russian), Funkc. Analiz i ego pril. 14, 58-59. Kondakov, V. P. : (1983), 'On Construction of Unconditional Bases in Certain Kothe Spaces (in Russian.) Studia Math. 76,137-157.
Kondakov, V. P.: (1983b), Problems of Geometry of Non-Normed Spaces (in Russian), Rostov University Publication. Krasnoselski, M. A'and Rutitski, Ya. B. : (1958), Convex Functions and Orlicz Spaces (in Russian), Gosud. Izd. Fiz.-Mat. Lit., Moscow.
Krein, M. G. and Milman, D. P.: (1940), 'On Extreme Points of Regularly Convex Sets', Stud. Math. 9, 133-138. Krein, M. G., Milman, D. P. and Rutman L. A.: (1940), 'On a Property of a Basis in a Banach Space' (in Russian), Zap. Khark. Matem. T-va 4,106-110. Krein, S. G.: (1960), 'Operators in a Scale of Banach Spaces', Rep. of Conference on Functional Analysis, Warsaw. Kwapien, S.: (1968), 'Complement au theoreme de Sazonov-Minlos', Compt. Rend. Acad. Paris 267, 698-700,
References
441
Labuda, I.: (1972), 'Sur quelques generalisation des thboremes de Nikodym et de Vitali-Hahn-Saks', Bull. Acad. Pol. Sci. 20,447-456. Labuda, I.: (1973),'Sur quelques th6oremes du type d'Orlicz-Pettis, I', Bull. Acad. Pol. Sci. 21, 127-132. Labuda, I.: (1975), 'Ensembles convexes dans les espaces d'Orlicz', Comp. Rend. Acad. Sci. Paris 281, 443-445. Labuda, I. and Lipecki, Z.: (1982), 'On Subseries Convergent Series and m-QuasiBases in Topological Linear Spaces', Manuscripta Math. 38,87-98. Landsberg, M.: (1956), 'Lineare topologische Raume die nicht lokalkonvex sind', Math. Zeitschr. 65,104-112. Leray, J. : (1950), 'Valeur propres d'un endomorphisms completement continue d'un espace vectoriels a voisinages convexes', Acta Sci. Math. Szeged, Pars B, 12, 177186.
Ligaud, J. P.: (1971), 'Solution d'un probleme de S. Rolewicz sur les espaces nucl6aires', Compt. Rend. Acad. Sci. serie A, 273, 113-114. Ligaud, J. P.: (1973), 'On example d'espace nucleaire dont le dual topologique est reduit a zero', Studia Math. 46, 149-151. Lindenstrauss, J. and Pelczyfiski, A.: (1968), 'Absolutely Summing Operators in P21Spaces and Their Applications', Studia Math. 29, 275-326. Lipecki, Z.: (1982), On Independent Sequences in Topological Linear Spaces (preprint). Lorch, E. R.: (1939),'Bicontinuous Linear Transformations in Certain Vector Spaces', Bull Amer. Math. Soc. 45, 564-569. Lusky, W.: (1976), 'Th^ Gurarij Spaces are Unique', Arch. Mat. 27, 627-635. Lusky, W.: (1979), 'A Note on Rotations in Separable Banach Spaces', Studia Math. 65, 239-242. Luxemburg, W. A. J. and Zaanen, A. C.: (1963), 'Compactness of Integral Operators in Banach Function Spaces', Math. Ann. 149 (2), 150-180.
Luxemburg, W. A. J. and Zaanen, A. C.: (1971), Riesz Spaces I, North-Holland, Amsterdam, London.
Mankiewicz, P.: (1972), 'On Extension of Isometrics in Normed Linear Spaces', Bull. Acad. Pol. Sci. 20, 367-371.
Mankiewicz, P.: (1976), 'On Isometrics in Linear Metric Spaces', Studia Math. 55, 163-173.
Mankiewicz, P.: (1979), 'Fat Equicontinuous Groups of Homeomorphisms of Linear Topological Spaces and their Application to the Problem of Isometrics in Linear Metric Spaces', Studia Math. 64, 13-23.
Marcus, M. and Woyczynski, W.: (1977), 'Domaine d'attraction normal dans les espaces de type stable p', Comp. Rend. Acad. Paris A285, 915-917.
Marcus, M. and Woyczynski, W.: (1978), 'A Necessary Condition for the Central Limit Theorem on Spaces of Stable Type', Proc. Conf. Dublin, Springer Lecture Notes 644, 327-339. Marcus, M. and Woyczynski, W.: (1979), 'Stable Measures and Central Limit Theorems in Spaces of Stable Type', Tran. Amer. Math. Soc. 251, 71-102.
442
References
Matuszewska, W. and Orlicz, W.: (1961), 'A Note on the Theory of s-Normed Spa ces of Integrable Functions' Studia Math. 21, 107-115. Matuszewska, W. and Orlicz, W.: (1968), 'A Note on Modular Spaces IX', Bull. Acad. Pol. Sci. 16, 801-807. Maurey, B.: (1972), 'Integration dans les espaces p-normes', Ann. Sci. Norm. Sup. Pisa 26 (4), 911-931.
Maurey, B. and Pisier, G.: (1973), 'Un theoreme d'extrapolation et ses consequences', Comp. Rend. Acad. Sci. Paris 277, 39-42. Maurey, B. and Pisier, G.: (1976), 'S6ries de variables aleatoires vectorielles independantes et properties geometriques desespacesdeBanach', Studia Math. 58,45-90. Mazur, S.: (1930), 'Ober die kleinste konvexe Mengen, die eine gegebene kompakte Menge enthalt', Studia Math. 2, 7-9. Mazur, S.: (1938), 'Sur les anneaux lineaires', Comp. Rend. Acad. Paris 207, 10251027.
