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!
. Then there exists fo E L1 (JR) such that
onto R We still have to verify that the Gel'fand topology of coincides with the usual topology on R For this, it is sufficient to show that the mapping C such that d(ab) c ----+ Spa, is a homeomorphism, and so, identifying 1>c with Sp a, we see that the Gel'fand representation of C is an isometric A). Define W
V(n, r) = {cpy E : sup le- iyt Itl::on
11 < r}
.
We claim that these sets V (n, r) form a base of neighbourhoods of CPo in the Gel'fand topology of . Clearly, each set V(n, r) is an open neighbourhood of CPo.
202
Introduction to Banach Spaces and Algebras
Now let N = {!py E : l!py(f)) -!Po(f))1 < c (j = 1, ... ,m)} be a basic open neighbourhood of !Po in , where il, ... , 1m E Ll(JR) and c > o. We may suppose that il, ... , 1m E Coo(JR). Take n E Nand r > 0 such that supp IJ <:;;; [-n, n] and r III) III < c for j = 1, ... , m. Then it is immediately checked that V(n, r) <:;;; N, giving the claim. But now, for each n E Nand r > 0, we see that !Py E V(n, r) if and only if IYI < (2/n)sin-l(r/2), and so the two topologies do coincide. 0 We see that the Gel'fand and Fourier transforms of Ll (JR) again coincide when we have made the identifications of the above theorem. 4.13 Boundaries and peak sets. Let A be a natural Banach function algebra on a compact space K. A subset E of K is a peak set for A if there exists I E A such that
I(x)
=
1
(x
E
E)
and
II(x)1 < 1 (x
E
K \ E);
the function I peaks on E. A point Xo E K such that {xo} is a peak set is a peak point, and the set of peak points is denoted by ro(A). A subset E of K is a boundary for A if, for each I E A, there exists x E E with II(x)1 = III K ; a closed boundary is a closed subspace of K that is a boundary. Clearly, each boundary contains ro(A). For I E A, define the maximum set of I by
M(f) = {x
E
K: II(x)1 = III K
}·
We shall show that, in the general case, the intersection rcA) of all the closed boundaries for A is itself a boundary for A (and so rCA) 1= 0). Let x E K. Clearly, x E rCA) if and only if, for each neighbourhood U of x, there exists I E A with M(f) <:;;; U, and so rCA) = rCA), where 11 is the uniform closure of A in C(K). We first give some examples. Examples 4.76 (i) Let (K; d) be a non-empty, compact, metric space, and let A = C(K). For Xo E K, the function Y f---7 1/(1 + d(xo, y)) belongs to A and peaks at xo, and so ro(A) = rCA) = K. (ii) Let A = A(~) be the disc algebra, as in Example 4.3. The maximum modulus theorem, Proposition 1.35, shows that '][' is a closed boundary for A. However, take Zo E ']['. Then the function (Z + zol)/2zo E A peaks at zo, and so every point of'][' is a peak point. Thus clearly ro(A) = rcA) = ']['. (iii) Let K be a non-empty, compact subset of C, and let A = R(K), as in Example 4.51. By the maximum modulus theorem, 8K is a closed boundary for A. Take Zo E 8K, and let U be a neighbourhood of Zo in
Banach algebras
203
Lemma 4.77 Let A be a natural Banach function algebra on a non-empty, compact Hausdorff space K, and take 11, ... , fn EA. Set
v = V(J"
.. ,fn) = {x
E
K : IfJ(x)1 < 1 (j = 1, ... , n)}.
Then either V n E =I- 0 for each boundary E, or else F \ V is a closed boundary for A whenever F is a closed boundary for A. Proof Suppose that there is a closed boundary F such that the closed set F\ V is not a boundary. We must show that V n E =I- 0 for each boundary E. There exists f E A with If IF = IflK = 1, but with IflF\v < 1. By replacing f by a suitably high power of itself, we may suppose that
If fJ IF\V < 1
(j = 1, ... , n) .
For each j = 1, ... , n, we have If fJ (x) I < 1 (x E V), and hence If fJ IF < 1. Thus IffJIK < 1 because F is a boundary. Take Xo E M(f), so that If(xo)1 = 1. Then IfJ(xo)1 < 1 (j = 1, ... , n), and so Xo E V. Hence M(f) <:;; V, and so VnE =I- 0, as required. 0 Theorem 4.78 Let A be a natural Banach function algebra on a compact space K. Then r(A) is a closed boundary for A. Proof Certainly f(A) is a closed subset of K (at this stage, possibly empty). Assume towards a contradiction that r(A) is not a boundary. Then there exists f E A with IflK = 1 and M(f) n r(A) = 0. Let y E M(f). Then there is a basic open neighbourhood, say of the form V = V(J, ,... ,fn)' of y and a closed boundary E for A with V n E = 0. By Lemma 4.77, F \ V is a closed boundary whenever F is a closed boundary. However, these sets V together form an open cover of the compact space M(f); let {VI, ... , Vn } be a finite subcover of this open cover. Certainly K is a closed boundary; hence K \ VI is a closed boundary, and, repeating the argument n - 1 times, K \ (VI U· .. U Vn ) is a closed boundary. Since M(f) <:;; VI U ... U Vn , this is a contradiction. Thus r( A) is a boundary. 0
The set r(A) is called the Shilov boundary of A. We now wish to prove that ro(A) is a boundary, and hence dense in r(A), whenever K is metrizable and A is a uniform algebra. Lemma 4.79 Let A be a natural uniform algebra on a compact space K. Suppose that f,g E A are such that: (i) IflK = 1; (ii) IgIM(J) = 1; (iii) L =I- 0, where
L = {x Then there exists h
E
E
K : g(x)
A such that M (h)
= f(x) = I}.
= L.
204
Introduction to Banach Spaces and Algebras
Proof We may suppose that f(x) = 1 (x E M(f)), for otherwise we replace f by (1 + f)/2 E A without changing anything else. Set m = IgIK' so that m 2 1. If m = 1, set h = (1 + g)f /2, and then h E A and M(h) = L, as required Now suppose that m > 1. For n E N, set
Kn = {x E K: 1 + Tn(m - 1) :::; Ig(x)1 :::; 1 + Tn+l(m - I)}. Each Kn is closed (possibly empty) in K and, by hypothesis (ii), If(x)1 any x E Kn. Further,
U{Kn : n E N}
=
< 1 for
{x E K: Ig(x)1 > I}.
For each n E N, there exists Pn E N with IfP" (x)1
< 1/2 for any x
E Kn-
Define
=
h = g + 4(m - 1) L T n f P"
;
n=l
we note that the series converges in A because IfP" IK We now consider two cases for a point x E K. First, suppose that Ig(x)1 :::; 1. Then
Ih(x)l:::; Ig(x)1
+ 4(m -
= 1 (n
EN).
=
1) LTn If(xW" :::; 1 + 4(m - 1). n=l
This inequality is strict unless Ig(x)1 = If(x)1 = 1. But f(x) = 1 when If(x)1 = 1, and so the inequality is strict unless also g(x) = 1. Thus, for such a point x, we have Ih(x)1 = 1 + 4(m - 1) if and only if x E L. Secondly, suppose that Ig(x)1 > 1. Then x E Kr for some r E N, and so Ig(x)1 :::; 1 + 2- r + 1 (m - 1) and If(xW' < 1/2, which gives
Ih(x)1 :::; Ig(x)1
+ 4(m -
1) (Tr If(xW'
+ LTn) nfr
=
Ig(x)1
+ 4(m -
< 1 + Tr+l(m -
+ 1 - Tr) 1) + 4(m - 1)(T r - 1 + 1 - Tr)
1) (Tr If(xW'
=
1 + 4(m - 1).
We conclude that IhlK = 1 + 4(m - 1) and M(h) = L, as required.
0
Theorem 4.80 Let A be a uniform algebra on a compact, metrizable space. Then the set ro(A) of peak points is a boundary for A, and ro(A) = r(A). Proof We suppose that A is a uniform algebra on K, a compact, metrizable space. Let fo E A. We must show that M(fo) contains a peak point. Take J to be the family of all maximum sets contained in M (fo) (so that J -I- 0), and order J by inclusion, so obtaining a partially ordered set. Let It
Banach algebras
205
be a chain in J, and set M = n{N : N E It}. Certainly M =I- 0 because K is compact. Since K is metrizable, M is in fact the intersection of countably many sets in It, say M = n{M(h n ) : n EN}, where Ihnl K = 1 (n EN). Choose Yo E M. By multiplying each h n by a suitable constant of modulus 1, we may suppose that hn(yo) = 1. Set h = L~=lhn/2n, so that h E A and h(yo) = 1 = IhIK. Since Ihn(y)1 < 1 for y E K \ M(h n ), we see that M(h) <;;; M, and so M(h) is a lower bound for It in (J; <;;;). By Zorn's lemma, bis, J has a minimum element, say M(J), where f E A. We may suppose that IflK = 1. Assume towards a contradiction that M(J) contains more than one point. Then there exists 9 E A that is not constant on M(J); again we may suppose that IgIM(f) = 1. By multiplying f and 9 by constants of modulus 1, we may suppose that f(x) = g(x) = 1 for some x E M(J). Thus L <;; M(J), where L is as in Lemma 4.79. However, by Lemma 4.79, L E J, a contradiction of the minimality of M(J). Hence M(J) is a singleton, say M(J) = {xo}. Clearly, Xo E M(Jo) n fo(A), and so f 0 (A) is a boundary for A. Certainly fo(A) <;;; f(A). But fo(A) is a closed boundary for A, and so fo(A) ;;2 f(A). Thus fo(A) = f(A). D Let A be a unital Banach algebra. Then we set
KA = {'\ E A*:
PII = '\(1) =
I},
so that KA is a closed, convex subset of (A*)[lJ' and hence KA is compact in (A*;O"(A*,A)). By Theorem 4.43, A <;;; KA. In particular, let A be a uniform algebra on a compact, Hausdorff space K. Then we regard K as a closed subset of KA, In the next theorem, closures are taken in the weak-* topology.
Theorem 4.81 Let A be be a uniform algebra on a compact space K. Then conv(fo(A)) = KA and fo(A) <;;; exKA <;;; f(A). Proof Clearly, conv(fo(A)) <;;; KA; we shall first prove the converse in the special case where A = C(K). Set E = ClR(K), a real-linear space. Take'\ E C(K)* with 11,\11 = '\(1) = 1. We borrow a result from Lemma 6.20 that shows that '\(J) E lR for all fEE, and hence that ,\ I E is a real-linear functional; clearly,\ lEE KE = {JL E e* : IIJLII = JL(1) = I}. Assume towards a contradiction that there exists'\o E KE with'\o conv(K). By Theorem 3.26(iii), there exists A E (E*; O"(E*, E))* = E such that A('\o) > 1 and SUPxEK A(cx) ::::: 1, and so there exists fEE with
rt
'\0(J) > 1
and
f(x)::::: 1
(x E K).
By replacing f by (J + a . 1)/(1 + a) for suitable a > 0, we may suppose that f E E+, and so IflK ::::: 1. But now 1'\0(J)1 : : : 1, a contradiction. Thus
conv(K)
=
K E.
We now return to the general case. The map f f--+ f is an isometry, and so we may suppose that K = rcA).
I f(A),
A ---- C(f(A»,
206
Introduction to Banach Spaces and Algebras
Let Ao
E
K A, and take a basic neighbourhood U of 0 in A *, say U = {A
E
A* : IA(f,)1 < 1 (i = 1, ... , n)},
where II, ... , fn E A. Set V = {A E C(K)* : IA(f,)1 < 1 (i = 1, ... , n)}. There is an extension flo E C(K)* of AO with Ilfloll = flo(1) = 1. By the above result, there is III E cony K with fll - flo E V. Set Al = fll I A. Then Al E cony K and Al - Ao E U. This holds for each such U, and so Ao E conv(K). By Theorem 4.80, ro(A) = r(A), and so conv(ro(A)) = KA. Now take Xo E ro(A). Then there exists f E A with f(xo) = 1 and If(x)1 < 1 for each x E K with x of Xo. Set Ko = {A E KA : A(f) = I}. Then Ko is a nonempty, closed extreme subset of KA, and so, as in the Krein-Milman theorem, Theorem 3.31, Ko contains an extreme point of K A . Since ex KA <;:; K, there exists x E K with x E exKA n Ko. Thus f(x) = 1, and so x = Xo. Hence Xo E exKA. We have shown that ro(A) <;:; exKA. D Notes A more general version of Theorem 4.46 states that, for each algebra A, the map 'P f--+ ker'P is a bijection from CPA onto the set of maximal modular ideals of co dimension 1 in A [47, Proposition 1.3.37]. A maximal ideal in even a commutative, unital algebra does not necessarily have co dimension 1. For example, there are many examples of fields_ven ordered fields [47, Definition 1.3.61]-of large dimension, and {OJ is a maximal ideal in these fields. These 'large' fields do have an important role in Banach algebra theory; see [47, 51]. Indeed, there are 'very large' algebras analogous to algebras of formal power series [51]. We have defined a 'Banach algebra of power series' (A; 11·11) within Example 4.52; the definition requires that the maps 7l"m : (A; 11·11) ---> C are all continuous. In fact, it is a recent result [55] that this time-honoured phrase is redundant. For a discussion of Banach algebras of power series, including a determination of their family of closed ideals in certain cases, see [47, Section 4.6]. We have defined the product in the Banach algebras Ll(1r) and Ll(JR) without using the theory of Lebesgue integration. Take G to be 1r or JR. The formulae given also define f * 9 directly for f,g E Ll(G). For this, one must show using Fubini's theorem that f * 9 E Ll(G) (and hence that IU * g)(x)1 < (Xl almost everywhere) for each f,g E Ll(G). Care must be taken over some measure-theoretic points. See [143, Section 7.13], for example. For f ELI (JR +) with f =1= 0, define aU)
= sup{ 8 ~ 0 : f I [0,8] = 0 almost everywhere} = inf supp f .
In Example 4.68, we essentially used the obvious fact that aU * g) ~ aU) + a(g) for f,g E L 1 (JR+) with f,g =1= O. In fact, it is a deep theorem of Titchmarsh (see [47, Theorem 4.7.22], for example) that aU * g) = aU) + a(g) for f,g E L1(IR+) with f,g =1= O. This implies that (Ll(IR+,W); *) is an integral domain for each continuous weight function w on 1R+. Let G be a locally compact group, so that G is a group with a locally compact topology and the group operations are continuous with respect to the topology. For a E G and f on G, define (Saf)(S) = f(a-1s) (s E G). Then there is a so-called left Haar measure m on G such that m =1= 0 and fG Saf dm = fG f dm for all a E G and f E Ll(G) = Ll(G, dm). For example, the Haar measure on (IR; +) is just the
Banach algebras
207
Lebesgue measure. The convolution product of formula
(f
* g)(s) =
fa
I, gEL 1 (G)
l(t)g(C l s) dm(t)
is now defined by the
(s E G).
In this case, (L 1(G); II . 111 ; *) is a Banach algebra; it is commutative if and only if G is abelian. It is shown in [47, Corollary 3.3.35] that Ll(G) is always semisimple. Next, suppose that G is abelian, so that G is an LeA group. A character on G is a group morphism from G to 1I'. The set of all continuous characters on G forms a group with respect to the pointwise product; this is called the dual group of G, and it is often denoted by r = G. The dual group r is itself an LCA group for a suitable topology, and the famous Pontryagin duality theorem states that the dual group of r can be identified with G. For example, it is implicit in our work that Z = 1I', that 'if = Z, and that i. is a 'second copy' of R The Fourier transform of 1 E L1(G) is the function on r defined by
1(r) =
fa 1
(s)')'( -s) dm(s)
(r
E
r, 1 E
L1(G)).
1
The (generalized) Riemann-Lebesgue lemma states that E Co(r), and then the Fourier transform is the map :F : 1 f--> Ll(G) -> Co(r). As in each of the specific examples that we have considered, this map is a monomorphism and the character space of Ll(G) can be identified with r in the sense that each character on Ll(G) has the form 1 f--> i( 'Y) for some (unique) 'Y E r. The identification of the character space of Ll(G) with r is a homeomorphism, and, with this identification, the Gel'fand transform coincides with the Fourier transform. The details of the above are given in various texts, including [47, 84, 94, 106, 145]. Our approach to the specific examples is rather close to that in [82]. In general, it is not true that E Ll(r) when 1 E L1(G). However, if E L 1 (r), there is an 'inversion formula' that recovers 1 from j Let K be a non-empty, compact subset of the plane. We have remarked that it is very easy to see that the uniform algebra R(K) is natural on K. It is true, but harder to prove, that the uniform algebra A(K) is also natural on K; this is a theorem of Arens, proved in [47, Theorem 4.3.14] and [79, Chapter II, Theorem 1.9]. There is a substantial study of when any two of the algebras R(K), A(K), and C(K) are equal in [79, Chapter VIII]; these questions encompass a considerable amount of classical approximation theory. Mergelyan's theorem [79, Chapter II, Theorem 9.1] asserts that P(K) = A(K) whenever e \ K is connected, and the Hartogs-Rosenthal theorem [79, Chapter II, Corollary 8.4] asserts that R(K) = C(K) whenever K has plane measure zero. Indeed, it is a theorem of Bishop [79, Chapter II, Theorem 11.4] that R(K) = C(K) if and only if the set of points that are not peak points has plane measure zero. An example for which R(K) i- A(K) is given in Exercise 4.16, below; criteria involving the 'capacity' of sets for R(K) = C(K) and for R(K) = A(K), and various exotic examples, are given in [79, Chapter VIII] and [152]. Conditions for a point in 8K to be a peak point for R(K) or A(K) are given also in [79, Chapter VIII]. There is one obvious example of a natural uniform algebra on the closed unit interval I, namely C(I). It is a conjecture that goes back to Gel'fand that this is the only natural uniform algebra on I. Let A be such an algebra. We do not even know if every point of 1 must be a peak point for A; however, we shall see in Corollary 9.12 that necessarily r(A) = I. Further, we do not know whether or not necessarily A = C(I) if we assume that ro(A) = I. This question is related to that of polynomial approximation on curves in en; see [152, Section 30J. Here is a related tantalizing open question of considerable
1,
1
1
208
Introduction to Banach Spaces and Algebras
antiquity. Let A be a natural uniform algebra on the closed unit disc ~. Does the Shilov boundary r (A) necessarily meet (or contain) the topological boundary '][' ? The proof in Theorem 4.80 is essentially due to Bishop. In the non-metrizable case, one can replace 'peak point' by 'strong boundary point'; the set of these points is dense in r(A). Let A be be a natural uniform algebra on a compact, Hausdorff space K. Then we showed in Theorem 4.81 that ro(A) <;;; ex KA. In fact, ex KA is equal to the set of strong boundary points for A (and hence to ro(A) in the case where K is metrizable). This set ex KA is called the Choquet boundary for A; it is characterized in several different ways in [47, Theorem 4.35]. A famous theorem of Choquet is relevant here. It states the following. Let K be a metrizable, compact, convex subset of a locally convex space E, and let Xo E K. Then there is a probability measure IL on K such that f(xo) = K fdlL (f E E*) and IL is supported on exK. For a natural Banach function algebra A on a compact, metrizable space K, the set ro(A) is not necessarily a boundary for A [43], but it is still true that ro(A) = r(A) [47, Corollary 4.3.7]. The so-called peak-point conjecture was outstanding for a long period. Let K be a compact, metrizable space. This conjecture stated that the only natural uniform algebra A for which ro(A) = K is A = C(K). In other words, for each proper, natural uniform algebra on K, there exists x E K that is not a peak point for A. This conjecture was shown to be false by Cole in his thesis [41]; see also [152]. An explicit counter example was given by Basener [26] a little later; it is the uniform algebra R(K) for a certain compact subset K of the unit sphere in C 2 • See [152, Example 19.8]; for related results of Feinstein, see [76]. We have given various standard results for commutative Banach algebras. Some, but not all, extend to Fn§chet algebras and other topological algebras. For example, there is an appropriate Gel'fand representation theorem for complete, commutative LMC algebras. However, there are examples that show that certain statements do not carry through; see [47, Section 4.10].
J
Exercise 4.11 (i) Let K be a non-empty, compact Hausdorff space, and let F be a closed subset of K. Set
J(F) = {J E C(K) : f I F = O} . Show that J(F) is a closed ideal in C(K), and that every closed ideal in C(K) has this form. (ii) Let K and L be non-empty, compact Hausdorff spaces. Prove that the Banach algebras C( K) and C( L) are isomorphic as algebras if and only if the topological spaces K and L are homeomorphic. Exercise 4.12 (i) Take kEN, and let A = (Ck(H); II· Ilk) be the Banach space introduced in Exercise 2.10. (See also Exercise 2.19.) Show that A is a natural Banach function algebra on H with respect to the pointwise multiplication of functions. (ii) Let A = CCXl(H) = n{Ck(H) : kEN} be the space of infinitely differentiable functions on H. Show that A is a Fn§chet algebra with respect to the sequence (II '1I kh2':l of seminorms on A and pointwise multiplication of functions. Show that A is functionally continuous, and that all characters on A are given by evaluation at a point of H.