Mazur, S. and Orlicz, W.: (1933), 'Ober Folgen linearen Operationen', Studia Math. 4, 152-157. Mazur, S. and Orlicz, W.: (1948), 'Sur les esnaces lineaires metriques I', Studia Math. 10, 184-208.
Mazur, S. and Orlicz, W.: (1953), 'Sur les espaces lineaires metriques 11', Studia Math. 13, 137-179. Mazur, S. and Orlicz, W.: (1958), 'On Some Classes of Linear Spaces', Studia Math. 27, 97-119.
Mazur, S. and Sternbach, L.: (1933), 'Ober die Borelschen Typen von linearen Mengen', Studia Math. 4, 48-53.
Mazur, S. and Ulam, S.: (1932), 'Sur les transformations isom6triques d'espaces vectoriels normes', Comp. Rend. Acad. Paris 194, 946-948. Metzler, R. C.: (1967), 'A Remark on Bounded Sets in Linear Topological Spaces', Bull. Acad. Pol. Sci. 15, 317-318. Mityagin, B. S.: (1960), 'Relation between e-Entropy, Speed of Approximation and Nuclearity of a Compact in a Linear Space' (in Russian), Dokl. A.N. S.S.S.R. 134,765-768. Mityagin, B. S.: (1961), 'Approximative Dimension and Bases in Nuclear Spaces' (in Russian), Usp. Matem. Nauk 16, 63-132. Mityagin, B. S.: (1969), 'Sur ('equivalence des bases inconditionelles dans les 6chelles de Hilbert', Comp. Rend. Acad. Paris 269, 426-428. Mityagin, B. S.: (1970), 'Frechet Spaces with a Unique Unconditional Basis', Studia Math. 38, 23-34. Mityagin, B. S.: (1971), 'Equivalence of Unconditional Bases in Hilbert Scales' (in Russian), Studia Math. 37, 111-137.
Mityagin, B. S. and Henkin, G. M.: (1963), 'Estimations between Diameters of Different Types (in Russian), Trudy Sem. Funkc. Analiza Voronezhskogo Univ. 7,97-103.
References
443
Mityagin, B. S. and Henkin, G. M.: (1970), `Linear Decomposition of Singularities and a Problem of Isomorphism of Holomorphic Functions' (in Russian), Usp. Matem. Nauk 25, 227-228. Mityagin, B. S. and Henkin, G. M.: (1971), `Linear Problems of Complex Analysis', (in Russian) Usp. Matem. Nauk 26 (4), 93-152. Mityagin, B. S. and Zobin, N. M. : (1974), 'Contre-Exemple a l'existence d'une base dans un espace de Fr6chet nucleaire', C. R. Acad. Sci. Paris, Ser. A 279, 255-258, 325-327.
Moscatelli, B.: (1978), 'On the Existence of Universal .%-Nuclear Fr6chet Spaces', Jour. Reine Angew. Math. 301, 1-26.
Musial, K., Ryll-Nardzewski, C. and Woyczynski, W. A.: (1974), 'Convergence presque sure des s6ries al6atoires vectorielles a multiplicateur born6s', Comp. Rend. Acad. Paris 279, 225-228.
Musielak, J.: (1978), Modular Spaces (in Polish), Wyd. UAM, Pozna6. Musielak, J. and Orlicz, W.: (1959), 'On Modular Spaces', Studia Math. 18, 49-65. Musielak, J. and Orlicz, W.: (1959b), 'Some Remarks on Modular Spaces', Bull. Acad. Pol. Sci. 7, 661-668. Nakano, H.: (1950), Modulared Semiordered Spaces, Maruzen Co. Ltd., Tokyo.
Ogrodzka, Z.: (1967), 'On Simultaneous Extensions of Infinitely Differentiable Functions', Studio Math. 28, 193-207. Orlicz, W.: (1929), 'Beitrage zur Theorie der Orthogonalentwicklungen II', Studia Math. 1, 241-255.
Orlicz, W.: (1933), 'Ober unbedingte Konvergenz in Funktionenraumen', Studia
Math. 4, I - 33-37, II - 41-47. Orlicz, W.: (1933b),'Uber die Divergenz von allgemeinen Orthogonalreihen', Studia Math. 4, 27-32. Orlicz, W.: (1936), 'Beitrage zur Theorie der Orthogonalentwicklungen V', Studia Math. 6, 20-38. Orlicz, W.: (1951),'On a Class of Assymptotically Divergent Sequences of Functions', Studia Math. 12, 286-307. Orlicz, W.: (1955), 'On Perfectly Convergent Series' (in Polish), Prace Matem. 1, 393-414.
Paley, R. and Zygmund, A.: (1932), 'A Note on Analytic Functions in the Unit Circle', Proc. Cambr. Phil. Soc. 28, 266-272.
Pallaschke, D.: (1973), 'The Compact Endomorphism of the Metric Linear Space )' ', Studia Math. 47, 123-133. Peck, N. T.: (1965), 'On Non-Locally Convex Spaces 1', Math. Ann. 161, 102-115. Peck, N. T.: (1968), 'On Non-Locally Convex Spaces II', Math. Ann. 178, 209-218. Pelczynski, A.: (1957), 'On the Approximation of S-Spaces by Finite-Dimensional Spaces', Bull. Acad. Pol. Sci. 5, 879-881.
Pelczyriski, A.: (1962), 'A Proof of Grothendieck Theorem on Characterization of Nuclear Spaces' (in Russian), Prace Matem. 7, 155-167.
444
References
Pelczynski, A.: (1969), `Universal Bases', Studia Math. 32, 247-269. Pelczynski, A. and Rolewicz, S.: (1962), Best Norms with Respect to Isometry Groups in Normed Linear Spaces, Short communication on International Math. Congress in Stockholm, 104. Pelczynski, A. and Semadeni, Z.: (1960), `Banach Spaces Non-Isomorphic to Their Cartesian Square II', Bull. Acad. Pol. Sci. 8, 81-84.
Pelczynski, A. and Singer, I.: (1964), 'On Non-Equivalent Bases and Conditional Bases in Banach Spaces', Studia Math. 25, 5-25.
Petrov, V. V.: (1975), Sums of Independent Random
Variables, Springer-Verlag,
Berlin.
Pettis, B. J.: (1938),'OnIntegration in Vector Spaces', Trans. Amer. Math. Soc. 44, 277-304.
Pettis, B. J.: (1939), `Absolutely Continuous Functions in Vector Spaces (abstract)', Bull. Amer. Math. Soc. 45, 677. Phelps, R. R.: (1966), Lectures on Choquet's Theorem, Van Noorstrand Co., Princeton. Pietsch, A.: (1963), `Absolut summierende Abbildungen in lokalkonvexen Raumen', Math. Nachr. 27, 77-103. Pietsch, A.: (1965), Nukleare lokalkonvexe Raumen, Akademie Verlag, Berlin.
Pietsch, A.: (1967), `Absolut p-summierende Abbildungen in normierten Raumen', Studia Math. 28, 333-353.
Popov, M. M.: (1984) On Codimension of Subspaces LD(y), p
Przeworska-Rolewicz, D. and Rolewicz, S.: (1968), Equations in Linear Spaces, Mon. Matem. 47, PWN, Warszawa. Raikov, D. A.: (1957), `On a Property of Nuclear Spaces' (in Russian), Usp. Matem. Nauk, vyp. 5, 12 (77), 231-236. Ramanujan, M. S.: (1970), `Power Series Spaces A(a) and associated A(a) Nuclearity', Math. Ann. 189, 161-168. Ritt, J. F.: (1928), `Certain Points in the Theory of Dirichlet Series', Amer. Jour. of Math. 50, 73-86. Roberts, J. W.: (1976), Pathological Compact Convex Sets in the Space LP, 0
Roberts, J. W.: (1977), 'A Compact Convex Set with no Extreme Points', Studia Math. 60, 255-266. Rolewicz, S.: (1957), `On a Certain Class of Linear Metric Spaces', Bull. Acad. Pol. Sci. 5, 471-473. Rolewicz, S.: (1959), `Some Remarks on the Spaces N(L) and N(l)', Studia Math. 18, 1-9.
References
445
Rolewicz, S.: (1959b), 'On Functions with Derivatives Equal to 0' (in Polish), Wiadomoici Matem. 3, 127-128.
Rolewicz, S.: (1959c), 'Remarks on Linear Metric Montel Spaces', Bull. Acad. Pol. Sci. 7, 195-197.
Rolewicz, S.: (1960), 'On Some Generalization of Dvoretzky-Rogers Theorem', Coll. Math. 8, 103-106. Rolewicz, S.: (1961), 'On the Characterization of Schwartz Spaces by Properties of the Norm', Studia Math. 20, 87-92. Rolewicz, S.: (1962),'On Spaces of Holomorphic Functions', Stud. Math. 21, 135-160.
Rolewicz, S.: (1962), 'On Cauchy-Hadamard Formula for Kothe Power Spaces', Bull. Acad. Pol. Sci. 10, 211-216. Rolewicz, S.: (1966), 'Plane Sections of Centrally Symmetric Convex Bodies', Israel Jour. of Math. 4, 135-138. Rolewicz, S.: (1968), 'A Generalization of the Mazur-Ulam Theorem', Studio Math. 31, 501-505.
Rolewicz, S.: (1971), 'An Example of a Normed Space Non-Isomorphic to Its Product by the Real Line', Studia Math. 40, 71-75. Rolewicz, S.: (1976), Funktionalanalysis and Steurungstheorie, Springer-Verlag, Berlin.
Rolewicz, S. and Ryll-Nardzewski, Cz.: (1967), 'On Unconditional Convergence in Linear Metric Spaces', Coll. Math. 17, 327-33 1. Rosenberger, B.: (1972),'O-nukleare Raume', Math. Nachr. 52, 147-160.
Rosenberger, B.: (1973), 'Universal Generator for Varieties of Nuclear Spaces', Trans. Amer. Math. Soc. 184, 275-290.
Ryll-Nardzewski, Cz.: (1962), 'Example of a Non-Separable Bo-Space in which every Bounded Set is Separable', Coll. Math. 9, 93-94.
Ryll-Nardzewski, Cz. and Woyczynski, W.: (1974), Convergence en measure des series aletoires vectoriels a multiplicateur borne, Sem. Maurey-Schwartz 1973-1974, Annexe Juin 1974.
Schauder, J.: (1927), 'Zur Theorie stetiger Abbildungen in Funktionalraumen', Math. Zeitr. 26, 47-65. Schauder, J.: (1930), 'Cber lineare, vollstetige Funktionaloperatoren', Studia Math. 2, 183-196.
Schwartz, L.: (1953), 'Homomorphismes et applications continue', Comp. Rend. Acad. Paris 236, 2472-2473.
Schwartz, L.: (1969), 'Un th6oreme de la convergence dans les LP, 0-
Shapiro, J.H.: (1969), 'Examples of Proper Closed Weakly Dense Subspaces in Non-Locally Convex F-spaces', Israel Jour. of Math. 7, 369-380.