209
Banach algebras
Exercise 4.13 Let (K; d) be a compact metric space, and take ex with 0 < ex S 1. Then let A = Lip"K be the Banach space of Lipschitz functions on K; this space was introduced in Exercise 2.13. Show that A is a natural, self-adjoint Banach function algebra on K with respect to the pointwise multiplication of functions. Exercise 4.14 (i) Let bv be the space of sequences x = (Xn)n2:1 in co(N) such that 00
Ilxll bv = Ilxli oo
+L
IXn+1 - xnl <
00.
n=l
Show that (bv; 11·ll bv ) is a Banach sequence algebra on N. (ii) We defined the James space (J;N) in Exercise 2.14. Let x,y E J. Show that xy E J with N(xy) S 2N(x)N(y), and hence conclude that J is a Banach sequence algebra on N for a norm equivalent to N. Exercise 4.15 Let U be a non-empty, open subset of the complex plane IC, and let (HOO(U); 1·l u ) be the Banach space that was introduced in Exercise 2.12. Prove that this space is a Banach algebra with respect to the pointwise multiplication of functions. This Banach algebra is studied extensively in the monograph of Garnett [80]. Let Ez : J f-+ J(z) be the evaluation character on HOO(U) for z E U, and let
+ ... + IJn(z)1
: z E U} > 0,
there exist gl, ... ,gn E H oo (U) such that 2:: 1J,g, = 1. The fact that this condition is indeed satisfied in the case where U is the open unit disc in IC is the famous corona theorem of Carleson [37]; for an exposition of this theorem, based on an approach of Wolff, see [80, Chapter VIII, Section 2], and, for a simpler proof, see [28]. Exercise 4.16 A Swiss cheese is a non-empty, compact subset of IC obtained from ~, the closed unit disc, by deleting a sequence (Dn)n2:1 of open discs, with each Dn contained in int~; we also require that Dm n Dn = 0 whenever m, n E N with m =1= n and that 2:~=1 Tn < 1, where Tn is the radius of Dn. We then set
K
=
~
\
U{ Dn : n E N} ,
so that K is a compact subset of ~. Give the details to show that, starting with a countable, dense subset S of int~, an inductive construction gives a sequence (Dn)n2:1 satisfying the above conditions, with the centre of each Dn in S, and so that K has empty interior. It follows that A(K) = C(K) in this case. However, we claim that R(K) =1= C(K). To see this, consider the continuous linear functional
A:J
f-+
1
8~
J(z) dz -
f= 1
J(z) dz
n=l 8Dn
on C(K). Verify that A is indeed an element of C(K)* and that (j, A) = 0 for each rational function J with poles off K, and hence that A I R(K) = O. However, show that
210
Introduction to Banach Spaces and Algebras
(z, A) This shows that
Z
=
27ri (1 - ~ rn) =I- o.
~ R(K), and so R(K)
=I- C(K).
Exercise 4.17 (i) We have discussed the algebra ~ = C[[X]] of formal power series in one variable in Example 4.52. This algebra is an integral domain, and so we know from standard courses in algebra that it has a quotient field. Verify that this quotient field can be identified with the field q(X)) of Laurent series in one variable; the latter is the algebra of formal series of the form L~=no anX n , where no E Z and each an E
(ii) Let IC{ X} be the collection of absolutely convergent power series, so that IC{ X} consists of the formal power series L~=o anXn E ~ such that L~=o Ian Ien < 00 for some 10 > O. Verify that IC{ X} is a subalgebra of ~ and that its quotient field, the field qX) of meromorphic functions, is a subfield of q (X)). Is there a norm II . II such that (IC{X}; 11·11) (respectively, (qX); 11·11)) is a normed algebra? Exercise 4.18 Let w be a weight function on Z+ , and let A = eI (Z+ ,w) be the Banach algebra of power series, as defined in Example 4.52. Set P = PA(X) = limn-->oo w(n)l/n. Show that SPA X = {Z E IC : Izi ::; p} = iPA, say SPA X = K. Show that A <;;; P(K) = A(K) and that A is semisimple whenever p> O. Exercise 4.19 For t E (0,1), set f(t) = 1/0 and get) = 1/yff"=t, and, for t:O:: 1 and t ::; 0, set f(t) = get) = O. Show that f and g are Riemann-integrable on JR, but that the integral for (f * g)(1) does not converge. Exercise 4.20 Let A be a Banach algebra. A (sequential) approximate identity in A is a sequence (a n )n2:1 in A such that aa n --> a and ana --> a as n --> 00 for each a E A. Show that each 'approximate identity' as defined in Exercise 2.36 is an approximate identity in the Banach algebra (L I ('f); II . III ; * ). Exercise 4.21 Let n E N. Compute explicitly the function hn := XI-n,n] Show that h n = fn (n EN), where
fn(t) = sin tsin(nt) 4t 2
Show that
Ilfnlll
--> 00
as n
--> 00,
but that IhnllR
*
XI-I,I]'
(n E N).
= 1 (n EN). Deduce from this that
A(JR) =I- Co(JR). Exercise 4.22 Let E be a closed subspace of the convolution algebra (LI(JR); *). Prove that E is translation-invariant if and only if E is an ideal.
211
Banach algebras
Exercise 4.23 Set IT = {z = x + iy E C : x > o}. For fELl (JR+) (see Example 4.67), we now define the Laplace transform £(f) by
£(f)(z) =
1
00
f(t)e- zt dt
(z E IT).
Show that £(f) E Co(IT) and that £(f) is analytic on IT. Show that
£ : (LI(JR+); *)
---->
(Co (IT); .)
is a monomorphism, and that the character space of LI(JR+) can be identified with IT in such a way that each character on L1(JR+) has the form f f---> £(f)(z) for some ZEIT. Exercise 4.24 Let w be a continuous weight function on JR+. Show that the limit p = limt~oo W(t)l/t exists. Let LI(JR+,W) be the commutative Banach algebra defined in Example 4.67. Show that LI(JR+,W) is semisimple whenever p > 0 and radical whenever p = o. Exercise 4.25 Let A be a natural uniform algebra on a metrizable, compact space K, and let Xo E K. Show that Xo is a peak point for A if and only if there exist a,j3 with o < a < 13 < 1 such that, for each neighbourhood U of xo, there exists f E A with IflK ::; 1, with f(xo) > 13, and with If(y)1 < a (y E K \ U). Exercise 4.26 Let A be a natural uniform algebra on a compact set K, and let Show that 8Sp(f) ~ f(r(A)).
f
E
A.
Exercise 4.27 Let K and L be the closed discs in C, each ofradius 1, and with centres at (0,1) and (0, -1), respectively, so that K n L = {(O,O)} and P(K U L) is natural on K U L. Show that (0,0) is a peak point for P(K U L), but that there is no polynomial that peaks at (0,0). Exercise 4.28 Let K = {(z, t) E C x JR : Izl ::; 1, 0::; t ::; I}, and let A be the set of functions in C(K) such that the function z f---> f(z,O) is analytic on int~. Show that A is a natural uniform algebra on K with r(A) = K. Identify the peak points for A, and note that they form a proper subset of K.
Runge's theorem and the holomorphic functional calculus 4.14 Runge's theorem. Let K be a non-empty, compact subset of Co We first recall the notations P(K), R(K), and A(K) from Example 4.3. These are uniform algebras on K, and R(K) and A(K) are natural. For a E
212
Introduction to Banach Spaces and Algebras
Proof (i) Set B = R(K)(Z, U/3), a closed subalgebra of A = R(K), and take U to be the component of e \ K to which a and f3 belong. Since K = SpAZ, the set U is a component of e \ SPAZ. But u/3 E B, so that f3 ~ SpBZ, and then, by Corollary 4.37(i), Un Sp BZ = 0. Thus a ~ SpBZ, i.e. Un E B. (ii) By Corollary 4.37(ii), the unbounded component of e \ K does not meet SpcZ for any closed subalgebra C of A that contains Z. In particular, this applies with C = P(K), and therefore Un E P(K) for every a in this unbounded component. 0
Let K be a non-empty, compact subset of C. Then we denote by OK the algebra consisting of all equivalence classes of functions analytic on some neighbourhood of K. [To be precise, for f, 9 E OK, we set f rv 9 to define the equivalence relation if there is an open neighbourhood W of K with f I W = 9 I W E O(W); further, if f E O(U) and 9 E O(V), where U and V are open neighbourhoods of K, we define f + 9 and fg pointwise on Un V; the algebraic operations are compatible with the equivalence relation, and so the family of equivalence classes is an algebra.] The elements of OK are called germs of analytic functions. There is a way of defining a topology on the algebra OK by regarding it as the inductive limit of the Fn§chet algebras O(U) as U runs through the family of open neighbourhoods of K. For us, it is sufficient to say that a sequence (fn)n>l in OK converges to 0 if there is an open neighbourhood U of K such that each fn is in O(U) and limn---+<Xl fn = 0 on a compact neighbourhood of K in U. Theorem 4.83 (Runge's theorem) Let K be a non-empty, compact subset ofe, and let A be a subset of e \ K that meets every bounded component of e \ K. Let f E OK. Then there is a sequence (rn)n~l of rational functions each having all its poles in A such that r n -+ f uniformly on K. Proof Let the function f be analytic on the open neighbourhood U of K. By Proposition 1.31, there is a contour, say "I, in U \ K that surrounds K in U. Let E be the closed subspace of R(K) generated by {un: a E e \ K}. Since the function a f---> Un, e \ K -+ E, is continuous, the integral
2~i j
f(a)u n da
defines an element, say g, in E. But, for each Z E K, the map h f---> h(z), E -+ e, is a continuous linear functional on E, so that we see from Theorem 3.10 that
.j
1 g(z) = -2 7rl
f(a)un(z) da = f(z)
(z E K),
'Y
now using the Cauchy integral formula. This proves that f I K = gEE. Set B = R(K)(S), the closed subalgebra of R(K) generated by S, where S = {Z, u" : A E A}. Then, by Lemma 4.82, Un E B for every a E e \ K. It follows that E r;;;; B. o The conclusion of the theorem now follows easily.
213
Banach algebras
Corollary 4.84 Let K be a non-empty, compact subset of re, and let A be a subset of re \ K that meets every bounded component of re \ K. Then the set of all rational functions havmg all their poles in A is dense in R( K). Proof We apply the theorem with having all its poles in re \ K. Corollary
4.85
f
taken to be an arbitrary rational function
D
Let K be a non-empty, compact subset of rc. Then:
(i) R(K) = O(K) ; (ii) R(K)
= P(K) if and only if re \ K is connected.
D
Let U be a non-empty, open subset of rc. We write R(U) for the subalgebra of O(U) consisting of all rational functions having their poles in re \ U. Corollary 4.86 (Runge's theorem for open sets) Let U be a non-empty, open subset of rc. Then R(U) is dense m O(U) in the topology of local uniform convergence. Proof Let f E O(U). We first claim that, for each non-empty, compact subset K of U and each c > 0, there is some r E R(U) with If - rlK < c. Indeed, choose n E N such that, for all z E K, we have both Izl ::; nand dist(z, re \ U) 2 lin, and then define
L = {z
Ere:
Izl ::; n, d(z,re \ U) 2 lin}.
Clearly, L is a compact subspace of re with U ~ L :2 K. Suppose that V is a bounded component of re \ L. Then av ~ L, and so Izl ::; n (z E V). But this implies that V must meet re \ U, since otherwise, for every z E V, we would have d(z, re \ U) 2 deL, re \ U) 2 lin, so that z E L, a contradiction. Thus, every bounded component of re \ L meets re \ U. We may then apply Theorem 4.83 to L with a set A ~ re \ U to see that f may be uniformly approximated on L, and so also on K, by functions in R(U). This gives the claim. As on page 114, we write U = U:=l K n , with each Kn a compact subset of U and Kn C int K n+ 1 (n EN). By the claim, for each n E N, there is a function rn E R(U) with If - rnlKn < lin. Clearly, rn ----+ f in the topology of local uniform convergence on O(U). D
4.15
The holomorphic functional calculus. Recall that, for each algebra A with an identity and for each a E A, we may define the element p( a) in A for each polynomial p; the map
e :p
1----*
p( a) ,
q Xl
----+
A,
is a unital homomorphism with e(x) = a. The map e is a 'polynomial calculus' for the element a. We wish to extend this idea to define f (a) for elements a in a
214
Introduction to Banach Spaces and Algebras
Banach algebra A and functions J that are in a unital algebra B that is strictly larger than qXj in such a way that the map 8 : J f---4 J(a), B ~ A, is a unital homomorphism, still with 8(X) = a. We start this process of defining more general functional calculus maps with a modest first step involving rational functions. Let A be a Banach algebra with an identity, let a E A, and take U to be an open neighbourhood of Spa in
r
f---4
r(a) ,
R(U)
~
A,
is a homomorphism. It will be shown in the next lemma that this mapping is continuous, and then Corollary 4.86 will give a unique extension to a continuous, unital homomorphism 8 a : O(U) ~ A such that 8 a (Z) = a. Notice that, for every r E R(U), it is clear that r(a) E C := {a}CC, the bi-commutant of {a} in A; since C is closed, it will then also follow that 1m 8 a ~ C. Now let a E A. Then we may define 00
exp a == ea =
L n=O
an n! '
where a O = 1. Note that the convergence of this series is an immediate consequence of the completeness of A, with the inequality Ilanll : : ; Iiali n (n EN). In a similar way, we may define sin a and cos a in A. More generally, we may define an element J(a) E A for each entire function J: if J has the Taylor expansion J(z) = L~=o An Zn for z E C, then J(a) = L~=o Anan. It is easily checked that the map
8: J
f---4
J(a) ,
is a unital homomorphism with 8( Z)
O(C)
~
A,
= a.
Lemma 4.87 Let A be a Banach algebra with an identity, and let a E A. Take "I to be a contour in C \ Sp a such that n( "I; z) = 1 Jor all z E Sp a. Then
2~i I(A1 -y
a)-l dA
=
1•
Proof Let I be the left-hand side of the above equation, noting that the existence of this integral is ensured by Theorem 3.10 and Corollary 4.12. Take
215
Banach algebras
R > Iiall, and let fR be the circle given by fR(t) = Re it (0 :::; t :::; 27r). Then n(fR;z) = 1 for all Z E Spa, and thus Theorem 3.11(i) shows that
I=~ 27rI
r (>.I-a)-ld>..
irR
But, on the circle [rR], we have (>.1 - a)-l = 2:::=0 >.-(n+l)a n , with uniform convergence. Termwise integration then immediately gives I = 1, the identity of A. D
Lemma 4.88 Let A be a Banach algebra with an identity, and let a E A. Let U be an open neighbourhood of Sp a, and let "( be a contour that surrounds Sp a zn U. Then, for every r E R(U), we have r(a) = Further, the homomorphism r
~1 2m
f--->
r(>') (>.1 - a)-l d>.. "Y
r(a), R(U)
--+
A, is continuous.
Proof By elementary algebra, there is an analytic A-valued function h on U with r(>')1 - r(a) = (>.1 - a)h(>') (>. E U). By Cauchy's theorem, we have J"Y h(>') d>' = 0, so that
~1 2m
r(>.)(>.1 - a)-l d>' = "Y
r(a)~ 1(>'1 27rI
a)-l d>' = r(a)
"Y
by Lemma 4.87. The function>. f---> Ra(>') = (>.1 - a)-l is continuous, and hence bounded, on the compact set b]' and so it is clear from the formula for r(a) that Ilr(a)11 :::; m Irlh]' where m is a constant not depending on r. This implies that the homomorphism r f---> r(a), R(U) --+ A, is continuous. D We can now establish what is called the holomorphic functional calculus in one variable for Banach algebras.
Theorem 4.89 (Holomorphic functional calculus) Let A be a unital Banach algebra, let a E A, let U be an open neighbourhood of Sp a in C. Then there is a unique continuous, unital homomorphism 8 a : O(U) --+ A such that 8 a (Z) = a. Set C = {a}ee. Then: (i) for every contour"( that surrounds Spa in U and every f E O(U), we have 1 . 1 f(>.)(>.1 - a)-l d>.; 8 a (f) = -2 7rI "Y (ii) 8 a (r) = r(a) (r E R(U)); (iii) im 8 a <:;;; {ay n {a}ee = {ay n C; (iv)
E
<1>c);
216
Introduction to Banach Spaces and Algebras
Proof It is clear from our preliminary remarks and Lemma 4.88 that there is a continuous, unital homomorphism 8 a : O(U) -4 A such that 8 a (Z) = a, and hence such that (ii) holds. Now suppose that () is any such map. Then, by elementary algebra, it follows from the fact that ()(Z) = a that we must have ()(r) = r(a) = 8 a (r) for every r E R(U). By Corollary 4.86, R(U) is dense in O(U), and so () = 8 a . Thus 8 a is uniquely specified by the stated conditions. It also follows that the formula of (i) holds for every f E O(U). Since Z f = f Z in O(U), we have a8 a (J) = 8 a (J)a, and so im8 a ~ {aVo It is further clear that r(a) E C for every r E R(U), and so, since C is closed in A, we have (iii). It remains to prove (iv). Let f E O(U) and
(2~i
1
2~i
f(A)(A1 - a)-l dA) =
1
f(A)(A -
and so
=
8;.' (J I V)
(J
E
E
A. Suppose that U
O(U)) .
Proof The mapping f 1-+ 8-:; (J I V), O(U) - 4 A, is evidently a continuous homomorphism mapping Z to a. The result follows from the uniqueness statement in Theorem 4.89. 0
217
Banach algebras
This last lemma justifies us in writing 'Sa' rather than 8~: if we simply have a function J analytic on an unspecified neighbourhood of Spa, then Sa(f) is well defined by setting it equal to 8~ (f) for any open neighbourhood U of Sp a in C on which J is defined. Thus we have a unital homomorphism
8 a : Osp a
---+
A
such that 8 a (Z) = a, and this homomorphism is continuous when OSpa has the topology described on page 212. Further, 8 a is the unique continuous homomorphism with the specified properties. The convention of writing J(a) := 8 a (f) is likewise unambiguous. In preparation for the next proposition, we remark that, if A is a commutative, unital Banach algebra and at, a2 E A satisfy at = a2, then it follows that SPA (at) = SPA(a2). Let U be an open set in C with U ::::l SpA(at}. Then there are the two functional calculus homomorphisms 8 al : J f-+ J(at) and 8 a2 : J f-+ J(a2) from O(U) into A. Proposition 4.92 Let A be a commutative, unital Banach algebra, and take elements at, a2 E A with at = a2. Let U be an open set in C with U ::::l SPA (at). Suppose that J E O(U) is such that
J(at) = J(a2) ,
while
f'(z)
#- 0 (z
E
SpA(at}).
Then at = a2. Proof Write K = SPA(at) = SPA(a2), and let "( be a contour in U \ K that surrounds K, but does not surround any point of C \ U. Then we have
0= J(at) - J(a2)
=
=
1 ~1 ~
J(z)((zl- at}-t - (zl - a2)-t) dz
2m
'Y
2m
'Y
J(z)(zl - at)-t(zl - a2)-t(at - a2) dz
=u·(at- a2), where
u=
~ 2m
1
J(z)(zl - at}-t(zl - a2)-t dz.
'Y
Let t.p E
.j J(z)(z - a)-2
1 t.p(u) = -2 7rl
'Y
dz
= a2, say a = t.p(at} = t.p(a2). = 1'(a) #- 0
since a E K. It follows from Theorem 4.59 that u is invertible in A, so that
at - a2 =
o.
D
218
Introduction to Banach Spaces and Algebras
Proposition 4.93 Let A be a unital Banach algebra, and let a E A. Suppose that f E OSpa and that an ----+ a in A. Then f(a n ) ----+ f(a). Proof We may suppose that f E O(U), where U is an open neighbourhood of Sp a in
II(AI - an) - (AI - a)11 :::;
~ II(AI -
a)-I
r
l
(A
E
[,,(D,
and so, by equation (*) in the proof of Corollary 4.12,
II(Al- an)-I - (AI - a)-III:::; 2M 2 c. It follows that (AI - an)-I ----+ (AI - a)-I uniformly on ["(], and so the result D follows from the formulae involving integrals for f(a n ) and f(a).