Shapiro, J.H.: (1977), 'Remarks on F-Spaces of Analytic Functions', Proc. of Kent. State Conf, Springer Lecture Notes 604, 107-124. Shapiro, J. H.: (1978),'Subspaces of LP(G) Spanned by Characters : O
446
References
Sierpifiski, W.: (1928), 'Sur les ensembles complets d'un espace (D)', Fund. Math. 11,203-205. Simmons, S.: (1964), 'Boundness in Linear Topological Spaces', Trans. Amer. Math. Soc. 113, 169-180. Singer, I.: (1961), 'On Banach Spaces with Symmetric Basis' (in Russian), Rev. Math. Pure Appl. (Acad. RPR) 6, 159-166. Singer, I.: (1962), `Some Characterization of Symmetric Bases in Banach Spaces', Bull. Acad. Pol. Sci. 10, 185-192. Slowikowski, W.: (1957), `On cS and cbci-Spaces', Bull. Acad. Pol. Sci. 5, 599-600. 9mulian, V.: (1940), 'Ober lineare topologische Raume', Matem. Sb. 7 (49), 425-448. Srinivasan, V.K.: (1966), Funktional Analysis and Its Applications, Ph.D. thesis at Univ. of Madras.
Sundaresan, K. and Woyczyfiski, W.: (1980), `Laws of Large Numbers and Beck Convexity in Metric Linear Spaces', Jour. of Multivariate Analysis 10, 442-457. Stiles, W. J.: (1970), `On Properties of Subspaces of 19, 0
Stiles, W. J.: (1971), 'On Non-Locally p-Convex Spaces,' Coll. Math. 23, 261-262. Szeptycki, P.: (1968), `On Functions and Measures Whose Fourier Transforms are Functions', Math. Ann. 179, 31-41.
Szeptycki, P.: (1979), `Domain of Integral Transformation on General Measure Spaces', Math. Ann. 242, 267-271.
Szeptycki, P.: (1980), Notes on Integral Transformation, Institute of Mathematics of the Polish Academy of Sciences, preprint 214.
Szeptycki, P.: (1980b), `On Some Problems Related to the Extended Domain of Fourier Transform', Rocky Mount. Jour. of Math. 10, 99-104. Szlenk, W.: (1968), `The Non-Existence of Separable Reflexive Banach Space Universal for All Separable Reflexive Banach Spaces', Studia Math. 30, 53-61. Tichomirov, W. M.: (1960), `Diameters of Sets in Functional Spaces and the Theory of the Best Approximation' (in Russian) Usp. Matem. Nauk vyp. 3, 15 (93), 81-120. Tillman, H. G.: (1963), 'Randverteilungen analytischer Funktionen and Distributionen', Math. Zeitschr. 59, 61-83. Turpin, Ph.: (1966), `Sur une classe d'algebres topologiques', Comp. Rend. Acad. Paris 263, 436-439. Turpin, Ph.: (1972), `Measures vectoriels topologiques', Comp. Rend. Acad. Paris 275, 647-649.
Turpin, Ph.: (1973), `Un criterie de compacite dans les espaces vectoriels topologiques', Studia Math. 46, 141-148.
Turpin, Ph.: (1973b), 'Operateur lineaires entre espaces d'Orlicz non localement convexes', Studia Math. 46, 153-163.
Turpin, Ph.: (1973c), `Espaces et intersection d'espaces d'Orlicz non-localement convexes', Studia Math. 46, 167-195. Turpin, Ph.: (1975), `Convexites dans les espaces vectoriels topologiques generaux', Diss. Math. 131, PWN, Warszawa.
References
447
Turpin, Ph.: (1975b), Integration par rapport d une measure d valeurs dans un espace vectoriels non suppose localement convexe, Coll. sur 1'Integration vectoriels, Caen, Mai 1975.
Turpin, Ph.: (1975c), 'Une measure vectorielle non borne', Comp. Rend. Acad. Paris 280, 509-511. Turpin, Ph.: (1978), 'Fubini Inequalities and Bounded Multiplier Property in Generalized Modular Spaces', Comm. Math., Tomus specialis in honorem Ladislai Orlicz, 331-353.
Turpin, Ph.: (1978b), 'Properties of Orlicz-Pettis or Nikodym Type and Barrelledness Conditions', Ann. Inst. Fourier 28 (3), 67-85.
Turpin, Ph.: (198 ), 'The Range of Atomless Vector Measures, Comm. Math. 23. Turpin, Ph. and Waelbroeck, L.: (1968), 'Sur l'approximation des fonctions diffbrentiables it valeurs dans les espaces vectoriels topologiques', Comp. Rend. Acad. Paris 267, 94-97.
Turpin, Ph. and Waelbroeck, L.: (1968b), 'Integration et fonctions holomorphes dans les espaces localement pseudo-convexes', Comp. Rend. Acad. Paris 267, 160-162.
Turpin, Ph. and Waelbroeck, L.: (1968c), 'Algebres localement pseudo-convexes a inverse continue', Comp. Rend. Acad. Paris 267, 194-195.
Urbanik, K. and Woyczynski, W.: (1967), 'A Random Integral and Orlicz Spaces', Bull. Acad. Pol. Sci. 15, 162-169.
Vogt, D.: (1967), 'Integration Theorie in p-normierten Rumen', Math. Ann. 173, 219-232.
Vogt, D.: (1982), An Example of a Nuclear Frechet Space without the Bounded Approximation Property, preprint. Waelbroeck, L.: (1967), 'Some Theorems about Bounded Structure', Jour. Funct. Anal. 1, 392-408.
Waelbroeck, L.: (1967b), 'Differentiable Mappings into b-Spaces', Jour. Funct. Anal. 1, 409-418. Waelbroeck, L.: (1977),'A Rigid Topological Vector Space', Studia Math. 59,227-234
Whitley, R. J.: (1967), 'An Elementary Proof of the Eberlein-Smulian Theorem', Math. Ann. 172, 116-118. Wiliamson, J. H.: (1954), 'Compact Linear Operators in Linear Topological Spaces', Jour. London Math. Soc. 29, 149-156.
Wobst, R.: (1975), 'Isometrien in metrischen Vektorraumen', Studia Math. 54, 41-53.
Wojtyr ski, W.: (1969), 'On Conditional Bases in Non-Nuclear Frechet Spaces', Studia Math. 35, 77-96. Wood, G. V.: (1981), Maximal Symmetry in Banach Spaces, preprint. Woyczynski, W.: (1969), 'Sur la convergence des series dans les espaces du type (L)', Comp. Rend. Acad. Paris 268, 1254-1257.
Woyczynski, W.: (1970), 'Ind-Additive Functionals on Random Vectors', Diss. Math. 72, PWN, Warszawa.
448
References
Woyczyriski, W.: (1974), `Strong Laws of Large Numbers in Certain Linear Spaces', Ann. Inst. Fourier, 24, 205-223.
Woyczynski W.: (1975), Geometry and Martingales in Banach Spaces, Springer Lecture Notes 472, 235-283.
Zahariuta, V. P.: (1967) 'On Bases and Isomorphism of Analytic Functions on Convex Domains of Several Variables' (in Russian), Teoriya Funktsii i Funkts. Analiz 5, 5-12. Zahariuta, V. P.: (1967b), 'On Prolongable Bases in Spaces of Analytic Functions of One and Several Variables' (in Russian), Sib. Matem. Zhur. 7, 277-292. Zahariuta, V. P. : (1968),'On Quasi-Equivalence of Bases in Finite Centre of Hilbert Scales' (in Russian), Dokl. A.N. S.S.S.R. 180, 786-788. Zahariuta, V. P.: (1970), 'Spaces of Functions of one Variable, Analytic on Open Sets and on Compacts' (in Russian), Matem. Sb. 82, (124), 84-89. Zahariuta, V. P.: (1970b), 'On Isomorphism of Cartesian Product of Linear Topological Spaces of Functions' (in Russian), Funkts. Analiz, vyp. 2, 4, 87-88. Zahariuta, V. P.: (1970c), 'Propolongable Bases in Spaces of Analytic Functions on Multicircular Domains' (in Russian), Sib. Matem. Zhur. 11, 793-809.
Zahariuta, V. P.: (1973), 'On the Isomorphism of Cartesian Products of Locally convex Spaces' Studio Math. 46, 201-221. Zahariuta, V. P.: (1975),'On Isomorphism and Quasi-Equivalence of Bases in Power Kothe Spaces' (in Russian), Dokl. A.N. S.S.S.R 221, 772-774. Zahariuta, V.P.: (1976), 'On Isomorfism and Quasi-Equivalence of Bases in Power Kothe Spaces' (in Russian), Proceedings of the 7-th Winter School on Mathematical Programming and Related Problems, Drogobych (1974, ed. in Moscow, 1976), 101-126.
Zelazko, W.: (1960), 'On the Locally Bounded and m-Convex Topological Algebras', Studia Math. 19, 333-356. Zelazko, W.: (1965), 'Metric Generalizations of Banach Algebras', Diss. Math. 47, PWN, Warszawa. Zelazko, W.: (1972), 'A Power Series with a Finite Domain of Convergence', Comm. Math. 15, 115-117. Added in proof:
Dragilev, M.M.: (1983), Bases in Kothe spaces (in Russian), Rostov University Publications.
Subject Index
abscissa of convergence of Dirichlet series 373 absolute
- basis 292 - p-convex hull 298
basic sequence 67
- - represented by a matrix 333 - sequences equivalent 70 basis 67
- absolute 292, 329
absolutely
- block 72
- convergent series 315 - p-convex set 94 - summing operator 317 absorbing set 40 additive operator 36
- functionals 69
admissible pseudonorm 326 affine group property 392 algebra 173, 174
- Schauder 67
- commutative 174
- F* 174 - locally bounded 174 - semisimple 177 almost transitive norm 410 analytic function 124, 125 approximative dimension 264, 268
- - diametral 274 - needle point 241
balanced set 1 Banach space 96
- Hamel 76 - regular 335
- of the type dl 292 ,
----d2 293 - standard 74
- unconditional 329 ,8F* - space 52 block basis 72 Bochner-Lebesgue integral 123 Bo-space 93 Bo -space 93
bonuded
- approximation property 332 - measure 130 - multiplier convergent series 154
- operator 37 - sequence 37
- set 37
bases - equivalent 70
- quasi-equivalent 335 - semi-equivalent 333
Cauchy condition 18 Cauchy-Hadamard formula 376
Subject Index
450
Cauchy sequence 18 C-bounding point 222 centre of a Hilbert scale 340
-----finite 340
-----infinite 140 chain 46 characteristic function of a random variable 146 C-internal point 222 closed convex hull 223 cluster point
--of aset 33
- - of a family of sets 33 compact -measure 132
- operator 206, 288 complementary function to a convex
- transitive norm 415 co-universal space 63 C-sequence 165 C*-space 73 C-space 166
8-divergent zone 245
derivative of a vector valued function 198 diametral approximative dimension 274 Dirichlet series 371
- -entire 371 distribution of a random variable 146 domain
- of an operator 35 - proper of an integral operator 84 dominated convergence theorem 84
function 199 complete
- space 18 - set in a linear topological space 34 completion
- of a linear topological space 35 - of a metric space 21 complex rational numbers 28 condition