Proposition 4.94 Let A and B be unital Banach algebras, and suppose that T : A ----+ B is a continuous, unital homomorphism. Then, for any a E A and f analytic on some neighbourhood of SPA a, we have T(J(a))
Proof Since SPA a Then
~
= f(Ta) .
Sp BTa, f is also analytic on a neighbourhood of Sp BTa.
e:f
f---4
T(J(a)) ,
O(U)
----+
B,
is a continuous, unital homomorphism such that e(Z) so that T(J(a)) = 8 T (a)(f) = f(Ta).
= Ta. Hence
e
=
8 T (a), D
The next theorem discusses the behaviour of the functional calculus with respect to compositions.
Theorem 4.95 Let A be a unital Banach algebra, let a E A, and let f E OSPAa· Then (g 0 f)(a) = g(f(a)) (g E OSPAf(a)). Proof Set b = f(a). By Theorem 4.89(iv), Spb = f(Spa), and so, for any 9 analytic on a neighbourhood V of Sp b, it follows that 9 0 f is well defined and analytic on some neighbourhood of Sp a. Define e : O(V) ----+ A by 8(g) = 8 a (g 0 f) (g E O(V)). Then clearly e is a continuous, unital homomorphism; also, Z 0 f = f, and so 8(Z) = 8 a (f) = b. It follows from the uniqueness assertion in Theorem 4.89 that 8 = 8b, and so (g
as required.
0
f)(a)
= (}(g) = 8b(g) = g(b)
(g E O(V)) , D
219
Banach algebras
Recall that an element p in an algebra A is an idempotent if p2 = p. Theorem 4.96 Let A be a unital Banach algebra, and let a E A be such that Sp a is the dzsjoint union of two non-empty, closed subsets K and L. Then there is a non-trivzal idempotent pEA such that ap = pa and such that Sp(pa)=KU{O}
and
Sp(a-pa)=LU{O}.
Proof There are open neighbourhoods U of K and V of L, respectively, with Un V = 0. Define f on U U V to be the characteristic function of U, so that f is an idempotent in O(U U V). Set p = GaU). Then p is an idempotent in A; by Theorem 4.S9(iii), ap = pa. By Theorem 4.S9(iv), Spp = f(Spa) = {O, I}, and so p is a non-trivial idempotent. Also, by Theorem 4.S9(iv), Sp (pa) = K U {O} andSp(a-pa)=LU{O}. D Corollary 4.97 Let E be a Banach space, and let T E B(E). Suppose that SpT is the disjoint union of two non-empty, closed subsets K and L. Then there are non-zero projections P, Q E B(E) such that P + Q = IE, PQ = QP = 0, T P = PT, and such that P(E) and Q(E) are proper invariant subspaces of E such that Sp 13(p(E))(T I P(E)) = K and Sp 13(Q(E))(T I Q(E)) = L. Proof Take P E B(E) as specified in Theorem 4.96, and set Q = IE - P, so that Q2 = Q; all is clear, save perhaps that Sp 13(P(E)) (T I P(E)) = K. Set Ko = Sp 13(P(E)) (T I P(E)); we shall show that Ko = K. Let A E K, and assume that A rf. Ko. Then there exists A E B(P(E)) with (AlE - T)Ax
= A(AlE - T)x (x
E
P(E)).
Let g(fJ) = 0 for fJ in a neighbourhood of K and g(fJ) = (A - fJ)-l for fJ in a neighbourhood of L, so that 9 E OSpT. Then g(T)(AlE - T)
Define B = AP
E
(B
=
(AlE - T)g(T)
= Q.
B(E). Then
+ g(T))(AIe -
T) = (AlE - T)(B
+ g(T))
= Ie,
and so A rf. Sp T, a contradiction. Thus K ~ Ko. Conversely, take A E C \ K. Define h(fJ) = 0 for fJ in a neighbourhood of L and h(fJ) = (A - fJ)-l for fJ in a neighbourhood of K not containing A. Then h(T)(AlE - T) = (AlE - T)h(T) = P. Set R = h(T) I P(E) and S = T I P(E). Then R(Alp(E) - S) = (Alp(E) - S)R = Ip(E) , and so A rf. Ko. Thus Ko
~
K.
o
220
Introduction to Banach Spaces and Algebras
Let E be a Banach space, and let T E K(E). We know from Theorem 4.34 that each non-zero ,X E SpT is an eigenvalue and that the eigenspace E('x) is finite-dimensional. By Proposition 3.8(i), there is a projection P).. : E ----) E('x); of course, SPS(E()"))P).. I E('x) = {A}. Let A be a Banach algebra with an identity, and let a E A. We have previously defined exp a, sin a, and cos a in A. Now regard exp, sin, and cos as entire functions on C. Then these functions belong to Osp a, and clearly exp a
= e a (exp) ,
etc. Note that exp(ia) = cosa + isina, for example. We shall sometimes write exp A = {exp a : a E A} .
Proposition 4.98 Let A be a unital Banach algebra. Then:
(i) for a, bE A with ab = ba, we have exp(a + b) = (expa)(expb); (ii) for every a E A, we have expa E G(A), with inverse exp(-a).
Proof (i) The identity of A is lA. For every n n
Xn
=
lA
+
L
k=l
k
Yn
= lA+
L
N, set n
bk k!'
n
~!'
E
Zn
lA
k=l
n Ilbll k Bn=L~'
= ~~, ~ k! k=O
+L
(a ;,b)k
k=l
and set An
=
en
=
t
(Hall + Ilbll)k k!
k=O
k=O
Then, by elementary algebra, since ab = ba, we have n
XnYn - Zn
= L
aJkaJbk,
J,k=l
where aJk :::: 0 for all j, kEN. Therefore, n
IIXnYn - Znll :::; L aJkllaWllbll k J,k=l
But limn~oo(AnBn - en) limn---;oo IIXnYn - Znll = O. The result follows.
= AnBn - en·
ellallellbll - e(llall+llblll
0, and so we see that
(ii) Take b = -a in (i), it being clear that expO = lA.
0
221
Banach algebras
We shall now show that some elements in a Banach algebra have square roots and logarithms. Recall that we write lR - = {x E lR : x ::; O}; also, we shall denote by II the open right-hand half-plane, so that
II = {z = x
+ iy E C : x > O} ,
as in Exercise 4.23. In the next two results, we set D
= C \ lR-.
Theorem 4.99 Let A be a unital Banach algebra, and let a E A be such that Sp a cD. Then there is a unique element b E A such that both b2 = a and Spb c II. Further, bE {a}C n {a}CC. Proof For ZED, we may write Z = re iO with r > 0 and
J(reiIJ)
-Jr
< () < Jr. Set
= rl/2eiO/2
for such zED. Then it is well known that J is the unique analytic function on D such that J(z)2 = Z (z E D) and J(I) = 1; note that J(z) E II for all zED. Set b = J(a). By Theorem 4.95, b2 = a, and this implies that b EA. Also, Sp b = J(Sp a), so that Sp b c II. Again from Theorem 4.89(iii), b E {a}C n {a }CC. Now let C E A satisfy c2 = a. Then c commutes with a = c2 , so that also c commutes with b, and hence
(b - c) (b + c) = b2
-
c2 = 0 .
By Corollary 4.48(i), Sp (b + c) ~ Sp b + Sp c. In addition, suppose that c also satisfies Spc c II. Then Sp(b+c) c II; in particular, 0 (j. Sp(b+c), and so b+c is invertible in A. Hence c = b, and so b is uniquely specified by the prescribed conditions. 0 Under the conditions of Theorem 4.99, we shall occasionally refer to b as the principal square root of a. Theorem 4.100 Let A be a unital Banach algebra, and let a Sp a cD. Then there exists b E A with exp b = a. Proof Let log be the unique analytic function
exp(f(z)) = z
(z E D)
J on D
and
E
A be such that
such that
J(I) = 0,
o
and set b = log a E A. By Theorem 4.95, a = exp b. In the above case, we set b = loga. Further, we set
ac'
=
exp((b)
(( E
q.
Then (aC, : (E C) is a semigroup in (A, .), and the map (f--> ac', C entire A-valued function.
--+
A, is an
222
Introduction to Banach Spaces and Algebras
Now suppose that a E A with p(l - a) uniquely specified by the formula
<
1. Then Spa
c
D, and loga is
10ga=_I=(1-a)n n=l
n
Thus we obtain the following result, which will be referred to later. Proposition 4.101 Let A be a unital Banach algebra. Then expA contains the neighbourhood {a E A : 111 - all < I} of 1. 0 The following is an attractive curiosity. Theorem 4.102 (Kahane and Zelazko) Let A be a commutative, unital Banach algebra, and let f be a linear functional on A with f (1) = 1 and such that f (exp a) -=f. 0 (a E A). Then f E A· Proof We first claim that f is continuous and that Ilfll = 1. For assume that there exists a E A with Iiall < 1 and f(a) = 1. By Proposition 4.101, there exists bE A with exp b = 1- a. By hypothesis, f(l-a) = f(exp b) -=f. 0, a contradiction. This gives the claim. For a E A, set G (z) = f (exp za) (z E q. Then G is an entire function, say
=f
G(z)
(~ z:~n) = ~ znf(,a n )
(z E
q.
N ow we see that IG(z)1
s::
I= Izl n=O
n
Ii,a n.
r = exp(lzlllall)
(z
E
q.
(*)
By hypothesis, G has no zeros, and so, by Proposition 1.37, G = exp F for some entire function F. Since G(O) = 1, we may suppose that F(O) = O. By (*), ReF(z) s:: Iziliall (z E q. By Theorem 1.40, F = aZ for some a E C, and so
= exp(az) =
n
anz L -,n. 00
G(z)
(z E
q.
n=O
Thus f(a n ) = an = f(a)n (n E N) by comparing coefficients in the two expansions of G(z); in particular, f(a 2 ) = f(a)2. This implies that f E A. 0
223
Banach algebras
4.16 The principal component of the group of units. This section follows on from Section 4.4; it comes at the present point because it makes some use of the functional calculus. Let A be a Banach algebra with an identity, and let G = G(A) be the group of units of A, topologized as a subset of A. We know that G is an open subset of A and that the mapping a f---+ a -1, G - t G, is a homeomorphism. Of course, we also know that the multiplication map (a, b) f---+ ab, G x G - t G, is continuous because it is the restriction of the multiplication map A x A - t A. It then follows that, for any given a E G, the maps b f---+ ab and b f---+ ba are homeomorphisms of G onto itself. (These statements may be summarized by saying that G is a topological group; but we shall not assume any knowledge of the theory of such structures. ) We define the principal component Go = Go(A) of G to be the component of G that contains the identity element 1 of A. Lemma 4.103 Let A be a unital Banach algebra. Then expA <;;; Go. Proof Let a E A. Consider the mapping f : t s,t E 1I, we have
Ilf(s) - f(t)11
=
f---+
exp(ta), 1I
-t
G. For each
Ilexp(ta) (exp (s - t)a - 1)11 :::; ellall (els-tiliall - 1) ,
and so f is continuous by the continuity of the real exponential function. Thus f defines a continuous path in G from 1 = f (0) to exp a = f (1). This shows that expa EGo. D Lemma 4.104 Let A be a unital Banach algebra. For each a E G(A) and bE A such that lib - all < Iia-lll- l , we have a-lb E exp A. Proof As in Corollary 4.11, b = a(1 - a- l (a - b))). Since Ila-l(a - b) II < 1, we D have a-lb = 1 - a-lea - b) E expA by Proposition 4.101. Theorem 4.105 Let A be a unital Banach algebra. Then: (i) Go is an open-and-closed, normal subgroup of G(A), and the components of G are precisely the cosets of Go in G(A); (ii) Go consists of all finite products of exponentials, so that
Go = {exp(adexp(a2)·· ·exp(ak): al, ... ,ak E A, kEN};
(iii) in the case where A is commutative, Go = exp A. Proof (i) Set G = G(A). Recall that the continuous image of a connected topological space is connected. Define fL: (a, b)
f---+
ab- l
,
Go x Go
-t
G.
Then fL is continuous, so that fL(G o x Go) is a connected subset of G. Also, fL(G o x Go) contains 1 = fL(l, 1). Thus fL(G o x Go) ~ Go, and this shows that
Introduction to Banach Spaces and Algebras
224
Go is a subgroup of G. Similarly, for each bEG, the mapping D:b : a f---+ b-1ab is a continuous map, and D:b (1 A) = 1A, so that again D:b ( Go) ~ Go. Thus Go is a normal subgroup of G. That Go is closed in G is a general result of point-set topology; it holds simply because Go is a component of G. This completes the proof of (i), except for showing that Go is open and giving the description of the components of G. We shall first prove clause (ii). (ii) By Lemma 4.103, exp A ~ Go. Let E be the set of all finite products of exponentials. Evidently, E is a subgroup of G, and so, by (i), E ~ Go. We shall next show that E is open in G, from which the rest of the proof will quickly follow. Thus let a E E, say
a
= exp(al)exp(a2)···exp(ak),
where aI, ... , ak E A. But, by Lemma 4.104, if b is sufficiently close to a, then we have b = aexp(c) for some c E A. Hence also bEE, so that E is open. But then every (left) coset of E is also open because each mapping a f---+ ba, G --+ G, is a homeomorphism of G. It follows that G \ E is also open because it is the union of all the cosets of E other than E itself. Hence E is closed in G. We have shown that E is a non-empty, open-and-closed subset of the connected set Go, so in fact E = Go. It then follows that each coset of Go in G is also connected and open-and-closed in G. Thus the cosets of Go are precisely the topological components of G. This proves (ii), and completes the proof of (i). (iii) This is immediate from (ii) and Proposition 4.98(i).
D
Notes For alternative treatments of the holomorphic functional calculus, see [47, Section 2.4]. See also [31,32, 79, 126, 143]. We have touched on its applications of the functional calculus to the study of Sp T when T E B(E). This is the beginning of the spectral theory of linear operators; see [63, 114] for much more. The theorem of Kahane and Zelazko, Theorem 4.102, is based on [104]. For an elementary proof, see [142]. Let A be a Banach algebra with an identity. Then the quotient group G(A)/Go is called the index group of A; the quotient map from G(A) onto G(A)/Go is an epimorphism of groups. This index is discussed in [62]. For 'automatic continuity' and uniqueness results for the holomorphic functional calculus, see [44] and [158]. Exercise 4.29 We used Runge's theorem, Corollary 4.86, in the proof of Theorem 4.89. Here is an alternative approach, which avoids an appeal to Runge's theorem. Let A be a Banach algebra with an identity, let a E A, and take U to be an open neighbourhood of Sp a in C. For j E O(U), define 8 a (f) as in clause (i) of Theorem 4.89. Then all the specified properties of the map 8 : O(U) -> A follow as before, save for the fact that 8 a (fg) = 8 a (f)8 a (g) whenever j,g E O(U). For this, first choose a contour 1'1 that surrounds Spa in U to specify 8 a (f). Let V be the open set bounded by bd, and choose a contour 1'2 that surrounds Sp a in V to specify 8 a (g). Then
225
Banach algebras
Use the facts that Ra()..) - Ra(P,) = -().. - P,)Ra()..)Ra(P,) and that
!
f()..)(p, - )..)-1 d)"
~2
to deduce that indeed 8 a(fg)
= 0
and
= 8 a(f)8 a(g)
g()..)
=!
g(p,)(p,l - a)-I dp,
~1
when f,g E O(U).
Exercise 4.30 Let A be a unital Banach algebra, so that expA <;::: G(A). (i) Let a E A. Prove that expa = limn~=(1 + a/n)n. (iii) Let A = MIn, the algebra of n x n matrices. Show that exp A = G(A). (ii) Let A = G(T). Show that expA =I- G(A). Indeed, G(A) has count ably many components, each containing exactly one of the functions Zk, where k E Z. Exercise 4.31 This exercise requires some knowledge of algebraic topology. Let K be a non-empty, compact Hausdorff space. Prove that the index group of the Banach algebra G(K) is naturally isomorphic to HI (K; Z), which is the first Cech cohomology group of K with integer coefficients. Exercise 4.32 Let A be a commutative Banach algebra. An analytic semigroup in A on TI is an analytic A-valued function z f---> a Z , TI ---> A, with a Z+ w = aZa w (z, wE TI). (i) Let (a Z : z E TI) be an analytic semi group in A. Show that a ZA = aA (z E TI). (ii) For z E TI and t E JR+, define
where r is the Gamma function. Show that /z E £I(JR+), and calculate the Laplace transform (see Exercise 2.39) of F for z E TI. Deduce that (F : z E TI) is an analytic semigroup in (£I(JR+); *). (This semi group is called the fractional integral semigroup.) (iii) For z E TI and t E JR, define 1 GZ(t) = 271'1/2 Z I/2 exp
(_t
2
~
)
Show that G Z E £I(JR), and calculate the Laplace transform of G Z for z E TI. Deduce that (G z : z E TI) is an analytic semigroup in (£I(JR); *). (This semigroup is called the Gaussian semigroup.)
5
Representation theory
Representations and modules 5.1 Modules. Let A be a complex algebra, and let E be a complex vector space. A representation of A on E is a homomorphism T : a f---+ Ta from A into the algebra £(E) of all linear endomorphisms of E. These representations are of great importance in pure algebra, especially when E is finite-dimensional, so that a representation can be regarded as a homomorphism into an algebra of matrices. A related theory is very important in Banach-algebra theory; however, usually we replace the finite-dimensional space E by an infinite-dimensional Banach space and consider continuous homomorphisms from our Banach algebra into 13(E). Given a representation T of an algebra A on E, we can give E the structure of a (left-) A-module by defining
a .
~ =
(Ta)(O
(a
E
A,
~ E
E).
We may also speak about the algebra A acting on the space E (via the homomorphism T). We are not supposing any knowledge of module theory, but it is often convenient to use some of the language of this subject. Indeed, a vector space E is an A-module for a product . if the map (a,~)
f---+
a . ~,
A x E
--->
E,
(*)
is bilinear and ab . ~ = a . (b . 0 (a, bE A, ~ E E). (Strictly, we should say 'left A-module'.) For example, a complex vector space is a ((>module in the obvious way, and so the notion of an A-module generalizes that of a vector space. Let E be an A-module, and let a E A. Then
a . E = {a . ~ : ~
E
E}
and
E.L = {a
E
A : a . E = {O}}.
A representation T : a f---+ Ta, A ---> £(E), is faithful if the map T is injective; equivalently the A-module E is faithful if E.L = {O}. Suppose that (A; I '11) is a normed algebra, (E; 1·1) is a normed space, that T : A ---> £(E) is a representation, and that each Ta is a bounded linear operator on E, so that Ta E 13(E). Then we say that the homomorphism T is a normed representation of A (or that E is a normed A-module). Thus, in this case, we require that, for each a E A, there is a constant Ka > 0 such that
la . ~I
:::; Kal~1
(~E
E).
Representation theory
227
This normed representation or the A-module is said to be continuous if the mapping T : (A; 11·11) ----> (8(E); 11·11) is continuous, where 11·11 also denotes the operator norm on (E; 1·1), or, equivalently (using Exercise 2.16), if the bilinear map in (*) is continuous, and so now we require that there be a constant K > 0 such that la . ~I ::; K lIalll~1 (a E A, ~ E E). It is easy to see that, in this case, we may replace the norm on E by an equivalent
norm to ensure that K
= 1,
la . ~I
::;
and hence that lIalll~1
(a
E
A, ~ E E);
(**)
we shall suppose that this has been done. In the case where (E; 1·1) is a Banach space and (**) holds, we say that E is a Banach left A-module. The great advantage of the module viewpoint is that we have the obvious (algebraic) notions of module homomorphism, isomorphism, submodu1e and quotient module. Let A be an algebra, and let E be an A-module Then F <:;; E is a submodu1e of E if F is a vector subspace of E and a . ~ E F whenever a E A and ~ E F; in this case, the quotient vector space ElF is an A-module for the operation a . (~+ F) = a . ~ + F (a E A, ~ E E). Let E and F be A-modules. Then a map T E £(E, F) is a module homomorphism if T( a . ~) = a . T~ (a E A, ~ E E) . Two A-modules E and F are isomorphic if there is a bijective module homomorphism T : E ----> F; we write E c:::: F in this case. Suppose that A is a normed algebra, that E is a normed A-module, and that F is a closed submodule of E. Then, in the quotient norm, ElF is also a normed A-module. The closed submodules of ElF correspond bijectively to those closed submodules of E that contain F. Suppose that E is a continuous normed Amodule. Then the normed A-modules F and ElF are also continuous. For example, let A be an algebra, and let I be a left ideal in A. Then I and AI I are left A-modules. In the case where A is a Banach algebra and I is a closed left ideal, I and AI I are Banach left A-modules. Let A be an algebra, and let E be an A-module. For ~ E E, set A .
~
= {a .
~
:a
E
A},
so that A . ~ is a submodule of E. A vector ~o E E such that E = A . ~o is a cyclic vector for E; the module E is cyclic if there is a cyclic vector, and E (or the corresponding representation) is irreducible (or simple) if it is non-zero and has no submodules other than {O} and E itself. Evidently, a non-zero module E is irreducible if and only if every non-zero vector in E is cyclic. Thus, in the case where A has an identity 1, we have 1 . ~ = ~ (~ E E) for an irreducible module E.
228
Introduction to Banach Spaces and Algebras
Example 5.1 Let E be a complex Banach space, and let Q( ~ 8(E) be a Banach operator algebra on E, as in Example 4.6. In particular, consider Q( = 8(E) itself. Then E is, of course, an Q(-module in a natural way, just taking T . x = Tx
(T E Q(, x E E) .
We shall show that E is an irreducible Q(-module. Thus take x, y E E with x -=I- o. We must show that there is some T E Q( with Tx = y. But this follows at once from the Hahn-Banach theorem. Indeed, there is an element J E E* with J(x) = 1, and we may define T by setting Tz = J(z)y (z E E), so that T = y ® J E Q(. Evidently Tx = y, as required. 0 Let A be a Banach algebra with an identity, 1. Then the left-regular representation of A is the homomorphism
T :a
f---+
La,
A
--t
8(A) ,
where, as before, La(b) = ab (b E A) for each a E A. Clearly, this representation is continuous, and it has 1 as a cyclic vector. In treating A as a (left-) A-module, we always intend this representation. For the left-regular representation, a submodule is just a left ideal of A. Let L be such a left ideal. Then AlLis also cyclic, with cyclic vector the coset 1 + L. It follows from Proposition 4.1 that the module AlLis irreducible if and only if L is a maximal left ideal. Since every maximal left ideal in a Banach algebra is closed, the module AlLis then a Banach left A-module.
Proposition 5.2 Let A be an algebra with an identity, and let E be an irreducible A-module. Then E ~ AIL for some maximal left ideal L of A. Proof Let 0 -=I- ~o E E, so that A . ~o = E. Set L = {a E A: a . ~o = O}. Then a f---+ a . ~o, A --t E, is a surjective module homomorphism with kernel L. So L is a left ideal and E ~ AIL. But L is maximal because the module E, and therefore AIL, is irreducible. 0 Corollary 5.3 Let A be a Banach algebra with an identity, let E be an zrreducible A-module, and let ~o E E with ~o -=I- o. For every ~ E E, define: I~I = inf{llall : a E A, a . ~o =
O·
Then 1·1 zs a complete norm on E such that 10, . ~I ::::; Ilalll~1 (a E A, ~ E E). Proof The ideal L of the above proof is closed in A, and so AlLis a Banach algebra with respect to the quotient norm. The given norm on E is just this quotient norm transferred to E by the isomorphism of Proposition 5.2. 0 The above corollary shows that every irreducible A-module E over a unital Banach algebra A is a Banach left A-module for a suitable norm on E.
229
Representation theory
5.2 Bimodules and derivations. Let A be a complex algebra, and let E be a complex vector space. Then E is an A-bimodule with respect to bilinear maps (a,~) f---+ a . ~ and (a,~) f---+ ~ • a from A x E to E whenever the following hold for all a, b E A and ~ E E:
(i) a . (b . ~) = ab . ~; (ii) (~ . a) . b = ~ . ab; (iii) a . (~ . b) = (a . ~) . b. For example, the algebra A itself, and each ideal in A, are A-bimodules with respect to the product map in A. Suppose that A is a subalgebra of an algebra B. Then B is an A-bimodule with respect to the product in B. Now suppose that (A; 11·11) is a Banach algebra and that (E; 1·1) is a Banach space such that E is an A-bimodule with respect to the above maps. Then E is a Banach A-bimodule if the two bilinear maps are continuous. In this case, by changing to an equivalent norm on E, we may and will suppose that
la . ~I
::;
Ilalll~l,
I~· al ::; Ilalll~1
(a E A, ~ E E).
For example, a Banach algebra A itself, and each closed ideal I in A, are Banach A-bimodules with respect to the product map in A, and the quotient space AI I is a Banach A-bimodule for the products
a . (b
+ I)
=
ab + I,
(b
+ I)
. a = ba + I
(a, bE A).
Examples 5.4 Let A be a Banach algebra. (i) Let E be a Banach left A-module. We define an action of A on the right by ~ . a = 0 (a E A, ~ E E). Then it is clear that E is now a Banach A-bimodule. (ii) Let E be a Banach A-bimodule. Then the dual space E* of E is a Banach A-bimodule for the operations defined by (a, J) f---+ a . f and (a, J) f---+ f . a, where
(x, f . a) = (a . x, 1) ,
(x, a . 1) = (x . a, 1)
(a
E
A, x
E
E,
f
E
E*) .
In this case, E* is the dual bimodule of E. (iii) Suppose that A is commutative, and take E to be a Banach left Amodule. We define an action of A on the right by ~ . a = a . ~ (a E A, ~ E E). Then it is again clear that E is a Banach A-bimodule. In this case, we say that E is a Banach A-mod ule. (iv) Let cp E
a . z = z . a = cp(a)z
(a
E
as is very easily verified; this bimodule is called
A, z
E
<eip.
o
Let A be an algebra, and let E be an A-bimodule. Then a derivation from A to E is a linear map D E £(A, E) such that
D(ab) = a . Db + Da . b (a, bE A). For example, let ~ E E, and set 0e(a) = a . ~ -~ . a (a E A). Then oe : A ---> E is easily checked to be a derivation; these derivations are called inner derivations.
230
Introduction to Banach Spaces and Algebras
In particular, consider the case where E = A (and the module operations are the product in A). Then a derivation D : A -> A is called a derivation on A. Examples 5.5 (i) Consider the A-bimodule C
= 'P(a)d(b) + 'P(b)d(a)
(a, bE A);
such maps are called point derivations at 'P. For example, let A = A(~), 'P = EO, and d(f) = 1'(0) (f E A). Then d is a continuous point derivations at EO. (ii) Let A = C l(n) and E = C(n). Then E is a Banach A-module with respect to the usual pointwise product of functions, and the map D : f f---+ 1', A -> E, is a continuous derivation. (iii) Let U be a non-empty, open subset of C. Then D : f f---+ I' is a continuous derivation on the Frechet algebra O(U). 0 The following version of Leibnitz's rule is proved by an easy induction; by convention, DO is the identity operator on A. Proposition 5.6 Let A be an algebra, and let D be a derivation on A. Then, for each n EN, we have Dn(ab) =
:t (~)
(Dka)(Dn-kb)
(a, bE A).
k=O
o
Theorem 5.7 (Singer and Wermer) Let A be a Banach algebra, and let D be a continuous derivation on A. Then exp D is a continuous automorphism on A. In the case where A is commutative, D(A) ~ J(A). Proof As before, exp D E B(A) is defined by Dn
L ---(!n. 00
(exp D)(a) =
(a E A).
n=O
Now take a, bE A. Then we calculate that (expD)(ab)
=
~ D~~b) = ~ ~! ~ (~)(Dka)(Dn-kb)
~~ 1 k 1 n-k = L L k! D a . ..D b= n=Ok=O
(f ~~a) (f ~~b) ,=0
}=o
= (exp D) (a) (exp D)(b) , and hence exp D is a homomorphism on A. The inverse of exp D is exp( - D), and so exp D is a continuous automorphism on A.
Representation theory
231
Let z E
Irp(exp(zD))(a)1 ::;
Iiall
(a
E
A).
Let a E A. Then the map j : z f---+ rp(exp(zD)(a)), C ~ C, is bounded, and it is clearly entire. By Liouville's theorem, Proposition 1.39, j is constant. In particular, the coefficient of z in the power-series expansion of j, namely rp(Da), is zero. Thus rp(Da) = 0 (rp E A)' Now suppose that A is commutative. By Corollary 4.60 (applied to A+), D(A) ~ Q(A) = J(A). D The following corollary will be strengthened in Corollary 5.34.
Corollary 5.8 Let A be a commutative, semisimple Banach algebra. Then there are no non-zero, continuous derivations on A. D
5.3 The Jacobson radical. Let A be an algebra, and let L be a left ideal of A. The standard representation of A on AlLis defined by (Ta)(b+L)=ab+L so that T : A by
~
(a,bEA),
£( AI L) is a homomorphism. We define the quotient L : A of A L :A
=
{a
E
A : aA
~
L} .
Thus L : A is the kernel of the above standard representation T of A on AIL, so that L : A is an ideal in A. Suppose that the algebra A has an identity. Then L : A ~ L, and indeed L : A is the maximal ideal that is contained in L. In this case, an ideal P in A is said to be primitive if P = L : A for some maximal left ideal L of A. It is evident that P = E.L is primitive if and only if it is the kernel of an irreducible representation of A on an A-module E. Also, every maximal ideal M of A is a primitive ideal because M ~ L for some maximal left ideal L, and then M ~ L : A, so that M = L : A by the maximality of M. The algebra A is said to be primitive if {O} is a primitive ideal. In the case where A is a commutative algebra with an identity, the primitive ideals of A are simply the maximal ideals; each character rp E A defines an irreducible module structure on C by the formula
a . z = rp(a)z
(a
E
A, z
E
q.
Let A be a Banach algebra with an identity, and let L be a closed left ideal in A. Then L : A is a closed ideal in A. Thus each primitive ideal in A is closed.
232
Introduction to Banach Spaces and Algebras
Let A be an algebra with an identity. Then we define the Jacobson radical, J(A), of A to be the intersection of all the primitive ideals of A or, equivalently, as the intersection of all the kernels of the irreducible representations of A. For an algebra A without an identity, we define J(A) to be J(A+). It is clear that, if A is a commutative Banach algebra, then the present definition of J(A) reduces to the previous one, given on page 193. An algebra A is said to be semisimple if J(A) = {O} and radical if J(A) = A; necessarily a radical algebra does not have an identity. It is immediate that J(A) is always an ideal in an algebra A. We remark that we use the term 'Jacobson radical' because, in pure algebra theory, there are several different radicals that can be defined; some authors just say 'the radical of A' for J(A). Nevertheless J(A) is the only radical that we shall discuss seriously. There is a variety of different characterizations of the Jacobson radical, some intrinsic, and we give some of these in the next theorem. A key point is that the above discussion of representations has been carried out in the context of left A-modules and maximal left ideals. (There is an entirely similar theory involving right A-modules and maximal right ideals.) Thus, at first sight, it appears that the definition of J(A) depends on the choice of 'left' or 'right' in the definition; however we shall see that this is not in fact the case. Let A be an algebra. Then the opposite algebra to A is the algebra B which is the same as A as a vector space, but such that the product in B is taken in the opposite order to that of A. Clearly, (maximal) left ideals of B correspond to (maximal) right ideals of A. The opposite algebra to A is denoted by AOP. Recall from Section 4.4 that G(A) denotes the group of units of an algebra A with an identity, so that G(A) = G(AOP), and that Aa = {ba : bE A}. Theorem 5.9 Let A be an algebra with an identity, and let J Jacobson radical of A. Then:
= J(A) be the
(i) J is the intersection of the maximal left ideals of A; (ii) J = {a
E
A : 1 + Aa
~
G(A)};
(iii) J = {a E A : 1 + aA ~ G(A)}; (iv) J is the intersection of the maximal right ideals of A. Proof (i) Let L be a maximal left ideal. Then J ~ P = L : A ~ L, so that J is contained in the intersection of the maximal left ideals. Conversely, suppose that ao ~ J. Then there is an irreducible representation T: A ----> £(E) such that Tao of- o. Thus there exists ~o E E with ao . ~o of- o. Set
L = {a
E
A :a .
~o =
O} .
Then L is a maximal left ideal of A and a ~ L. This proves (i). (ii) Let a E J and b E A. Assume that 1 + ba is not left invertible in A. Then there is a maximal left ideal L of A with 1 + ba E L. However, ba E L (by (i»,
Representation theory
233
and so 1 E L, contrary to the fact that L =1= A. Thus 1 + ba is left invertible in
A. Now choose C E A so that (1 + c)(1 + ba) = 1, so that c = -ba - cba E J. By the argument of the previous paragraph, 1 + c is left invertible in A. But we know that 1 + c is right invertible, and so 1 + c is invertible, and hence so is 1 + ba. Thus 1 + Aa c:: G(A). Conversely, suppose that a E A has the property that 1 + ba is invertible (or even just left invertible) for every b E A. Let L be a maximal left ideal of A, so that L + Aa is a left ideal in A containing L. Assume that a if- L. Then L + Aa =1= L, and so L + Aa = A by the maximality of L. Thus there is some b E A with 1 + ba E L, which contradicts the left invertibility of 1 + ba. Hence a is in every maximal left ideal, so that a E J by (i). This proves (ii). (iii) This follows from (ii) since, by Proposition 4.16(ii), 1 + ba is invertible if and only if 1 + ab is invertible. (iv) From (ii) and (iii), we see that J(AOP) = J, and so (iv) follows from the equivalence of (i) and (ii). D Corollary 5.10 Let A be an algebra. Then J(A) is an ideal in A and AIJ(A) is semisimple. Proof We may suppose that A has an identity. Set J = J(A), an ideal in A. Let 7r : A -+ AI J be the quotient map. For each irreducible A-module E, J is included in the kernel of the representation of A on E, so that E becomes an irreducible AIJ-module in a natural way (by setting 7r(a) . ~ = a . O. Suppose that ~ E J (AI J). Then ~ = 7r( x) for some x E J, and so ~ = O. This shows that J(AI J) = {O}, and so AI J is semisimple. D
Here is an alternative proof that the quotient AI J is semisimple. Take x E J(AI J) and y E AjJ, say x = 7r(a) and y = 7r(b), where a, b E A. Since 1 + xy is left invertible in AI J, there exists some e E A such that the element u := e(1 + ab) - 1 E J. But then e(1 + ab) = 1 + u is left invertible in A, so that 1 + ab is left invertible in A. Hence a E J and x = 7r( a) = O. Thus J(AI J) = {O}, and so AI J is semisimple. Let A be an algebra. Recall that we are writing Q(A) for the set of quasinilpotent elements of A. A subset S of A is said to be quasi-nilpotent if S c:: Q(A). Corollary 5.11 Let A be an algebra. Then J(A) is a quasi-nilpotent ideal that is the union of all the quasi-nilpotent left ideals in A. Proof We may suppose that A has an identity. Take a E J(A) and A E
234
Introduction to Banach Spaces and Algebras
Corollary 5.12 Let I be an zdeal in an algebra A. Then J(I)
= In J(A).
Proof We may suppose that A has an identity. Take a E J(I) and b E A. Then ba E I and 1 + Iba <;;; 1 + Ia <;;; G(I+) by Theorem 5.9(ii). By the same result, ba <;;; J(I). Thus J(I) is a left ideal in A. Clearly, J(I) is quasi-nilpotent in A, and so J(I) <;;; J(A) by Corollary 5.11. Now suppose that a E I n J(A) and b E I. Then 1 + ba E G(A), and so 1 + ba E G(I + C . 1) by Lemma 4.9, and so, by Theorem 5.9(ii) again, a E J(I). Hence In J(A) <;;; J(I). 0
We now see that the situation for Banach algebras is particularly satisfactory. Theorem 5.13 Let A be a Banach algebra. Then J(A) is a closed ideal in A, and AI J(A) is a semisimple Banach algebra. Proof We may suppose that A has an identity. We know that J(A) is an ideal in A. By definition, J(A) is the intersection of the primitive ideals of A. However, each primitive ideal of A is closed, and so J(A) is closed. By Corollary 5.10, AI J(A) is semisimple. 0 Example 5.14 Let E be a Banach space, and let 2l be a Banach operator algebra on E, as in Example 4.6. We already know from Example 5.1 that 2l is an irreducible B(E)-module, and so 2l is semisimple. As in Proposition 5.2, Lx := {T E 2l : Tx = O}
is a maximal left ideal for each x E E, and
nxEE
Lx = {O}.
o
We have shown in Corollary 4.60 that, for a commutative Banach algebra A, we have J(A) = Q(A). The above example, even in the simplest case where A = M 2 , shows that, for a non-commutative Banach algebra A, it may happen that J(A) S;; Q(A). Indeed, here A is semisimple, but the matrix
(~ ~) is a non-zero nilpotent element of A. This example also shows that a closed, unital, commutative sub algebra of a semisimple algebra need not be semisimple. Indeed, matrices in M2 of the form
(~ ~),
where
A, JL
E
C,
form a closed, unital, commutative subalgebra A of M2 such that J(A) is the one-dimensional space of matrices of the form (
~ ~),
where
JL E Co
The next lemma will enable us to give an attractive new proof of the most important result of the next section, the Jacobson density theorem. We continue to suppose that all our Banach algebras are over the complex field.
235
Representation theory
Theorem 5.15 (Harris and Kadison) Let A be a Banach algebra with an zdentity, let L be a maximal left ideal of A, and let a E A. Suppose that La L. Then there exists A E C such that a - Al E L.
s:
Remark: The converse to this lemma is, of course, trivial. It is also clear that, for given L and a, the complex number A is uniquely determined and belongs to Spa. Proof By Theorem 4.14, the maximal left ideal L is closed in A, and so AIL is a Banach A-module; set E = AIL. We first show that, if ba E L for some b E A, then either a E L or bEL. For suppose that ba E L, but that b 1- L. By Proposition 4.1, there exists e E A such that 1 + eb E L. Hence a + eba E La L. But eba E L, and so a E L. Since La L, we can act on AIL by right multiplication by a, i.e. define Ra E 8(E) by Ra(b+L) = ba+L (b E A). We claim that, if a 1- L, then Ra is an invertible operator. Indeed, Ra is injective, for if Ra (b + L) = L, then ba E L, so that bEL by our second paragraph. Also, Ra is surjective because there exists e E A with ea - 1 E L, and so, given b E A, we have Ra(be + L) = b + L. We have shown that Ra is bijective; by Banach's isomorphism theorem, Corollary 3.41, Ra is invertible in 8(E). For each A E C, we have L(a - AI) Land R a-)..l = Ra - AI. Certainly Sp Ra -=I- 0 by Theorem 4.17; take A E Sp Ra. Then R a-)..l is not invertible, and so a - Al E L. 0
s:
s:
s:
Corollary 5.16 Let A be a Banach algebra with an identity, let L be a maximal left ideal of A, and let a E Q(A) be such that La L. Then a E L.
s:
Proof By Theorem 5.15, there exists A E C with a - Al E L. But, as noted above, A E Sp a. Since a E Q(A), necessarily A = 0, so that a E L. 0 The first application of Theorem 5.15 will be to prove a result of Zemanek that is an important step in proving his characterization of the radical (see Theorem 5.40). First, we need a simple algebraic lemma. Throughout, for an algebra A, we write [a, bJ = ab - ba (a, bE A) . Lemma 5.17 Let A be a unital algebra, let L be a maximal left ideal of A, and let a E A. Suppose that [b, aJ E Q(A) for every bEL. Then La L.
s:
Proof Assume towards a contradiction that La CZ L. Then there exists c E L with ea (j. L. Again by Propositon 4.1, there exists bE A with 1 + bea E L, and hence also 1 + [be, aJ = 1 + bea - abc E L. Since be E L, the hypothesis implies that [be, aJ E Q(A). But then 1 + [be, aJ is invertible, which contradicts the fact that L -I A. Thus La <;;:; L. 0
236
Introduction to Banach Spaces and Algebras
Corollary 5.18 Let A be a Banach algebra, and let a E Q(A) be such that [b,aj E Q(A) for every b EA. Then a E J(A). Proof We may suppose that A has an identity. By Lemma 5.17 and Corollary 5.16, a belongs to every maximal left ideal of A, and so a E J(A). 0 5.4 Irreducible representations. We now return to consideration of irreducible representations or modules. It is useful to make some further definitions. Let A be an algebra, and let E be a left A-module. Take n E N. Then E is n-transitive if, for every linearly independent sequence (6, ... '~n) of elements in E and every sequence (1]1, ... ,1]n) in E, there is some a E A such that a . ~k = 1]k (k = 1, ... , n). Evidently E is irreducible if and only if it is I-transitive. Theorem 5.19 (Jacobson's density theorem) Let A be a Banach algebra with an identity, and let E be an irreducible A-module. Then E is n-transitive for every n E N. It is convenient to prove first two lemmas. Lemma 5.20 Let A be an algebra with an identity, and let E be an irreducible A-module. For each n :::: 2, the property of n-transitivity is equivalent to its special case (*)n, defined as follows: (*)n: for every linearly independent sequence (6, ... , ~n) of elements in E, there exists a E A such that a . ~k = 0 (k = 1, ... , n - 1) and a . ~n =1= o.