- Cauchy 18 - (JE) 10
-(A,) 11 -(1Q) 12 -(0) 173 - R 288 conjugate space 39, 199 consistent family of F-norms 46 continuous linear
- - functional 39 - - operator 39 continuously imbedded subspace 77 convergent sequence 2 convex
- function 111 - hull 221 - set 89, 221
element invertible 175 e-capacity 264 e-net 265 equicontinuous family of operators 39 equivalent - bases 70
- basic sequence 70
- F-norms 5 - metrics 2 extreme - point 238 - subset 238
finite
- centre of a Hilbert scale 340 - dimensional operator 206
- e-net 265 finitely supported function 46 F*-algebra 174 O-operator 294
F-norm 4 F-norms equivalent 5 F-pseudonorm 15
Subject Index F-space 22 F*-space 5 F*-space quotient 5 functional
451
identically distributed random variables 183 independent random
- - measures 145
- - variables 138
- continuous linear 39 - dimension 369
index 290
- linear 39
inequality
- Minkowski 188 - multiplicative linear 176 - non-trivial linear 75, 187 function
- analytic 124, 125 - complementary to a convex function 199 - convex 111
- finitely supported 46 - measurable 133 - Riemann integrable 120 functions
- equivalent on an interval (0, +oo) 112
- equivalent at infinity 112
--at0 112 fundamental
- family of sets 33 - sequence 18
F-closed set 226 F-closure 226 -compact set 226 F-continuous functional 226 F-topology 226 G. -set 20
Haar system 75 Hamel basis 76 Hermitian derivative 350 Hilbert scale 340 Hilbert-Schmidt operator 320 Hilbert space 22 homogeneous random measure 146
- Kolmogorov-Kchintchin 169
- Markov 346 - Paley-Zygmund 138 - Tchebyscheff 137
-Young 199 integral
- Bochner-Lebesque 123 - of a simple function with respect to a vector measure 134
- Riemann 120, 126 integration of a scalar valued function with respect to L'-bounded measure 143, 145
invariant metric 2 inverse 175 invertible element 175 isometry 390 isomorphic spaces 44
Kolmogorov-Kchintchin inequality 169 Kothe power spaces 285
- - - of finite type 285 - - - of infinite type 285 Levy-Kchintchin formula 147 linear
- codimension 62 - dimension 45 - functional 39
- operator 36 - space 1 - topological space 33, 223 locally
Subject Index
452
- bounded algebra 174
- - space 95
- compact space 250 - convex space 93, 225 - p-convex space 90 - pseudoconvex space 90, 93
multiplication 173 mulitplicative-linear functional 176
near isomorphism 290 needle point 241
- - approximative 241
- - space 250 Markov inequality 346 maximal norm 409 M-basis sequence 76 measurable function 133 measure 128 - bounded 130
-compact 132 - independent random 145 - Lm-bounded 135 - non-atomic vector 145
- separable 27 - or-finite 10
- space 10 - variation of 128 - vector valued 128 metric 1
- invariant 2 - linear space 1 - stronger than 2 metric equivalent 2 metrizable topological linear space 33 metrized modular space 10 metrizing modular 6 Minkowski functional 188 m-quasi-basis 76 modular 6
- metrizing 6 - space 10 modulus of concavity
- - - of a set 89 - - - of a space 96 monotone convergence theorem 83
- norm 48
Montel space 251
non-atomic vector measure 145 non-decreasing norm 7 non-trivial continuous linear functional 75, 187
norm 4 - almost transitive 410 - convex transitive 415
- equivalent to 5 - maximal 409 - monotone 48
- non-decreasing 7 - nuclear 309 - of a basis 103 - submultiplicative 174 - symmetric 421, 425 - stronger than 5 - transitive 410 normal sequence of subdivisions 120 normed space 96 nowhere dense set 18 nuclear
- norm 309 - operator 308
- space 296
operator - absolutely summing 317
- additive 36 - bounded 37
- compact 206, 288 - continuous linear 36 - finite dimensional 206 - Hilbert-Schmidt 320
- linear 36
Subject Index
- nuclear 308 Orlicz space 11
Paley-Zygmunt inequality 138 perfectly bounded sequence 173 p-homogeneous pseudonorm 90 plane of symmetry 420 point extremal 238 polynomials Tschebyscheff 137 precompact set 255 pre-Hilbert space 18 product space 6 proper domain of an integral operator 84 property P 135
453
Schwartz space 256 section 46 semi-equivalent bases 335 semisimple algebra 177 separable
- measure 27 - space 26 sequence
- bounded 37 - Cauchy 18 - convergent 2 - fundamental 18 - linearly independent 75 - linearly m-independent 76
- of independent random variables 138, 183
pseudoconvex set 89
- perfectly bounded 173 - Rademacher 138 - topologically linearly independent
pseudonorm 93
- admissible 326 - p-homogeneous 90