Proof Let (6, ... , ~n) be a linearly independent sequence in of elements E, and let (1]1, ... ,1]n) be a sequence of elements in E. We are supposing that (*)n holds. By renumbering the elements, we see that there is, for each k = 1, ... ,n, an element ak E A with ak . ~J = 0 for all j =1= k and with ak . ~k =1= o. We then use the case n = 1 to find, for each k = 1, ... , n, an element Ck E A with Ck . (ak . ~k) = 1]k, and finally define a = L~=l Ckak E A. Clearly, a . ~k = 1]k (k = 1, ... , n), and so E is n-transitive. 0 The next lemma, proving 2-transitivity, is where use will be made of Theorem 5.15, which applies to Banach algebras.
Lemma 5.21 Let A be a Banach algebra with an identity, and let E be an irreducible A-module. Then E is 2-transitive. Proof Let (6,6) be a linearly independent sequence in E. Set L = {a E A : a . 6 = O}, so that L is a maximal left ideal in A. Now choose bE A with b· 6 = 6. Then, for>. E C, we have (b->.I) ·6 = 6->'6 =1= 0, and hence b->'1 ~ L. By Theorem 5.15, it follows that Lb g L, i.e. there is some a E L with ab ~ L. Thus a . 6 = 0, but a . 6 = ab . 6 =1= o. This proves (*)2. By Lemma 5.20, E is 2-transitive. 0
237
Representation theory
Proof of Theorem 5.19 It remains to prove n-transitivity for every n ~ 3. Take n ~ 3, and make the inductive hypothesis that E is (n - l}-transitive; we shall prove that (*}n holds. Thus, let (6, ... ,~n) be a linearly independent sequence of elements in E We set I={aEA:a'~k=O
(k=1, ... ,n-2)},
B = 1+
= 0 (k = 1, ... , n - 2) and a·
~
= 1],
i.e. a E I ~ B and a . ~ = 1]. But now a . I = a . 7r(O = 7r(1]), and so B acts irreducibly on ElF. Clearly, (7r(~n-l),7r(~n)) is linearly independent in ElF and so, by Lemma 5.21, there is some b E B with b . 7r(~n-d = 0 and b . 7r(~n) = 7r(~n); i.e. b . ~n-l E F and b . ~n - ~n E F. By the inductive hypothesis, there is some c E I with c . ~n = ~n' Since IF = {O}, we have cb . ~k = 0 (k = 1, ... , n - 2). Since b . ~n-l E F, we have cb . ~n-l = O. Also, cb . ~n = C • ~n = ~n -I=- O. Thus (*)n holds (with a = cb). This continues the induction. By induction, E is n-transitive for each n E N, and this completes the proof of Theorem 5.19. 0 The following is a classical theorem of Wedderburn. Theorem 5.22 (Wedderburn's theorem) Let A be a finite-dimensional, semisimple algebra (over
238
Introduction to Banach Spaces and Algebras
with PI = J2 + P2' Since h <::; PI, necessarily P2 = PI -12 E PI, and hence P2 E PI n P 2 . Continuing, we eventually obtain Pk-1 = jk + Pk with jk E Jk and Pk E H n ... n P k = {a}. Our claim follows. It follows that A ~ MInI EEl· .. EEl MInk' The identity of A is e1 + ... + ek. 0
Proposition 5.23 Let A be a Banach algebra with an identity, and let E be an irreducible A-module. Suppose that {~n : n E N} is a linearly independent set of distinct elements in E. Then there is a sequence (a n )n21 in A such that an'" a1 . ~m = 0 for m, n E N wzth m < n and an'" a1 . ~n -1= 0 for n E N. Proof Let {17n : n E N} be a linearly independent set of distinct elements in E. By Lemma 5.21, there exists b1 E A such that b1 . 171 = 0 and b1 . 172 -1= o. It follows from Theorem 5.19 by induction that there is a sequence (b n )n>l in A such that, for each n ~ 2, we have Ilbnll ~ 2- n , bn . 171 = ... = bn . 17n = 0, and bn . 17n+ 1
Set b =
rt lin {b
t .
17] : i = 1, ... , n - 1, j = 1, ... , n
L::1 bt ; the series converges in A.
+ I} .
( *)
Then n-1
b . 171
= 0 and b· 17n = ~ bt
.
17n
(n ~ 2).
t=l
Clearly, b· 172 = b1 . 172 -1= o. Assume that b· 17m+1 E lin{b· 171,··" b· 17m} for some m ~ 2. This contradicts the case in which n = m of (*), and so {b . 17n : n ~ 2} is a linearly independent set of distinct elements in E. Now choose a1 = 1 so that a1 . ~ = ~ (~E E). Next, choose a2 E A such that a2al . 6 = 0 and {a2al . ~n : n ~ 2} is a linearly independent set of distinct elements in E. Choose a3 E A such that a3a2a1 . 6 = 0 and {a3a2al . ~n : n ~ 3} is a linearly independent set of distinct elements in E. Continue in this way to obtain the required sequence (a n )n21 in A. 0
Proposition 5.24 Let A be a Banach algebra with an identity, and let (E k )k2 1 be a sequence of fimte-dimensional, irreducible Banach left A-modules. Suppose that Et n· .. n E;, Cl Et for each k > n in N. Then, for each n E N, there exists an E Et n ... nE;, such that an' Ek = Ek (k > n). Proof For kEN, take Tk to be the representation of A on Ek defined by (Tka)(O = a . ~ (a E A, ~ E Ek). Fix n EN, and set B = Et n ... n E;" a Banach space. For k > n, set Uk
=
{b E B : det(Tkb) -1= O};
since det is a continuous function on the space £(Ek ), the set Uk is open in B. Since n ... n E;, Cl Et, there exists bk E Uk. For each bE B \ Uk, we have b = limm~(X)(b + bk/m), and b + bk/m E Uk for sufficiently large mEN, and so bE Uk. Thus Uk is dense in B. By the Baire category theorem, Theorem 1.21, there exists an E B with an E Uk (k > n). Clearly, an . Ek = Ek (k > n), as required. 0
Er
Representation theory
239
Notes For an introductory account of the theory of modules, see [47, Section 1.4] and many books on algebra, including [97]. For an introduction to the theory of Banach left A-modules and Banach A-bimodules, see [47, Section 2.6], where a somewhat different terminology is used; for an extensive account, couched in the language of representations, see [126]. Many examples of Banach A-bimodules are given in [47]; the concept is ubiquitous in more advanced work on Banach algebras. There are many results about derivations on particular Banach algebras in [47]. Here are two examples; for another example, see Section 5.7. (i) Let Ql be a Banach operator algebra on a Banach space E, and let D : Ql -+ B(E) be a derivation. Then there exists T E B(E) such that D(A) = AT - TA (A E Ql) [47, Theorem 2.5.14].
(ii) Let G be a locally compact group. Then the following question was discussed and only partially resolved in [47, Section 5.6]: Does every continuous derivation from Ll(G) to itself have the form D : f f-+ f * f-t - f-t * f for some measure f-t on G? Subsequently, this question has been resolved positively by Losert in [118]. The question whether or not all continuous derivations from a Banach algebra A into a specific Banach A-bimodule are inner is the first step in the enormous cohomology theory of Banach algebras; see [47, Section 2.8] for a preliminary account. A Banach algebra A is amenable if all continuous derivations from A into each dual Banach A-bimodule are inner. This concept was introduced by Johnson [100], where it was proved that a group algebra L 1 (G) is an amenable Banach algebra if and only if G is an amenable locally compact group. Intrinsic characterizations of amenable Banach algebras are given in [47, Theorem 2.9.65]. For example, the Banach algebras C(K) are all amenable. A recent advance is a theorem of Runde [147]: for each p:O:: 1, the Banach algebra B(fP) is not amenable. The Singer~Wermer theorem, Theorem 5.7, was first proved in 1955 in [150]. In this paper, it was conjectured that it was not necessary to assume that the derivation be continuous. This conjecture was finally proved by Thomas in 1988 in [159]: for each derivation on a commutative Banach algebra A, we have D(A) ~ J(A). See [47, Theorem 5.2.48] for a proof. A 'non-commutative Singer~Wermer theorem' would establish that, for each derivation on a Banach algebra A, we have D(P) ~ P for each primitive ideal P; this is true for continuous derivations [148]. This question has defied solution for 55 years; the strongest partial result is in [160]. Let A be an algebra with an identity. What we have called a primitive ideal should, more precisely, be called a left primitive ideal in A because it is defined using left Amodules or, equivalently, maximal left ideals. Similarly, there are right primitive ideals in A, defined using right A-modules or maximal right ideals. In pure algebra, there are left primitive ideals that are not right primitive, but it seems to be a long-standing open question, going back to at least [31, page 125], whether or not the classes of left primitive and right primitive ideals coincide in every unital Banach algebra. Let A be an algebra, not necessarily with an identity. A proper left ideal I in A is modular if there exists an element U E A such that a - au E I for each a E A; in this case, u is a right modular identity of I. A maximal modular left ideal in A is a maximal element in the family of proper, modular left ideals in A. Since each left ideal containing a modular left ideal is also a modular left ideal, a maximal modular left ideal in A is a maximal left ideal. It follows from Zorn's lemma that each proper, modular left ideal is contained in a maximal modular left ideal. We may define the Jacobson radical, J(A), of A without reference to A+: it is equal to the intersection of the maximal modular left ideals of A. This gives the same ideal as that defined in the text, but this version may be a little 'cleaner' and more general. Different radicals from the Jacobson radical are sometimes considered: see [47, 126]. For example, let A be an algebra with an identity. Then the strong radical of A is the
240
Introduction to Banach Spaces and Algebras
intersection of the maximal ideals of A and the prime radical of A is the intersection of the prime ideals of A. (A proper ideal P of an algebra A is prime if, for ideals I and J in A, either I <:;: P or J <:;: P whenever I J <:;: P; an ideal P is prime if and only if either a E P or b E P whenever aAb <:;: P.) Theorem 5.15 is due to Harris and Kadison [90]. Some results, including Theorem 5.15 and Jacobson's density theorem are true for a wider class of algebras than Banach algebras, and are so formulated in texts on algebra; the key fact that we have used is the fundamental theorem of Banach algebras that the spectrum of an element in such an algebra is always non-empty. Jacobson's density theorem, Theorem 5.19, and the Wedderburn theorem, Theorem 5.22, are classical algebraic results. The latter is a rather special case of the more general Wedderburn~Artin theorem, given in many texts, including [97, Chapter IX, Section 3] and [126, Section 8.1]. The related Wedderburn principal theorem states that a finitedimensional algebra A (over q can be written as A = B EB J(A), where B is a unital subalgebra of A with B ~ A/J(A); see [47, Theorem 1.5.18]; possible generalizations to Banach algebras are discussed in [47]. Exercise 5.1 Let A be a Banach algebra, and let E and F be Banach left A-modules. For a E A and l' E B(E, F), define
(a . T)(x) = a . Tx,
(1'. a) (x) = T(a . x)
(x E E).
Verify that B(E, F) is a Banach A-bimodule with respect to the maps (a, 1') and (a, 1') f-> l' . a.
f->
a .T
Exercise 5.2 Let A be an algebra, let E be an A-bimodule, and let D : A ----> E be a derivation. (i) Suppose that pEA is an idempotent. Prove that p . Dp . p = 0, and that Dp = 0 if, further, p . Dp = Dp . p. (ii) Suppose that a E A with a . Da = Da . a. Prove that D(a n ) = nan~l . Da for each n E N. Exercise 5.3 (i) Let A be a natural Banach function algebra on a non-empty, compact Hausdorff space K, and let x E K. Show that a linear functional on A is a point derivation at Ex if and only if d(l) = 0 and dUg) = 0 whenever 1,g E Mx = kerE x . (ii) Let C 1 (II) be the Banach function algebra of continuously differentiable functions introduced in Exercise 4.12(i), so that A is a natural Banach function algebra on II. Show that every continuous point derivation at EO has the form 1 f-> 001'(0) for some a E C. Also show that the space of discontinuous point derivation at EO is infinite-dimensional. (iii) Prove that all point derivations on the disc algebra A(~) are continuous. Exercise 5.4 Let A be a unital Banach algebra, and let E be an irreducible A-module. Extend Jacobson's density theorem by showing that, for linearly independent sequences (6, ... , ~n) and ("11, ... , "In) of elements in E, there is some a E A such that (expa) .
~k
=
"Ik
(k
= 1, ... ,n).
Representation theory
241
Automatic continuity There are remarkable results in which algebraic properties related to Banach algebras, Banach modules, and linear maps between them actually force the relevant map to be continuous. This is called the 'automatic continuity' of the linear map. A basic example of this phenomenon was given in Section 4.10: every character on a Banach algebra is automatically continuous. A particular aspect of automatic continuity theory concentrates on the 'uniqueness-of-norm' property for Banach algebras, and we shall now discuss thisi in the present section, we shall consider commutative Banach algebras, and then, in some later sections, we shall turn to general Banach algebras. The first such result is very easYi it is almost immediate from the fact that characters on Banach algebras are continuous. Theorem 5.25 Let A and B be Banach algebras, and suppose that B zs commutative and semisimple. Then every homomorphism from A into B is automatically continuous. Proof Let () : A ~ B be a homomorphism. We use the language of the separating space 6(()), which was introduced on Section 3.13 i by Theorem 3.51, we must show that 6( ()) = {O}. Thus, suppose that (an)n>l is a null sequence in A with ()(a n ) ~ bin B. Then, for every character tp on B, tp 0 () is either a character on A or is identically zeroi in either case, it is continuous. Thus tp(()(a n )) = (tp 0 ())(a n ) ~ 0 as n ~ 00. But also tp is continuous on B, so that tp(()(a n )) ~ tp(b) as n ~ 00. It follows that tp(b) = 0 for every tp E
242
Introduction to Banach Spaces and Algebras
5.5 Johnson's uniqueness-of-norm theorem. In this section, we shall prove a dramatic theorem of B. E. Johnson showing the uniqueness of the norm topology on every semisimple Banach algebra, extending the result for commutative Banach algebras that we have just proved; we shall give Johnson's now classical proof. In Section 5.10, we shall give an alternative proof of this theorem. The following is the main result of Johnson. Theorem 5.27 Let A be a Banach algebra with an identity, and let E be an irreducible normed A-module. Then E is a continuous A-module. Proof Let P be the kernel of the representation of A on E, so that P is a primitive ideal in A, and hence P is closed. Then AlP is a Banach algebra, and we may consider E as a faithful, irreducible AlP-module. Clearly, the continuity of the corresponding representation of AlP would imply the same for A. Thus, without loss of generality, we may suppose that E is a faithful A-module; i.e. we suppose that A ~ B(E) (of course, we do not hypothesize any relation between the norm of A and the operator norm on B(E)). Next, if dimE < 00, then dimB(E) < 00, and the result is trivial. Thus we suppose that E is infinite-dimensional. We are considering the normed space (E; 1·1); let (it; I ·1) be the completion of E. By Proposition 2.21, there is an isometric mapping j : U ~ j(U),
B(E)
--+
B(E).
For each ~ E E, define R~ : S ~ S~, B(E) --+ E, so that R~ is a continuous linear operator. We again set (Ta)(~) = a· ~ (a E A), and define e~ : a ~ a·~, A --+ E, so that e~ E .c(A, E) and R~ 0 T = e~, where T = joT. Set F = {~E E: e~ E B(A,E)}. Then F is clearly a subspace of E, and also, if ~ E F and b E A, then the map is the composition of the continuous maps a ~ ab on A and e~, so that b . ~ E F. Thus F is a submodule of E. Since E is irreducible, either F = E or F = {O}. First, assume towards a contradiction that F = {O}, and take ~ E E with ~ -=I- o. Then 6(e~) -=I- {O}. By Proposition 3.54(ii), 6(T) . ~ -=I- {O}. There is a linearly independent set {~n : n E N} of distinct elements in E. By Proposition 5.23, there is a sequence (a n )n2:1 in A such that an··· al . ~m = 0 for m, n E N with m < n and an··· al . ~n -=I- 0 for n E N. This implies that
eb. ~
6(T) . (Tan)··· (Tan)~n -=I- {O},
6(T). (Tan+l)··· (Tad~n = {O},
(*)
for each n E N. We now apply the stability theorem, Theorem 3.55(ii), taking each En of that theorem to be A, taking Rn : b ~ ban, so that Rn E B(A), taking F to be B(E), with Sn : U ~ _U 0 Tan, so that Sn E B(F), and taking T of that theorem to be the present T; we verify that TRn = SnT, as required.
Representation theory
243
Thus, by Theorem 3.55(ii), (6('T) . (Tan)··· (Tan))n>l is a nest which stabilizes, a contradiction of (*). Secondly, suppose that F = E, so that 0E E 8(A, E) for each ~ E E. For each a E A and ~ E E[l], we have IOE(a)1 = I(Ta)(~)1 ::::: IITall. By the uniform boundedness theorem, Theorem 3.34, there is M > 0 such that IIOd : : : M (~ E E[l]), and so IITal1 : : : M Iiall (a E A). Thus T is continuous, the required conclusion. The proof is complete. D Corollary 5.28 Let A and B be Banach algebras, and suppose that B is semisimple. Let 0: A -+ B be a homomorphism wzth O(A) = B. Then 0 is continuous. Proof We use the closed graph theorem. Indeed, take (an)n>l to be a null sequence in A with O(a n ) -+ b in B. Assume towards a contradiction that b "I- o. Then there is some irreducible representation T of B on a vector space E with Tb "I- o. By Corollary 5.3, E can be given a complete norm I . I such that T becomes a continuous, normed representation. But now ToO is a normed representation of A on E. Since O(A) = B, it follows that ToO is irreducible, and then, by Theorem 5.27, ToO is continuous. This implies that
Tb = lim T(O(a n )) = lim (T n----+oo
n----+oo
0
O)(a n ) = 0,
which is a contradiction. Thus b = 0, and so 0 is continuous by the closed graph theorem, Theorem 3.45. D The following corollary is Johnson's famous uniqueness-oE-norm theorem. Corollary 5.29 Let (A; unique complete norm.
11·11)
be a semisimple Banach algebra. Then A has a
Proof This follows from Corollary 5.28 in the same way as Corollary 5.26 followed from Theorem 5.25 in the commutative case. D
5.6 Eidelheit's theorem. Let E be a Banach space, and let Q( <;;; 8(E) be a Banach operator algebra on E, as in Example 4.6. Then Q( is a semisimple Banach algebra, and so Q( has a unique complete norm. We now give a delightful proof of Eidelheit from as long ago as 1940 that shows a somewhat stronger result in this special case. In the result, the operator norm on 8(E) is denoted by 11·11. Theorem 5.30 (Eidelheit's theorem) Let E be a Banach space, and let subalgebra of 8(E) that contains F(E).
Q(
be a
(i) Suppose that III . III is an algebra-norm on Q(. Then there is a constant C such that IITII : : : C IIITIII for each T in Q(. (ii) Any two complete algebra-norms on Q( are equivalent.
244
Introduction to Banach Spaces and Algebras
Proof (i) Assume towards a contradiction that there is no such constant C. Then there exists a sequence (Snk::':l in ~ such that IllSnll1 = 1 (n E N) and IISnl1 ---+ 00 as n ---+ 00. By the uniform bounded ness theorem, Theorem 3.34, there exists Xo E E such that the sequence (1ISnxoll)n>l is unbounded. By Corollary 3.35, there exists fo E E* such that the sequence (lfo(Snxo)l)n>l is unbounded; set Zn = fo(Snxo) (n EN). Now define T = Xo ® fo, so that T E ~ and T -1= o. Let n E N. Then (TSnT)(x)
= fo(x)(TSn)(xO) = znfo(x)xo = znTx (x
E
E),
and so znT = TSnT, whence IZnllllTll1 :::; IIITII12. Since IIITIII -1= 0, it follows that IZnl :::; IIITIII (n EN), a contradiction of the fact that the sequence (Zn)n21 is unbounded. (ii) Suppose that ~ is a Banach algebra with respect to III . 1111 and III . 1112' Take (Tnk:':l in ~ such that Tn ---+ 0 in (~; 111·1111) and Tn ---+ T in (~; 111.111 2), By (i), Tn ---+ 0 in (S(E); 11·11) and Tn ---+ Tin (S(E); 11·11). Thus T = 0, and so, D by Corollary 3.52,111'111 1 and 111·1112 are equivalent. 5.7 The continuity of derivations. Our aim in this section is to prove a theorem of Johnson and Sinclair on the automatic continuity of derivations on a Banach algebra. Recall that the notion of a nest which stabilizes was introduced on page 136. Let A be a Banach algebra, and let E be a Banach A-bimodule. Then E is a separating module for A if, for each sequence (a n )n21 in A, each of ( a1 ... a n . E) n21
and
(E· a n ... a1) n21
is a nest which stabilizes. In the case where E is a closed ideal in A, we speak of a separating ideal in A. Proposition 5.31 Let A be a Banach algebra. (i) Let B be a Banach algebra, and let e : B ---+ A be an epimorphism. Then 6(e) is a separating ideal in A. (ii) Let E be a Banach A-bimodule, and let D : A ---+ E be a derivation. Then 6(D) is a separating module for A. Proof Let (a n )n21 be a sequence in A. (i) Since e is an epimorphism, 6(e) is a closed ideal in A and there exists a sequence (b n )n21 in B such that e(b n ) = an (n EN). For n E N, define Rnb
= bbn (b
E
B)
and
Sna
= aa n (a
E
A).
Then (Rn)n21 and (Sn)n21 are sequences in S(B) and SeA), respectively, and eRn - Sne = 0 (n EN). It follows from Theorem 3.55(ii) that (al ... a n 6(e))n21 is a nest which stabilizes.
245
Representation theory
Similarly, (6(B)a n ··· at}n21 is a nest which stabilizes. (ii) Clearly, 6(D) is a closed sub-bimodule of E. For n
Rna = ana
(a E A)
and
Snx = an . x
E
N, define
(x E E),
so that (R n )n21 and (Sn)n21 are sequences in B(A) and B(E), respectively. Also,
(DRn - SnD)(a)
=
D(ana) - an . Da
=
Dan· a
(a
E
A),
and so DRn - SnD E B(A, E) (n EN). It follows from Theorem 3.55(ii) that (al ... an . 6(D))n>1 is a nest which stabilizes. Similarly, (6(D) . an··· at}n>l is a nest which stabilizes. D
Proposition 5.32 Let A be a Banach algebra with identity such that A zs a separating ideal in itself. Then AI J(A) is finite-dimensional. Proof Let E be an irreducible left A-module. By Corollary 5.3, E is a Banach left A-module for a suitable norm. Assume towards a contradiction that the space E is infinite-dimensional, say {~n : n E N} is a linearly independent set of distinct elements in E. By Proposition 5.23, there is a sequence (an)n>l in A with an+l ... al . ~n = 0 and an··· al . ~n =I- 0 for each n E N. Define In = Aan ··· al (n EN). Since Aan · .. al . ~n = E, we have In . ~n = E. Since Aan+l ... al . ~n = {O} and E is a Banach left A-module, I n+l . ~n = {O}. Thus I n+1 £.; In (n EN). This contradicts the fact that the nest (In)n>l stabilizes. Thus each primitive ideal in A has finite co dimension in A. Assume towards a contradiction that J(A) is not a finite intersection of primitive ideals of A. Then it is clear that there is a sequence (Pn)n>l of primitive ideals in A such that PI n· . ·nPn ~ Pk for each k > n in N, say Pk ~ Et (k EN), where (Ek)k>l is a sequence of finite-dimensional, irreducible Banach left Amodules. By -Proposition 5.24, for each n E N, there exists an E Ef n ... n E;, with an . Ek = Ek (k > n). Again, set In = Aan ··· al (n EN). For n E N, we have In+! <;;; Pn+!, and so In+l . En+l = {O}. However, an··· al . En+l = En+l' and so In . En+l = En+l · Thus In+l £.; In (n EN), a contradiction of the fact that (In)n>l stabilizes. Thus J(A) is an intersection of finitely many primitive ideals, each of finite co dimension in A, and so AI J(A) is a finite-dimensional algebra. D Theorem 5.33 (Johnson and Sinclair) Let A be a semisimple Banach algebra, and let D be a derivation on A. Then D is automatically continuous. Proof Set I = 6(D). By Proposition 5.31(ii), I is a closed ideal in A such that I is a separating ideal. By Corollary 5.12, I is semisimple, and I is a separating module in I. By Proposition 5.32, I is a finite-dimensional algebra.
246
Introduction to Banach Spaces and Algebras
Assume towards a contradiction that I =1= {a}. By Theorem 5.22, I contains a non-zero idempotent, say p. Let (a n )n21 be a null sequence in A such that Dan -+ pas n -+ 00. Then pan -+ 0 in I and
D(pa n ) = p(Da n ) + (Dp)a n
-+
p2 = P as
n
-+ 00.
However, D I I is continuous because I is finite-dimensional, and so p = 0, a contradiction. Thus 6(D) = {O}, and so D is continuous. 0 Corollary 5.34 Let A be a commutative, semisimple Banach algebra. Then there are no non-zero derivatwns on A.
Proof This is immediate from Corollary 5.8 and Theorem 5.33
o
Corollary 5.35 The algebras coo(lI) and O(U), where U is a non-empty, open subset of C, are not Banach algebras with respect to any norm. 0 Notes Johnson's proof [98] of his uniqueness-of-norm theorem, Corollary 5.29, answering a long-standing open question, was the seed from which much automatic continuity theory grew; see [47]. A second proof of the theorem will be given in Corollary 5.47, below. The theorem of Eidelheit, given as Theorem 5.30, is from [67]. It is natural to consider whether Johnson's uniqueness-of-norm theorem extends to Banach algebras with finite-dimensional radicals. An example, given as Exercise 5.7 below, will show that this is not always the case. The general case was investigated by Dales and Loy in [49], and further results and examples are given by Pham in [129]. Despite a substantial amount of work on automatic continuity and uniqueness-ofnorm questions, some obvious questions remain unanswered. For example, let A be a commutative Banach algebra which, as an algebra, is an integral domain, in the sense that ab oJ 0 whenever a and b are non-zero elements of A. Then it is not known whether or not A necessarily has a unique complete norm. Theorem 5.33 of Johnson and Sinclair is from [101]. For more general forms of Corollary 5.29 and Theorem 5.33, see [47, Theorem 5.2.28]. Other results, and some open questions, on the continuity of derivations are given in [23, 54]. It is an unproved conjecture that all derivations from a group algebra L 1 (G) into a Banach L 1 (G)-bimodule are automatically continuous. This is proved or stated for some groups G in [47, Section 5.6]; there seems to have been no further progress on this question. It is also an unproved conjecture that all derivations from an amenable Banach algebra A into a Banach A-bimodule are automatically continuous. There is a discontinuous derivation from the disc algebra A(~) into a Banach A(~)-bimodule, although all point derivations on the disc algebra are continuous (see Exercise 5.3). There is a generalization of this result in [22] and [47, Theorem 5.6.79]. A different technique of proving automatic continuity results flows from the main boundedness theorem of Bade and Curtis [21]. This leads to a theorem of Johnson [99]: for a Banach space E such that E is homeomorphic to its square (see Exercise 2.9), every homomorphism from 8(E) into a Banach algebra is automatically continuous. It also leads to a theorem of Bade and Curtis [21] that shows that a homomorphism from a commutative Banach algebra of the form C(K), where K is a compact Hausdorff space, into a Banach algebra is necessarily continuous on a dense subalgebra of C(K), and which gives a very precise description of such a homomorphism; the 'discontinuity' of e is 'concentrated on a finite singularity set'. It is a theorem of Esterle [74] that every
247
Representation theory
epimorphism from C(K) is automatically continuous. See [47, Section 5.4] for details and further results. The theorem of Bade and Curtis led to the conjecture that every homomorphism from C(K) into a Banach algebra is automatically continuous. However, it was proved independently by Dales [46] and Esterle [73] that, with CH, there is a discontinuous homomorphism from C(K) into certain Banach algebras whenever K is an infinite, compact Hausdorff space. It follows that, for each such K, there is an algebra-norm on C(K) that is not equivalent to the usual uniform norm; this answers a question of Kaplansky. The construction of such a homomorphism grew out of an earlier theorem of [10] that showed that the algebra J = iC[[X]] of formal power series in one variable can be embedded into certain Banach algebras; the latter construction used the notion of an element 'of finite closed descent' in a Banach algebra. For more on elements of finite closed descent, see [11-13]. Full details of these constructions and a characterization of the commutative Banach algebras that arise as the codomain of a discontinuous homomorphism from a maximal ideal of C(K) into a radical Banach algebra, are given in [47, Section 5.7]. Note that the above-mentioned construction of a discontinuous homomorphism requires the assumption of the continuum hypothesis, cn. It is a remarkable theorem of Woodin that this set-theoretic assumption is not redundant: there are models of set theory ZFC (in which CH does not hold) such that, in this model, all homomorphisms from C(K) into a Banach algebra are continuous. An account of this result is given in [51].
Exercise 5.5 Let w be a continuous weight function on JR+, and let L1(JR+,W) be the commutative Banach algebra defined in Example 4.67 and discussed in Exercise 4.24. Show that every derivation on L 1 (JR+ ,w) and every epimorphism from a Banach algebra onto L1 (JR+, w) is continuous, and hence that L1(JR+, w) has a unique complete norm. Indeed, let I be a non-zero, closed ideal in L1(JR+,W). Use Titchmarsh's theorem (given in the notes on page 206), to show that I cannot be a separating ideal in L1(JR+,W). Exercise 5.6 Let A be a commutative Banach algebra, and let E be a Banach Amodule. Set 2l = A EEl E, with the product specified by (a,x) . (b,y) = (ab, a . y
+b
. x)
(a,b E A, x,y E E)
and the norm specified by II(a,x)11 = lIall + Ilxll (a E A,x E E). (i) Show that (2l; II '11) is a commutative Banach algebra with J(2l) (ii) Let D : A ---> E be a derivation, and set
111(a,x)111 = lIall + IIDa - xii Show that (2l;
III . III)
=
J(A) EEl E.
(a E A,x E E).
is a Banach algebra.
(iii) Show that the norms continuous.
11·11
and
111·111
are equivalent on 2l if and only if D is
(iv) Use Exercise 5.3 to construct a commutative algebra with a one-dimensional radical which is a Banach algebra with respect to two non-equivalent norms.
248
Introduction to Banach Spaces and Algebras
Exercise 5.1 Let A be the sequence algebra (£2,11·112)' so that A is a commutative Banach algebra. Set 2l = A EEl IC as a vector space. For a, bE A and z, W E IC, define
(a,z) . (b,w) = (ab, 0)
and
II(a,z)11 = IIal12
+ Izl
.
Verify that 2l is a Banach algebra with respect to this product and norm, and that J(2l) is the one-dimensional space {a} EEl IC. Now let f be a linear functional on £ 2 such that f I £ 1 is the linear functional 00
f : a = (an) and set
f-+
L an , n=l
111(a, z)111 = max{ll a I1 2 , If(a) - zl} ((a,z) E 2l). Verify that (2l, 111·111) is also a Banach algebra and that the norm 111·111 is not equivalent to 11·11. A Banach algebra A is strongly decomposable if there is a closed subalgebra B such that A = B EEl J(A). Show that (2l, 111·111) is not strongly decomposable. This exercise is based on an example from 1951 of Feldman [77].
Variation of the spectral radius Let A be a Banach algebra. We have defined the spectrum, SPA a, and spectral radius, PA(a), of an element a of A in Section 4.5. In particular, we proved the fundamental theorem, Theorem 4.17, that SPA a is a non-empty, compact subset of C. We shall now study how this spectrum varies as the element a varies in the Banach algebra, and give some implications of this variation. 5.8 Continuity properties. The first point to notice is that the spectral radius function is upper semi-continuous on A. We shall give three simple proofs of this; we recall that upper semi-continuity was defined on page 18.
Proposition 5.36 Let A be a Banach algebra. Then PAis upper semi-continuous on A. Proof First proof. Let a E A, take c > 0, and set U = {A E C: 1,\1 < p(a) +c}, so that U is open and Sp a C U. By Proposition 4.21, there exists 8 > 0 such that Sp be U whenever lib - all < 8. But then p(b) < p(a) + c whenever lib - all < 8. Second proof. Let a E A, and take c > o. By Corollary 4.25, there is a norm III . III which is equivalent to II . II and such that (A; III . III) is a Banch algebra and Illalll < p(a) + c. But then, by the continuity of 111·111 at a, there exists 8 > 0 such that p(b) ~ Illblll < p(a) + c whenever lib - all < 8. Third proof. For each n E N, set fn(a) = IlanIl I / n . Then certainly each function fn : A -+ 1R+ is continuous on A. From Theorem 4.23, p(a) = infn fn(a). Then just this fact (Le. that p is the pointwise infimum of a collection of continuous
249
Representation theory
functions) implies the upper semi-continuity of p. For let a E A and E > O. Then there is some N EN such that fN(a) < p(a) + E/2. By the continuity of fN at a, there is then a 5 > 0 such that, for every b E A with lib - all < 5, we have p(b) ::; fN(b) < fN(a) + E/2 < p(a) + E. D As you will have guessed from the discussion of semi-continuity, the spectral radius function is not in general continuous. Of course, if A is a commutative Banach algebra, then PAis even uniformly continuous, since then
IPA(a) - PA(b)1 ::; PA(a - b) ::;
Iia - bll
(a,b E A).
An example showing that the spectral radius is not a continuous function on 8(£2) will be given in Exercise 5.lD.
5.9 Analytic properties. In this section we shall look at the following situation: A will be a Banach algebra, U will be an open subset of C, and f : U ---+ A will be an analytic A-valued function. We shall see that the function Zf---4PA(J(Z)) ,
U---+lR+,
has some very analytic-like properties, of which important use will be made. (In fact this function is 'subharmonic', and most of the properties that we shall prove in this section are true for subharmonic functions in general; but we emphasize that we still do not generally have continuity of the function-merely upper semi-continuity. ) We start with a trivial, but crucial, observation. Let A be a Banach algebra, and take a E A. As in the third proof of Corollary 5.36, above, let us write fn(a) = Ilanl1 1 / n for n E N. Now consider just the subsequence of N in which n runs through the powers of 2. Specifically, define gn = h n (n EN). Then of course gn(a) ---+ p(a) as n ---+ 00. However, in addition, (gn(a)k::':l is a decreasing sequence. This is just because 11·11 is a submultiplicative function, and so
gn+1(a) =
lIa2n+lIITcn+l) ::; lIa 2n II T (n+l) Ila 2" II Tcn +
1
)
= gn(a)
(n
E
N). '
We now need an elementary lemma which is a variant on Dini's theorem. Lemma 5.37 (Dini's lemma) Let K be a non-empty, compact Hausdorff space, and let (fn)n?l be a decreasing sequence in CIR(K)+. Set
f(x) Then
sUPxEK
fn(x)
---+ sUPxEK
=
lim fn(x)
n->oo
f(x) as n
(x
---+ 00.
E
K) .
250
Introduction to Banach Spaces and Algebras
Proof The sequence
(SUPK
fn)n?l is decreasing and
supfn2':supf2':O K
So L := limn--><Xl(suPK fn) exists, and L 2': En
(nEN).
K
=
SUPK
f. For n E N, define
{x E K : f n (x) 2': L} .
Since each f n is continuous, with sup K f n 2': L, we see that each En is compact and non-empty. But also the sequence (En)n?l is decreasing, and so, by the finite intersection property, nnEN En -I- 0. Take Xo EnnEN En. Then fn(xo) 2': L for all n E N. Hence f(xo) = limn--><Xl fn(xo) 2': L, and so SUPK f 2': L. This shows that SUPK f = L = limn--> <Xl (SUPK fn). 0 As a matter of interest, notice that the more familiar Dini's theorem is an immediate consequence of the above lemma.
Corollary 5.38 (Dini's theorem) Let K be a non-empty, compact Hausdorff space, and let (gn)n?l and 9 be continuous, real-valued functions on K such that gn(x) ---> g(x) monotonically for each X E K. Then gn ---> 9 uniformly on K. Proof We apply Dini's lemma to (fn)n?l, where fn
Ign -
=
gl
(n EN).
0
The first application of the above to Banach algebras gives a form of a maximum principle. Recall that oK denotes the frontier of a compact plane set K.
Theorem 5.39 Let K be a non-empty, compact subset of C, and take U to be an open neighbourhood of K. (i) Let E be a complex Banach space, and let f : U ---> E be an analytic function. Then Ilf(z)ll::::: sup Ilf(w)11 (z E K). wEaK
(ii) (Weak spectral maximum principle) Let A be a Banach algebra, and let f : U ---> A be an analytic function. Then
p(J(z))::::: sup p(J(w))
(z E K).
wEaK
Proof (i) Let z E K, and choose X E E* with II xii = 1 and X(J(z)) = Ilf(z)ll· Now X 0 f : U ---> C is an analytic function, so that, by the classical maximum modulus principle,
Ilf(z)11 = l(xof)(z)l::::: (ii) Apply (i) to the function we have
sup wEaK
l(xof)(w)l::::: sup IIf(wll· wEaK
f 2n , and take the 2n th root: for each n E N,
IIf(z)2 nII Tn
:::::
sup
Ilf(w)2n 11 2 -
n •
wEaK
Then, by Dini's lemma, Lemma 5.37, we deduce the required conclusion by letting n -+ 00. 0
251
Representation theory
As an immediate application we give the following characterization of the radical of a Banach algebra. Recall that J(A) and Q(A) denote the Jacobson radical and the set of quasi-nilpotents, respectively, of an algebra A.
Theorem 5.40 (Zemanek's characterization of the radical) Let A be a Banach algebra, and let a E A. Then the following are equivalent:
(a) a E J(A); (b) Sp (a + x) = Spx for every x E A; (c) p(a + x) = p(x) for every x E A; (d) p(a + x) = 0 for every x E Q(A). Proof We may suppose that A is unital. (a) =?(b). Let a E J(A), let x E A, and take A tf- Spx. Then x - A1 is invertible in A, and a + x - Al = (x - AI) (1 + (x - A1)- l a), which is invertible, using Theorem 5.9(ii). Thus A tf- Sp (a + x). The reverse inclusion also follows because x = (-a) + (a + x). The implications (b) =?(c) =?(d) are trivial. The main task is to show that (d) =?(a). Thus, let a E A be such that p(a + x) = 0 for every x E Q(A); in particular p(a) = 0 (taking x = 0), i.e. a E Q(A). Now let b E A be arbitrary, and define
F(z) = exp( -zb) a exp(zb)
(z E
q.
Then F is an A-valued entire function, F(O) = a, and F'(O) = ab - ba = [a, b]. Let
G(z) = {
F(Z)-a z
(z ~ 0),
[a, b]
(z
= 0).
Then G is also an A-valued entire function. Since Sp (F(z)) = Spa = {O} for all z E C, hypothesis (d) shows that p(G(z)) = 0 for all z E C \ {O}. By the weak spectral maximum principle, Theorem 5.39(ii), it follows that
p(ab - ba) = p(G(O)) = 0, i.e. that [a, b] E Q(A). We deduce from Corollary 5.18 that a E J(A). This completes the proof.
D
It should be noted that the use of the weak spectral maximum principle in the last argument seems to be quite essential; the semi-continuity of the spectral radius yields only the fact that p( G(O)) 2: o. For the further development of our analytic methods, we need a form of the three-circles theorem. For 0 < Rl < R 2 , define the closed annulus
Ll(Rl' R 2 ) := {z
E
C : Rl :s; Izl :s; R 2 }
.
Introduction to Banach Spaces and Algebras
252
Theorem 5.41 (Spectral three-circles theorem) Let A be a Banach algebra, let 0< Rl < R 2, let U be an open nezghbourhood of 6.(Rl' R 2 ), and let f : U ----7 A be an analytic function. For j = 1,2, set M J = sUPlwl=R, p(J(w)). Then p(J(z)) :::; MiMi- t where t E (0,1)
is
the unique number wzth
(Rl
< Izl < R 2),
Izl =
RiR~-t.
Proof Take z E C with Rl < Izl < R 2, and write r = 14 For p E Z and q E N, we apply Theorem 5.39(ii) to the function w f---+ IwIPp(J(w))q = p(w Pf(w)q) on a neighbourhood of 6.(Rl' R 2 ), and then take the qth root, so obtaining r P/ qp(J(z)) :::; max{ Rf/q M I , R~/q M 2 } .
(*)
Let 0: E lR be such that Rf Ml = R';f M 2 , and then apply (*) to a sequence of pairs (Pn, qn) E Z x N with Pn/qn ----70:. We deduce that rQ;p(J(z)) :::; RrMl = R~M2.