76
- weakly convergent 76, 130 quasi-norm 211, 252 quasi-equivalent bases 335 quotient
series
- absolutely convergent 315
- bounded multiplier convergent 154 - unconditionally convergent 152
- F*-space 5
- space 5
set
- absolutely p-convex 94
- absorbing 40 Rademacher sequence 138 random variable 136, 183
- - symmetric 147, 183
- balanced I
- bounded 37 - convex 89
- - identically distributed 183
- F-closed 226 - F-compact 226
- - independent 138, 183
- Gd 20
random variables
reflexive space 229
regular
- basis 333 - space 283
Riemann
- integrable function 120 - integral 120, 126 rigid space 210 rotation 390
- nowhere dense 18 - of the first category 18 - of the second category 18 - precompact 255 - pseudoconvex 89
- or-finite 10
- solid 77 - starlike 89 - starshaped 89
Subject Index
454
- totally bounded 255 - tree-like 46 - unfriendly 79 smooth point 401 solid
- set 77 - space 77
- LV(Q, --,.u) 12
- L°(n, E, u) 16,
-L(S1,I, u) 14
- metric 1 - metric linear 1 - metrizable 33 - metrized modular 10
- Banach 96
- modular 10 - Montel 251
- BB 93 - B° 93
- M 14
-,6F* 52
- M [a, b] 14 - M (am,n) 17
space
- complete metric 18
-m 14
-conjugate 39, 199
- M (S1, E, u) 14 - needle point 242
- c 15 - c° 15 - C(SI) 14
-N(l) 13
space complete metric linear 19
- C (S1/S1°) 15
- C [a, b] 15
- normed 96 - nuclear 296
- N (L) 13 - N (L[a, b]) 13
- C°°(S1) 16
- N(L (.Q, .E, µ)) 11
- e,(D) 16
- of Dirichlet series 371
- couniversal 63
- of the type dl 292 - of the type d8 293 - of the type d;, i = 1, 2 338 - Orlicz 11 - pre-Hilbert 18
-F 22
- F* 5
- having trivial dual 39
- Hilbert 22 - -)f (D) 354 - =l((D) 355 - C1 355 - Kothe power of finite type 285 - Kothe power of infinite type 285
- quotient 5 - reflexive 229
- regular 283
-rigid 210
- linear 1 - - topological 33, 223
- Schwartz 256 - second conjugate 229 - separable 26
- locally bounded 95
- solid 77
- - compact 250 - - convex 93, 225 - - p-convex 90
- strictly galbed 157
- - pseudoconvex 90, 93
- S(S1, E, µ) 12 - c5(En) 17
1P 13
- LP 12
-(s) 12 - S [a, b] 12
- LP [a, b] 12
- S(,) 371 - S(A) (R) 373
- LP am,n) 17
- S('A.) 375
Subject Index
- S""1 375
- a 340 - topological linear 33, 223
- universal 45 - - with respect to isometry 436
- - - - - isomorphism 45 - - - - - linear codimension 63
455
topology of bounded convergence 39 totally bounded set 255 total family of linear functionals 44, 195, 226
transitive norm 410 tree-like set 46 triangle inequality 1
- - - - - - dimension 45
- with arbitrarily short lines 196
- - bounded norms 389 - - non-trivial dual 187 - without arbitrarily short lines 52 - with strong Krein-Milman property 391
- with trivial dual 187 - (Xi)(.) 31 spaces
- isomorphic 44 - nearly isomorphic 290 sprctrum 181 standard basis 74 starlike set 89 starshaped set 89 subgroup
unconditional basis 329 unconditionally convergent series 152 unfriendly set 79 unit of na algebra 174 universal space 45
- - with respect to isometry 426
- - - -- - isomorhism 45 - - - - - linear codimension 63 - - - - - linavr dimension 45 variation of a vector valued measure 128 vector valued measure 128
- equicontinuous 391
- fat 391 submultiplicative norm 174 subspace 5 surjection 400 symmetric norm 421, 425 Tchebyscheff inequality 137
weakly convergent sequence 76, 230 weak - topology 226
- - of functionals 226 - *-topology 226
- polynomials 137 topological linear space 33, 223
Young inequality 199
Author Index
Alaoglu, L. 228, 434 Albinus, G. 95, 434 Antosik, P. 350, 434 Aoki, T. 95, 434 Arnold, L. 124, 434
Douady, A. 291, 436 Dragilev, M.M. 286, 287, 335--338, 436,
Aronszajn, N. 79, 80, 85-87, 434 Atkinson, F.V. 291, 434 Auerbach, H. 298, 408, 434
Dunford, N.S. 202, 445, 435, 437 Dubinsky, E. 238, 307, 331, 437 Duren, P.L. 193, 437 Dvoretzky, A. 316, 235, 437 Dynin, A. 313, 328, 330, 437
448
Drewnowski, L. 76, 77, 124, 133, 436, 437
Banach, S. 5, 39, 42, 43, 86, 100, 101, 168, 410, 427, 434, 435
Baire, R. 18 Bartle, R. 435 Beck, A. 435 Bessaga, C. 52, 65, 66, 107, 167, 168, 196, 237, 238, 252, 254, 274, 275, 277, 385, 389, 427, 428, 435
Bohnenblust, H.F. 191, 445 Bourgin, D.G. 252, 435 Burzyk, J. 307, 436
Charzyfiski, Z. 390, 436 Cowie, E.R. 416, 436 Crone, E.R. 335, 436
Day, M.M. 195, 436 Dieudonne, J. 252, 254, 436 Djakov, P.V. 331, 333, 436
Eberlain, W.F. 231, 437 Eidelheit, M. 205, 250, 385, 437 Egorov, D.F. 134
Fenske, Ch. 305, 437 Figiel, T. 238, 401, 438 Frechet, M. II, 438