(**)
But log r = t log Rl + (1 - t) log R2 and 0: log Rl + log Ml = 0: log R2 + log M 2, and so we easily deduce from (**) that logp(J(z)):::; tlogMl + (l-t)logM2' which is equivalent to the stated result. 0
Corollary 5.42 Let A be a Banach algebra, take R > 1, and let f be an analytzc A-valued function on a neighbourhood of the annulus 6.(1/ R, R). Then (p(J(I)))2:::;
sup p(J(w))· sup p(J(w)). Iwl=l/R Iwl=R
Proof This is the case of the theorem in which Rl Izl2 = RIR2 = 1.
1/ R, R2
R, and 0
The first application of the theorem gives a Liouville-type theorem; we write the proof to show that, for this result, Corollary 5.42 is sufficient.
Corollary 5.43 Let A be a Banach algebra, and let f be an A-valued entire function. Suppose that p 0 f is bounded on C. Then p 0 f is constant. Proof Let M = sUPzEC p(J(z)) and m = inCEc p(J(z)) 2: 0; we may clearly suppose that M > o. Assume to the contrary that p 0 f is not constant. Then m < M, and so we may choose E > 0 such that m + E < M. Take Zo E C with p(J(zo)) < m + E. By considering f(z + zo) if necessary, we may suppose that Zo = o. By the upper semi-continuity of p 0 f at 0, there exists 8 > such that p(J(z)) < m + E whenever Izl :::; 8.
°
253
Representation theory
Let 0 =I- a E
(p(f(a))2::;
sup p(f(w))· sup p(f(w))::; (m + E)M. Iwl=lall R Iwl=Rllal
Thus (m+E)M is an upper bound for (p 0 J)2 on
p(f(z)) ::; p(f(zo)) Then p
0
(z
E
D).
f is constant on D.
Proof We first show that p 0 f is constant on a neighbourhood of zoo Let R > 0 be such that {z E
m = inf{p(f(z)) : Iz - zol < R/2}, take E > 0, and choose Zl E
0
f is
E = {z ED: p(f(z)) = p(f(zo))} = {z ED: p(f(z)) 2: p(f(zo))}. By the upper semi-continuity of p, the set E is relatively closed in D. By the argument of the last paragraph, E is open. Since Zo E E, we have E =I- 0, and so, by the connectedness of D, it follows that E = D. The theorem is proved. 0
254
Introduction to Banach Spaces and Algebras
5.10 Aupetit's lemma. We recall a definition from Chapter 3. Let E and F be Banach spaces, and let T : E - F be a linear map. As on page 135, the separating space SeT) of T is the set of all y E F such that TX n - y for some null sequence (Xn)n~l in E. It was noted in Theorem 3.51 that SeT) is a closed subspace of F and that T is continuous if and only if SeT) = {a}. Now suppose that A and B are unital Banach algebras and that e: A - B is a unital homomorphism with e(A) = B. Then it is easy to see that See) is a closed ideal of B. We remark also that, in the above case, certainly Sp B (ea) ~ Sp A(a), so that PB(ea) :::; PA(a), for each a E A. We give the following lemma for any linear mapping T : A - B for which PB(Ta) :::; PA(a) (a E A); by our remark, this includes the case in which T is a homomorphism. Theorem 5.45 (Aupetit's lemma) Let A and B be Banach algebras, and let T : A - B be a linear mapping such that PB(Ta) :::; PA(a) (a E A). Then, for every a E A and every b E SeT), we have
PB(Ta) :::; PB(Ta
+ b) .
Proof (Ransford) Let f :
PB(f(l)) :::; sup (PB(f(W)))t. sup (PB(f(W))/-t, Iwl=r
(*)
Iwl=R
where t = log Rj(log R - logr). Let a E A and b E SeT), say (an)n~l is a null sequence in A with Tan - b in B. Take E > O. By Corollary 4.25, we may, by changing to a norm on B equivalent to the original norm (which does not affect the conclusion), suppose that IITa + bll < PB(Ta + b) + E. For each n EN, apply inequality (*) to the function In, where
In(z) = T(a
+ an) - zTa n (z
E
certainly each In :
PB(fn(1)) :::; sup (PB (T(a
+ an) - zTa n )) t
. sup (PA (a
Izl=r
+ an - zan)) I-t
Izl=R
:::; (1IT(a
+ an)11 + rIITanll)t(lla + anll + Rllanll)l-t.
Since In(1) = Ta (n EN), we may let n -
PB(Ta) :::; (11Ta Now, for each fixed r > 0, let R t = log Rj(logR -logr) - 1, so that
00
to deduce that
+ bll + rllbll)tllaill-t. 00
and note that in this case we have
PB(Ta) ::; IITa + bll + rllbll Finally, let r - O. Then PB(Ta) ::; IITa every E > 0, so that the result follows.
(r > 0) .
+ bll < PB(Ta + b) + E.
This holds for 0
255
Representation theory
Corollary 5.46 (Generalized Johnson's theorem) Let A and B be Banach alge-
bras, and let T : A ----+ B be a linear s1trjection such that PB(Ta) :s; PA(a) (a E A). Then 6(T) s;:; J(B). In particular, T is continuous whenever B is semisimple. Proof Let b E 6(T), and let c E Q(B). Since T(A) = B, there exists a E A with Ta = -(b+c). By Theorem 5.45, PB(c+b) :s; PB(C) = O. By Theorem 5.40, (d)=}(a), bE J(B). In the case where B is semisimple, it follows that 6(T) = {O}, and so T is continuous. D Let A, B, and T : A ----+ B satisfy the conditions of the above theorem. Take bE 6(T). Then, for every sequence (bnk:~l in T(A) such that PB(b n - b) ----+ 0, it follows that PB(bn) ----+ O. But this does not imply that PB(b) = O. Corollary 5.47 Let A and B be Banach algebras with B semisimple, and let be an epimorphism. Then 0 is automatically continuous. D
o: A ----+ B
5.11 A notorious open problem. We would like to offer a few thoughts about a notorious unsolved problem in this area, often called the dense range problem. The problem is as follows:
Let A and B be Banach algebras, with B semisimple, and let 0 : A ----+ B be a homomorphism with dense range, so that O(A) = B. Is it true that 0 is automatically continuous? Remarks (i) By Theorem 5.25, the problem has a positive solution in the case where B is commutative. (ii) An equivalent formulation is to drop the requirements in the above statement that B be semisimple and that 0 have dense range, and ask the following question.
Let A and B be Banach algebras, and let 0 : A it always true that 6(0) s;:; Q(B)?
----+
B be a homomorphism. Is
It follows from Exercise 5.9, below, that, for each b E 6(0), the set Spb is connected, and, of course, 0 E Sp b. But this does not exclude the possibility that PB(b) > 0 because the function PB might not be continuous at b. The proof in Theorem 5.45 shows that 6(0) n O(A) s;:; Q(B). (iii) We may divide the problem into two parts: (a) show that kerO is closed; (b) with the additional hypothesis that ker 0 is closed, prove that 0 is continuous. Each of these two sub-problems appears to be open.
Elements analytically attached to subspaces. Let E be a complex Banach space, and let F be a vector subspace (not necessarily closed) of E. Then an element x of E is analytically attached to F if there is some open neighbourhood N of
o in
C and an analytic function f : N
----+
E such that: (i) f(O)
=
x; and (ii)
Introduction to Banach Spaces and Algebras
256
J(z) E F (z E N\ {O}). We shall denote by a(F) the set of all elements of E that are analytically attached to F. The following comments are almost immediate: (i) F
<;;;;
a(F)
<;;;;
F;
(ii) a(F) is a vector subspace of E, and, if A is a Banach algebra with F a subalgebra of A, then a(F) is a subalgebra of A; (iii) without loss of generality, we can always take N = D = {z E CC : Izl < I}, simply by replacing J(z) by J(l5z) for a sufficiently small positive 15. For example, let E be any infinite-dimensional, complex Banach space, and let F be a vector subspace that is the union of a strictly increasing sequence (En)n?l of closed subspaces; we remark that, by the Baire category theorem, Theorem 1.21, the vector subspace F is not closed. (For a specific example, take E = £2 and F = coo.) Now suppose that x E a(F). Then there is an analytic function J : D --+ E with J(O) = x and with J(z) E F for all zED \ {O}. But then there is some N E N such that J (z) E EN for uncountably many zED \ {O}. It is then elementary that J(z) E EN (z E D \ {O}). But EN is closed, and so also x = J(O) E EN <;;;; F. Hence a(F) = F. Let A be a complex, unital Banach algebra, and let a be element of A such that lIall ::; 1 and pea) -=I- 0 for each monic polynomial p. Define
J(z) = a(l - za)-l
(z
E
D).
It is an elementary exercise to show that {fez) : zED} is a linearly independent subset of A. Define F = lin{f(z) : 0 < Izl < I}. Then it is easy to see that a = J(O) E a(F) \ F. Thus a(F) -=I- F. Lemma 5.48 Let B be a unital Banach algebra, let F be a vector subspace of B, let a E a(F), and take c > O. Then
pea) ::; sup{p(b) : bE F, lib - all < c}. Proof We may suppose that a rt. F and that there is an analytic function J : D --+ B with J(O) = a and J(z) E F for 0 < Izl < 1. Choose 0 < 15 < 1 such that IIJ(z) - J(O) II < c whenever Izl < 15. By the weak maximum principle, Theorem 5.39(ii), we have
pea)
=
p(f(O)) ::; sup{p(f(z)) : Izl
=
15} ::; sup{p(b) : bE F, lib - all < c},
o
giving the result. Corollary 5.49 Let A and B be Banach algebras, and let homomorphism. Then 6(e) n a(e(A)) <;;;; Q(B).
e:A
--+
B be a
Proof Let b E 6(e) n a(8(A)), and take c > O. By Theorem 5.45, PB(C) < c for every c E e(A) with IIc - bll < c. Hence
PB(b) :S SUp{PB(C) : c E 8(A), IIc - bll < c} ::; c. Since this holds for every c > 0, we have PB(b) = 0, and so b E Q(B).
0
257
Representation theory
There is much that I do not know about the mapping F f---+ a(F) when F is a vector subspace of a Banach space or a subalgebra of a Banach algebra. In order to make use of Corollary 5.49 in approaching the dense range problem, it would be necessary to have a useful sufficient condition for an element to belong to a(F). Two specific questions are as follows (with F a subspace of a Banach space E):
Question 1. Is it ever the case that a( a( F))
-=I=-
a( F)?
a(F) = F? We remark, finally, that the definition of a(F) can be iterated transfinitely (though, obviously, this is not of interest unless the answer to Question 1 is 'Yes'). Let us say that a subspace F of a Banach space E is analytically closed if a(F) = F. Clearly, every closed subspace is analytically closed, while examples given above show that, for a non-closed subspace, both situations are possible. Question 1 is equivalent to asking whether a(F) is always analytically closed. It is clear that every intersection of analytically closed subspaces is itself analytically closed, and so there is a unique smallest analytically closed subspace of E that includes F: we shall denote that subspace by F"', and shall call it the analytic closure of F. It is clear that, for every subspace F of E, we have Question 2. Can it happen that F
F
~
-=I=-
a(F)
~
F'"
~
F.
We may also reach F'" 'from below' by transfinite recursion. We define aJ.L(F) for every ordinal JL as follows. (i) Let a1 (F) = a(F); (ii) for every ordinal JL, define aJ.L+1(F) = a(aJ.L(F)); (iii) define aJ.L(F) = U{av(F) : v < JL} for a limit ordinal JL. There is then a least ordinal, say v := v(F), such that av(F) is analytically closed. It is evident that av(F) = F"', as previously defined. If a(F) is not always analytically closed, then we may ask:
Question 3. Does there exist a non-closed subspace F for which F'" = F? The following is an easy extension of Corollary 5.49. Corollary 5.50 Let A and B be Banach algebras, and let 8 homomorphism. Then 6(8)) n 8(A)'" ~ Q(B).
A
--->
B be a o
Notes For more continuity and analytic properties of the spectrum, see [20, Chapter III, Section 4] and the earlier work [18]; this leads to the theory of analytic multifunctions, discussed in [20, Chapter VII]. For a general discussion of harmonic functions, see [136]. The earliest result in this area seems to be Vesentini's theorem, given in [47, Theorem 2.3.32]' which states that PA 0 f : U -> lR is a subharmonic function whenever A is a Banach algebra, U is a non-empty, open set in C, and f : U -> A is an analytic function. The idea of proving Johnson's uniqueness-of-norm theorem by the approach of the present section is contained in Aupetit's important paper [19]; see also [20, Chapter V, Section 5]. Ransford's beautiful simplification of the proof is in [135]. Here is a striking result of Aupetit [18] that is proved by the methods of this section; see [47, Theorem 2.6.28]. Let A be a semisimple Banach algebra, and suppose that there
258
Introduction to Banach Spaces and Algebras
is a real-linear subspace H of A such that A = H + iH and Sp a is finite for each a E H. Then A is finite-dimensional. Theorem 5.40 is due to Zemanek [169]; for an elementary proof, see [170]. The proof of Theorem 5.41 adapts one of the standard proofs of the Hadamard three-circles theorem; Corollary 5.42 is given in [135], and Theorem 5.45 is contained in [19]. It appears that the more special three-circles lemma of Ransford, Corollary 5.42, leads to the inequality PB(1'a? S PA(a)PB(1'a + b), which is slightly weaker than that of Theorem 5.45, but it should be remarked that this latter inequality is still sufficient for the deduction of Corollary 5.46. We shall sketch in Exercise 5.10 an example that shows that the spectral radius function PA is not necessarily continuous on a Banach algebra A. A more complicated example of Muller [123] exhibits a Banach algebra A with Q(A) = {O} such that the spectral radius function is discontinuous at each point of a line in A. In [137]' Ransford proved that there exist S, l' E A = B(£2) such that the map A f--4 PA(T - AS) is discontinuous at almost every point of the open unit disc; the argument avoids the combinatorial intricacies of Muller's example and uses some elementary potential theory. An example of Dixon [60], given as [47, Example 2.3.15], exhibits a semisimple Banach algebra A such that Q(A) <;; Q(A) = A, so that PA is not continuous on A; this semisimple Banach algebra contains as a dense, continuously embedded subalgebra a radical Banach algebra. The equivalent formulation of the dense range problem in the first remark, above is due to Albrecht and Dales in [3]; see also [47, p. 601]. Suppose that we assume (with B semisimple and T(A) = B) just that T : A -> B is linear with PB(Ta) S PA(a) (a E A). Then it is not the case that l' is automatically continuous: a counter example to this possibility is given in [47, p. 601]. It is also easy to see that, in that example, ker l' is not closed. But it appears that the question of whether or not it is always the case that 6(1') <;;; Q(B) is open in this more general case. Exercise 5.8 Show that the following conditions on a Banach algebra A are equivalent: (a) J(A) = Q(A); (b) Q(A) + Q(A) <;;; Q(A); (c) Q(A) . Q(A) <;;; Q(A). Exercise 5.9 Let A be a unital Banach algebra, and let a E A. Suppose that K is a non-empty, open-and-closed subset of Sp a. Prove that, for each open neighbourhood U of K in C, there exists 0 > 0 such that Sp (a + b) n U =1= 0 whenever b E A with Ilbll < O. Deduce that: (i) Sp a is a connected, compact set containing 0 whenever there is a sequence (a n )n2 1 in A such that an -> a and p(an ) -> 0 as n -> 00; (ii) the spectral radius PA is continuous at ao E A whenever Sp ao is totally disconnected. Exercise 5.10 We consider a weighted right shift operator l' on the Hilbert space H = £2, as defined in Exercise 4.10. The operator l' is defined by a sequence (w n ) in (0,1]' and then TOn = W nOn +l (n E N). To specify (W n )n21' note that each n E N has the form 2k(2£ + 1) for some uniquely specified numbers k,£ E Z+; in this case, set Wn = e- k . Then a calculation shows that, for each mEN, we have (WIW2'"
W2 m _l)1/(2"'-1)
> ( [ ( exp
C/+l ))
2
Representation theory
259
Set a = 2:;:1 j /2J+1. Then it follows from the above formula that the spectral radius p('1') satisfies p(T) :::: e- 2 0" > 0, and so the operator T is not quasi-nilpotent. By Exercise 4.10, SpT is a disc with centre 0 and with strictly positive radius. Next, for each k E fiI, define an operator 1", E B(H) by setting 1",e n = 0 when n = 2k(2L' + 1) for some L' E z:+ and 1"e n = 0 otherwise. Then Tk is nilpotent. However, liT - Tkll = e- k (k E fiI), and so 1" --> '1' in B(H). This shows that Q(B(H)) is not closed and that the spectral radius function p is not continuous on the algebra B(H) at the element T. Exercise 5.11 Show that the two formulations of the 'notorious open problem' in Section 5.11 are equivalent.
6
Algebras with an involution
Banach algebras with an involution 6.1
Let A be an algebra over
General Banach *-algebras.
rc.
Then an
involution on A is a mapping
* :a
f---+
a*,
A
-->
A,
such that:
= a for all a E A; (ii) (>.a + p,b)* = "xa* + "jib* for all a, bE A and >., p, (iii) (ab)* = b* a* for all a, b E A. (i) a**
E C;
An algebra with an involution is a *-algebra. Note that, immediately from (ii), 0* = 0 and, from (iii), that, if A has an identity 1, then 1* = 1. It then follows that, if A has no identity, then the involution on A is uniquely extendible to an involution on A+. So, we may suppose without loss of generality that A has an identity. Let A be a *-algebra. A subset S of A is *-c1osed or a *-subset if a* E S whenever a E S. A *-closed subalgebra or ideal in A is a *-subalgebra or *-ideal, respectively. Note that a *-closed left ideal in A is an ideal. A homomorphism () : A --> B between *-algebras is a *-homomorphism if (()a)* = ()(a*) (a E A). Suppose that A is a Banach algebra. Then an involution * on A is isometric if Ila*11 = II all (a E A). The algebra (A; *) is a Banach *-algebra if * is an isometric involution on A. Example 6.1 The following are all examples of Banach *-algebras, as is easily checked. (i) Set A = Co(K), the algebra of all complex-valued, continuous functions which vanish at infinity on a non-empty, locally compact Hausdorff space K, and define
j*(x) = f(x)
(f
E
A, x
E
K),
so that j* = 7 in a previous notation. (ii) Set A = A(,6.), the disc algebra, as in Example 4.3, and define
j*(z) = f(z) (iii) Set A
(f E A, z E ,6.).
= (Ll(JR.); 11.11 1 ), as in Example 4.65, and define j*(t)
= f( -t) (f
E
A, t
E
JR.).
Algebras with an involution
261
(In some notations, this 1* is denoted by /.) Note that F(f*) = F(f) for each and so the Fourier transform F is a *-homomorphism from Ll(lR) into Co(IR).
f E £1 (1R),
The above examples (i), (ii), and (iii) are all commutative algebras. The most important non-commutative example is the following. (iv) Set A = B(H), the algebra of all bounded linear operators on a Hilbert space H, with the involution taken as the Hilbert-space adjoint; see Corollary 2.56. Then A is a Banach *-algebra. A *-representation of a *-algebra A on a Hilbert space H is a *-homomorphism () : A ---. B(H). Our main aim in this section is to construct *-representations of arbitrary Banach *-algebras. (v) Any closed sub algebra of B(H) that is invariant under taking the adjoint is a Banach *-algebra. In particular, let T E K(H), the operator algebra of all compact linear operators on H. By Corollary 3.66, the adjoint T* is compact, and so K(H) is a Banach *-algebra. (vi) As a special case of the above, let A = (a'J) E MIn, the algebra of all n x n-matrices over C. Then the adjoint A* of A is (aJ ,). (The matrix (aJ ,) is sometimes called the conjugate transpose matrix of (a'J)') (vii) Let A be a Banach *-algebra. Suppose that J is a closed *-ideal of A, and set (a + J)* = a* + J (a E A). Then (AjJ; *) is a Banach *-algebra. In particular, for a Hilbert space H, the so-called Calkin algebra B(H)jK(H) is a Banach *-algebra. D On page 76, we defined 'self-adjoint' (or 'hermitian'), 'normal', and 'unitary' for operators on a Hilbert space. We now give more abstract versions of these definitions. Let A be a *-algebra, and let a E A. Then a is self-adjoint or hermitian if a = a*, and normal if a*a = aa*; in the case where A has an identity 1, a is unitary if a*a = aa* = 1. The set of self-adjoint elements of A is a real-linear subspace of A; it is denoted by Asa. In the case where A is commutative, Asa is a real subalgebra of A. The set of unitary elements is denoted by U(A). In the case where A is a Banach *-algebra, Asa and U(A) are closed in A. Proposition 6.2 Let A be a unital *-algebra, and let a E A. Then a is invertible if and only if both a*a and aa* are invertible; if a is normal, then a is invertible if and only if a*a is invertible.