Gawurin, M.K. 438 Gelfand, I.M. 176, 369, 438 Gohberg, I.C. 290, 438 Goldberg, A.A. 379, 382, 438 Goldstine, H.H. 229, 438 Gramsch, B. 124, 183, 438 Grothendieck, A. 317, 438 Gri nbaum, B. 438 Gurarij, W.I. 412, 439
Author Index Hahn, H. 189, 439 Henkin, M.G. 296, 365, 442, 443 Holsztyfiski, W. 439 Hyers, D.M. 211, 252, 439
Iyachen, S.O. 439
James, R.C. 234, 235, 439
Kakutani, S. 2, 439
Kalton, N.J. 46, 47, 51-53, 63, 76, 99, 193, 197, 199, 206, 210, 211, 213, 220, 248, 307, 419, 420, 439, 440 Klee, V. 19, 76, 197, 440
Klein, Ch. 120, 440 Kolmogorov, A.N. 96, 264, 440 Komura, T. 341, 440 Komura, Y. 341, 370, 371, 440 Kondakov, V.P. 333, 335, 336, 440 Krasnosielski, M.A. 202, 440 Krein, M.G. 103, 240, 290, 440 Krein, S.G. 340, 438, 440 Kwapiefi, S. 169, 440
457
Mazur, S. 21, 39, 40, 100, 101, 116--118, 121, 161, 168, 176, 202, 250, 251, 385, 390, 401, 427,442, 434, 435, 437 Metzler, R.C. 162, 442 Mikusifiski, J. 350, 434 Milman, D.P. 103, 240, 440 Mityagin, B.S. 274, 275, 277, 296, 313, 314, 328, 330, 331, 340, 354, 365, 436, 442, 443
Moscatelli, B. 307, 326, 443 Musial, K. 140, 443 Musielak, J. 6, 8, 443
Nakano, H. 6, 443 Ogrodzka, Z. 349, 443 Orlicz, W. 6, 8, 39, 40, 101, 116-118, 121, 153, 161, 169, 173, 202, 442, 443
Paley, R. 138, 443 Pallaschke, D. 206, 208, 443 Peck, N.T. 76, 197, 211, 213, 248, 439, 443
Pelczyfiski, A. 46, 52, 106, 107, 167, 168, 196, 237, 238, 268, 274, 275, 277, 321,
Labuda, I. 76, 437, 441 Landsberg, M. 94, 441 Lebesgue, H. 134 Leray, J. 224, 441 Levi, E.E. 182 Ligaud, J.P. 297, 298, 305, 441 Lindenstrauss, J. 321, 441 Lipecki, Z. 76, 441 Lorch, E.R. 441 Lusky, W. 412; 441 LuxemburgW.A.J. 77,79,82--84,441 Mankiewicz, P. 391--393, 441 Marcus, M. 186, 441 Matuszewska, W. 136, 442
330, 385, 389, 410, 412-415, 418-420,435,443,444 Petrov, V.V. 184, 186, 444 Pettis, B.J. 444 Phelps, R.R. 401, 444 Pietsch; A. 321, 324, 444 Pisier, G. 136, 442 Popov, M.M. 63, 444 Prekopa, A. 146, 444 Przeworska-Rolewicz, D. 124, 183, 290, 444
Raikov, D.A. 444 Ramanujan, M.S. 307, 437, 444 Retheford, J.R. 435
Author Index
458
Ritt, J.F. 384, 444 Roberts, J.W. 210, 211, 220, 241, 242,
Szeptycki, P. 77, 79, 80, 85--87, 446 Szlenk, W. 446
245, 246, 248, 439, 444
Robinson, W.B. 307, 335, 436, 437 Rogers, C.A. 316, 325, 437 Rolewicz, S. 14, 52, 89, 95, 96, 106, 108, 113, 120, 124, 135, 155, 183, 193, 194, 196,
198, 238, 252, 254, 260--262,
274, 275, 277, 290, 355, 376, 385, 389,
Talagrand, M. 131 Tichomirov, W.M. 274, 440, 446 Tillman H.G. 446 Turpin, Ph. 64, 88, 126, 127, 131, 135, 157,16392789280-283,446,447
410, 412--415, 418--420, 435, 440, 444,445
Romberg, R.G. 193, 437 Rosenberger, B. 236, 445 Rutitski, Ja.B. 202, 440 Rutman, L.A. 103, 440 Ryll-Nardzewski, C. 135, 140, 141, 155,
Ulam, S. 390, 442 Urbanik, K. 150, 447 Vilenkin, N.Ja. 369, 438 Vogt, D. 332, 447
186, 254, 443, 445
Schauder J. 67, 445 Schields, A.L. 193, 437 Schock, E. 305, 437 Schwartz, J. 202, 435, 437 Schwartz, L. 169, 171-173, 445 Semadeni, Z. 238, 444 Shapiro, J.H. 193, 197, 206, 248, 439, 445 Sierpinski, W. 20, 446 Simmons, S. 28, 446 Singer, I. 330, 421, 444, 446 Slowikowski, W. 258, 446 9mulian, V. 231, 446 Sobczyk, A. 191, 435 Srinivasan, V.K. 376 Steinhaus, H. 39, 435 Sternbach, L. 21, 442 Stiles, W.J. 101, 103, 446 Sundaresan, K. 184
Waelbroeck, L. 126, 127, 210, 447 Whitley, R.J. 231, 447 Wiener, N. 181 Wiliamson, J.H. 291, 447 Wobst R., 390, 447 Wojtyriski, W. 330, 340, 447
Wood, G.V. 417-421, 447 Woyczynski, W. 140, 141, 150, 152, 184, 186,440,441,443,445--448
Zaanen, A.C. 82--84, 441 Zahariuta, V.P. 288, 291, 295, 338, 340, 365, 368, 448
Zobin, N.M. 331, 443 Zygmund, A. 138, 443 2e1azko, W. 90, 124, 174, 176, 177, 181, 182, 448
List of Symbols
A+B 1 p (x, y) I
M(E) 128 E(X) 137
(m1),...,(m3) 1 IA I
V(X) 137 A(X) 157
(X, p) 1
conv(A) 221
x.--->x 2
conv(A) 223
P
xn --> x 2
n(X) 229
Ilxll 4 (n 1), ... , (n 6) 4 (X II II)5 X/Y 5 (md 1), ... , (md 5) 6
X* * 229 E(K) 238
(i 1), ... , (i 5) 17
M(A, B, e) 263 M(A, B) 264 M(X) 264 M'(X) 268 M;(X) 268 M(X) 270 6 (A, B, L) 274
G8 20
6n(A, B) 274
(X3(,) 31, 100
6(X) 274 a(X) 275 6'(A, B) 284 6'(X) 284 6'(X) 284 T(X) 295
X, 6
(md 5') 6 (x, y) 17
DA 36
Y) 38 Be(X B0(X) 38
B,(X - Y) 39 Y) 39 X* 38, 199 dimiX 45 E [c] 46 B (X -
dn(T) 308
a(T) 318
suppx 47
Ct,t 327
codimiX 62
DK 84 c(A) 89 c(X) 96
d; 338 H(X) 391 T(X) 391 Lin (X) 392 Aff(X) 392 Inv(G) 391
n(t) 107
G(II
c , 77 E.'\, o 81 X, 81
ID 408