Proof Suppose that a E G(A). Then a* E G(A), and hence a*a, aa* E G(A). Conversely suppose that a*a, aa* E G(A), say b = (a*a)-l. Then ba*a = 1, and a has a left inverse. Similarly, a has a right inverse, and so a E G(A). The remark about a normal element a is immediate, since then a*a = aa*. D Let A be a *-algebra. Then each a E A can be written uniquely as a = b + ic, with b, c E A sa , by taking b = (a + a*)j2 and c = (a - a*)j2i. It follows that A = Asa EB iA sa . Note that a = b + ic is normal if and only if bc = cb. It is also a
262
Introduction to Banach Spaces and Algebras
simple, purely algebraic, fact that (in the case where A is unital), a is invertible if and only if a* is invertible, in which case (a*)-l = (a- 1 )*. Hence Sp(a*)
= {:-\: A E
Spa} .
Thus, if a E A sa , then Sp a is symmetrical about the real axis. (N.B.: it does not follow that Sp a <;;; JR., as is shown by considering the element Z in the disc algebra, with the involution as in Example 6.1(ii); here Z is self-adjoint, but SpZ = fl.) In the case where A is a Banach *-algebra, p(a*) = p(a) (a E A). Lemma 6.3 Let A be a *-algebra. Then its radzcal J(A) is a *-ideal, and so AI J(A) has a natural involution. Proof We may suppose that A is unital. Let a E J(A), and take b E A. Then (1 + ba*)* = 1 + ab*, which is invertible by Theorem 5.9(iii). By the above remarks, 1 + ba* is also invertible, so that a* E J(A) by Theorem 5.9(ii). Since a** = a, the converse holds. 0 Theorem 6.4 (Johnson) Let A be a semisimple Banach algebra with an involution. Then the involution is continuous on A, and A is a Banach *-algebra for an equivalent norm. Proof Let 11·11 be the norm on A. The theorem is immediate from Corollary 5.29 because the formula
Illalll := max{llall, Ila*11}
(a E A)
defines a second Banach algebra-norm on A, and this must then be equivalent to 11·11. Clearly, (A; 111·111) is a Banach *-algebra. 0 Let A be a Banach algebra, and let a E A be such that Sp a n JR.- = 0. Recall that we showed in Theorem 4.99 that there is a unique element bE A such that both b2 = a and Sp b c II. The element was called the 'principal square root of a'. There is a simple further point to be made in the case of an algebra with an involution; note that we do not suppose that the involution is continuous in the following result. Again, we set D =
= a* = a, and also
Sp b* = {X : A E Sp b} <;;; II . Hence, by the uniqueness of the principal square root, b* = b.
o
Algebras with an involution
263
Proposition 6.6 Let A be a Banach algebra with an involution, and let a be a normal element of A. Then there is a closed, commutative *-subalgebra C of A, containing a and such that Spca = SPA a. Proof We may suppose that A is unital. Take C = {a, a*}Cc. Since {a, a*} is a commutative subset of A, it follows from Corollary 4.42 that C is a closed, commutative subalgebra of A such that Spca = SPA a. Clearly, a E C and C is *-closed in A. 0 6.2
Positive linear functionals.
A+
Let A be a *-algebra. Then we define
= {ta;aJ: a1, ... ,an
E
A, n E
J=l
N} .
Clearly, (A + , +) is a semigroup and aA + <;;; A + for each a we have 4ab
E
jR+. For a, b E A,
= (b+a*)* (b +a*) - (b - a*)* (b - a*) + i(b+ia*)* (b+ ia*) - i(b- ia*)* (b - ia*) ,
and so A2 = linA+. A linear functional f on a *-algebra A is positive if and only if f(a*a) :::: 0
(a E A).
In this case, f(A+) <;;; jR+. A positive linear functional f on a unital *-algebra A is a state if f(l) = 1; the set of states on A is the state space of A, and it is denoted by S(A). For example, every character on C(K) for a compact Hausdorff space K is a state, and the map T f---7 (Tx, Xl is a positive linear functional on 8(H) for each x E H; the latter map is a state if and only if Ilxll = l. Note that, for each b E A and each positive linear functional f on a *-algebra A, the functional a f---7 f(b*ab) is also a positive linear functional on A. Proposition 6.7 Let A be a *-algebra, and let f be a positive linear functional on A. Then: (i) f(a*) = f(a) (a E A2);
(ii) f(a*b) = f(b*a) (a, bE A); (iii) (Cauchy-Schwarz inequality) If( a* b) 12 ::; f( a* a )f(b* b) (a, b E A) ; (iv) in the case where A is unital, If(a)1 2 ::; f(l)f(a*a) (a E A). Proof (i) Let a E A2, say we have a = 'L;=1 AJa;aJ , where A1,'" An E C and al,··.,an E A. Then a* = 'L;=lXJa;aJ, and we know that f(a;a J ) :::: 0 for j = 1, ... , n. The result follows. (ii) This is a special case of (i).
264
Introduction to Banach Spaces and Algebras
(iii) Take a,b E A. Then f((Aa + p,b)*(Aa + p,b)) :::: 0 for each A,p, E Co In particular, take p, = f(a*b) and A = t E lR. Then
t 2f(a*a)
+ 2t If(a*b)1 2 + If(a*b)1 2 f(b*b)
:::: 0,
and this implies the Cauchy-Schwarz inequality. (iv) This is a special case of (iii).
o
Proposition 6.8 Let A be a unital Banach *-algebra, and let f be a positive
linear functional on A. Then: (i) If(a*ba)1 ::; f(a*a)p(b) (a E A, bE Asa); (ii) If(a*ba)1 ::; f(a*a)p(b*b)1/2 (a, bE A); (iii) If(a)1 ::; f(l)p(a*a)1/2 ::; f(l) Iiall (a E A); (iv) If(a*a)1 ::; f(l)p(a*a) ::; f(l) IIal1 2 (a E A); (v) f is continuous, wzth Ilfll = f(I), and f(A+) C
jR+.
Proof (i) Let a E A and b E Asa; we may suppose that p(b) < 1, and so Sp (1 ± b) ~ TI. By Proposition 6.5, there exist y, Z E Asa with (1 - y)2 = 1 - b and (1 - Z)2 = 1 + b. Set u = a - ya and v = a - za. Then u*u = a*a - a*ba and v*v = a*a + a*ba. Thus f(a*a) ± f(a*ba) :::: 0, and so If(a*ba)1 ::; f(a*a). (ii) Let a, b E A. By the Cauchy-Schwarz inequality,
If(a*ba)12 ::; f(a*a)f(a*b*ba), and so the result follows from (i). (iii) The first inequality is a special case of (ii); for the second, note that
p(a*a) ::; Ila*all ::; Ila11 2 . (iv) This follows from (iii) because p(a*aa*a) = p(a*a)2 (a E A). (v) Let a E A. Then If(a)1 ::; f(l) Iiall by (iii), whence Ilfll ::; f(l). Since Ilfll :::: f(l), we have Ilfll = f(l). Since f(A+) ~ jR+ and f is continuous, it follows that f(A+) C jR+. 0 Corollary 6.9 Let A be a unital Banach *-algebra. Then
S(A) = {f
E
A* : f(A+) ~
Let A be a *-algebra, and let define
jR+,
Ilfll =
f(l) = I}.
o
f be a positive linear functional on A. Then we
L j = {a E A: f(a*a) = O}. It is clear from the Cauchy-Schwarz inequality that L j is a left ideal in A and that f(ba) = f(a*b) = 0 (a E L j , bE A). Suppose that A is unital. Then we define the *-radical of A by J*(A) = n{L j
:
f
E
S(A)}.
The *-algebra A is *-semisimple if and only if J*(A) = {o}.
265
Algebras with an involution
Proposition 6.10 Let A be a unital *-algebra. Then: (i) J*(A) = n{ker f : f E S(A)}; (ii) J*(A) is a *-ideal in A; (iii) Aj J* (A) is *-semisimple. Proof (i) Take a E J*(A). For f E S(A), we have a E ker f by Proposition 6.7(iv). Take a E n{ker f : f E S(A)}. For f E S(A) and ,\ E C, we have f(('\l
+ a)*a('\l + a))
=
0,
and so '\f(a*a) +).f(a 2 ) = 0. By taking'\ = 1 and'\ = i, we see that f(a*a) = 0, and so a E Lt. Thus a E J*(A). (ii) Certainly J*(A) is a left ideal in A. Let a E J*(A), and take f E S(A). Then the linear functional g : b 1-+ f(aba*) is also positive, and so f(aa*aa*) = 0, whence a* E Lg s;;: J*(A). Thus J*(A) is *-closed, and hence a right ideal.
o
(iii) This follows easily.
Proposition 6.11 Let A be a unital Banach *-algebra. Then J*(A) is a closed ideal in A, AjJ*(A) is a *-semisimple Banach algebra, and J(A) s;;: J*(A). Proof It follows from Proposition 6.8(iv) that ker f is closed for each f E S(A), and so J*(A) is closed. Thus AjJ*(A) is a *-semisimple Banach algebra. Let a E J(A). Then a*a E J(A) s;;: Q(A), and so f(a*a) = (f E S(A)) by Proposition 6.8(iv). Thus a E J*(A). 0
°
Corollary 6.12 Let A be a unital Banach *-algebra such that A is *-semisimple. Then A is semisimple. 0 Theorem 6.13 (Kelley and Vaught) Let A be a unital Banach *-algebra. Then J*(A)
=
{a E A: -a*a E A+ } .
Proof Take a E A with -a*a E A+, and let f E S(A). Then f(a*a) s;;: lR- nlR+ by Proposition 6.8(v). Thus f(a*a) = 0, and a E Lt. It follows that a E J*(A). Conversely, suppose that a E J* (A). For a E Asa , let fJ,(a)
= inf{lla + bll : b E
A+}
= d( -a, A+).
Then fJ, is a sublinear functional on the real vector space Asa , and so, by the Hahn-Banach theorem, Theorem 3.1, there is a real-linear functional f on Asa with f(a*a) = fJ,(a*a) and f(b) :::::: fJ,(b) (b E Asa). Extend f to be a complexlinear functional, also called f, on A. For b E A, we have - f(b*b) :::::: fJ,( -b*b) = 0, and so f(b*b) ;::: 0. This shows that f is a positive linear functional, and so f(a*a) = because a E J*(A). Thus fJ,(a*a) = 0, and so -a*a E A+. 0
°
266
Introduction to Banach Spaces and Algebras
6.3 The GNS construction. We now give the so-called GNS-construction for Banach *-algebras; the letters G, N, S honour Gel'fand, Naimark, and Segal, who developed the theory. Let A be a unital Banach *-algebra, and let J E S(A), so that L f is a closed left ideal in A. We denote by 7rf : A ----+ AIL f the quotient mapping, and consider the equation (7r f (a), 7rf ( b)) f = J (b* a) (a, b E A) . (* ) Suppose that aI, a2 E A with al - a2 ELf. Then J(b*aI) But also, if bl , b2 E A with bl - b2 ELf, then J(bra)
=
J(a*b l )
=
J(a*b2)
=
JW2a)
=
J(b*a2) (b E A).
(a E A).
Thus the formula (*) gives a well-defined value to (7r f (a), 7rf (b)) f' It is then clear that ( " . ) f is an inner product on AIL f; the corresponding norm on AIL f is denoted by 11·lI f , so that l17rf(a)lI~ = J(a*a) (a E A), and the Hilbert space which is the completion of (AILf; 1I·lI f ) is called Hf; see Section 2.12. For a E A, define a mapping Tfa: AILf ----+ AILf by (Tfa)(7rf(b)) = 7rf(ab)
(b E A).
That is, we make AILf into a left A-module in the natural way, as in Section 5.1 In this notation, we have the following 'single representation' result, corresponding to a fixed J E S(A). Lemma 6.14 (i) For each a E A, the map Tfa is a bounded lmear operator on (AI L f ; 1I·lI f ), and so extends uniquely to a bounded linear operator (also denoted by Tfa) on H f .
(ii) The mapping Tf : a and IITfall :::::
lIall
f-4
Tfa, A ----+ B(Hf), is a unital *-homomorphism
(a E A).
(iii) kerTf = Lf : A = {a E A: J(b*a*ab) = 0 (b E A)} ~ L f . (iv) J*(A)
= n{kerTf
: J E S(A)}.
Proof (i) For a, bE A, we have II(Tfa)(7rf(b))IIJ = J(b*a*ab). For b E A, the function g : a f-4 J(b*ab) is a positive linear functional on A, and so, using Proposition 6.8(iv), we have J(b*a*ab)
=
g(a*a) ::::: g(1)lIaIl 2 = J(b*b)lIaIl 2 = l17rf(b)IIJ
lIall 2 .
Hence II(Tfa)(7rf(b))lIf = g(a*a)I/2::::: l17rf(b)lIf lIall, and so Tfa E B(AILf)· (ii) Clearly, the map Tf is a unital homomorphism with IITfall ::::: lIall (a E A), and so IITf II = 1. Now take a, b, c E A. Then ((Tfa)(7rf(b)), 7rf(c)) f
and so (Tfa)*
= Tf(a*).
=
J(c*ab)
=
J((a*c)*b)
=
(7rf(b), (Tf(a*))(7rf(c))) f'
Thus T f is a *-homomorphism.
267
Algebras with an involution
(iii) This is immediate from the definition of T f . (iv) Let a E kerTf . Then a ELf, and so n{kerTf : f E S(A)} ~ J*(A). Conversely, let a E J*(A), and take f E SeA). Then f(b*a*ab) = 0 (b E A), and so a E kerTf . Thus J*(A) ~ n{kerTf : f E S(A)}. 0 The final step in the proof of our main representation theorem is to fit together the homomorphisms Tf for f E SeA). For this we need to be able to form a 'Hilbert-space direct-sum' of a collection of Hilbert spaces. The basic idea is quite simple, and we shall be brief over some details. Let (H).,; 11·II).,hEA be a collection of Hilbert spaces. Then the Hilbertian sum, H = L).,EA H)." of these spaces is the space H consisting of all those families
{ (x).,) E
IT H)., : L)., Ilx).,ll~ < oo} .
).,EA
It is an exercise to prove that, with respect to the obvious coordinatewise operations and with inner product given by (x, y)
=
L
(x)." Y).,).,
(x
=
(X).,».,EA, Y
= (Y).,».,EA)'
).,EA H becomes a Hilbert space; by Proposition 2.45, l(x)",y)")"1 ::::: Ilx)"lllly)"11 for oX E A, and so the series used to define the inner product is absolutely summable by Holder's inequality, Theorem 2.8(i). Let A be a unital Banach *-algebra, and suppose that T)., is a unital *representation of A on H)., with liT)., II = 1 for each oX E A. For a E A, define
(Ta)((x).,».,EA)
=
(T).,a(x)"»)"EA
((X).,».,EA E H).
Then Ta E B(H) with IIT).,all ::::: IITal1 ::::: Iiall (oX E A), and the map T is a unital *-representation of A on H with IITII = 1. It is called the direct sum of the *-representations T)." and is denoted by T = E8{T)., : oX E A}. Now let Hf and Tf be as above for f E SeA), and set H = LfES(A) Hf· Then T = E8{Tf : f E S(A)} is a unital *-representation of A on H with IITII = 1. Let a E A. It is clear that Ta = 0 if and only if Tfa = 0 (f E SeA»~, and so kerT
=
J*(A)
=
n{Lf : f E S(A)}.
Hence T is an injection if and only if A is *-semisimple. Let A be a unital *-algebra. Then a *-representation T of A on a Hilbert space H is faithful if it is injective, and universal if T is unital and each state on A has the form a ~ (Ta(x), x) for some x E H with Ilxll = 1. Let T be a unital *-representation of a unital Banach *-algebra on a Hilbert space H, and take x E H with Ilxll = 1. Then f : a ~ (Ta(x), x) f is a state on A and IITa(x)11 2 = f(a*a) ::::: lIa11 2 , and so IITII = 1. Suppose that T is faithful and that a E J*(A). Then Ta = 0, and so a = O. Thus A is *-semisimple. By combining the above remarks we establish the following theorem.
268
Introduction to Banach Spaces and Algebras
Theorem 6.15 Let A be a unital Banach *-algebra. Then there is a Hilbert space H and a universal *-representation T of A on H with IITII = 1. Further, there is a faithful universal *-representation if and only if A is *-semisimple. 0 Notes A magisterial account of *-algebras and Banach *-algebras is the two-volume work of Palmer [126, 127J. For a brief introduction, see [47, Sections 1.10, 3.1J and [140, Chapter 4J. There is a variation in the terminology regarding Banach *-algebras: the point is that a Banach algebra with an involution is not necessarily a Banach *algebra. For example, for Palmer [127], the term 'Banach *-algebra' does not imply that the involution is continuous. There is an example in [47, Theorem 5.6.83], from [45], of a Banach algebra A with a discontinuous involution * and an element a E A such that (expa)* =1= exp(a*). For a study of topological algebras with an involution, see [78J. An involution * on a unital algebra A is symmetric if 1 + a*a E G(A) for all a E A and hermitian if Sp a <;;; lR for each a E Asa. A symmetric involution is always hermitian. In the case where A is a Banach algebra, the converse is true, and this holds if and only if p(a)2 ::; p(a*a) (a E A); see [47, Corollary 3.1.11J. There seems to be no standard notation for what we have called J*(A), the *radical. It is called the pre-reducing ideal in [127, Definition 9.7.14J. It seems to be unknown whether or not J(A) <;;; J*(A) for every *-algebra A. The theorem of Kelly and Vaught is from [108J. For more on Hilbertian sums, see [102, Section 2.6J. The GNS construction is extensively discussed in the literature. Exercise 6.1 Let A be an infinite-dimensional Banach space, and define ab = 0 for a, b E A. Then A is a commutative Banach algebra. Let {Xn : n E N} U {y, : 'Y E r} be a basis of A consisting of elements of norm 1. Define
y; = y,
('Y E
r) ,
* X2n
= nX2n-1 ,
X;n-1
= X2n/n
(n
E N),
and extend * by conjugate-linearity to A. Verify that * is a discontinuous involution on A. Exercise 6.2 Let A be a Banach *-algebra. Show that (expa)* = exp(a*) (a E A). Exercise 6.3 Let (H; ( . , . )) be a Hilbert space, and suppose that S, T E .c(H) are such that (Tx, y) = (x, Sy) (x,y E H). Use the closed graph theorem to prove that both Sand T are continuous. Thus it is sensible to define T* for T E .c(H) only if T E B(H). Exercise 6.4 Let G be a group. We defined the group algebra (f 1 (G); Example 4.4. Now define
* : '~ " f(8)8 5 sEC
f--> " ~ ' - 5 -1,
f(8)8
f 1 (G)
-+
II . III ; *)
in
f 1 (G).
sEC
Verify that * is an involution on f1(G) and that (f1(G); algebra.
11·111;
*; *) is a Banach *-
269
Algebras with an involution
Exercise 6.5 Let §2 be the free semigroup on two generators, u and v, so that elements of §2 are 'words' in u and v. Show that there is a unique involution U on (l1(§2) such that 8t = 8v and 8t = 8u and such that ((l1(§2); 11.11 1 ; 0 is a Banach *-algebra. For each word w E §2, let nu (w) and nv (w) be the number of times that the letters u and v, respectively, occur in the word w. Show that, for each (1, (2 E ~, the continuous linear functional 'P defined by requiring that 'P(8 w ) = (~,,(w)Gv(w) for each word w is a character on (11 (§2), and that each character arises in this way. Deduce that 8uv is self-adjoint, but that Sp 8uv :;;> ~.
C* -algebras 6.4 Elementary theory of C* -algebras. nach algebra with an involution such that
Ila*all =
IIal1 2
A C *-algebra is a (non-zero) Ba-
(a E
A).
The above equality is called the C *-condition. Let A be a C *-algebra. For each a E A, we have IIal1 2 =
Ila*all : : :
Ila*llllall,
and so Iiall ::::: Ila* II. Since a** = a, we deduce that Ila* II = Iiali. Thus A is a Banach *-algebra. Suppose that A has an identity 1. Then 111112 = 111*111 = 11111, so that 11111 = 1, i.e. a C* -algebra with an identity is necessarily a unital Banach algebra. Further, for each unitary u, we have IIul1 2 = Ilu*ull = 11111 = 1, and so Ilull = 1. Let A be a C *-algebra. Then a *-subalgebra of A which is closed in the norm topology of A is called a C *-subalgebra of A. Examples 6.16 (i) Let K be a locally compact Hausdorff space, and consider CoCK), as in Example 6.1(i). It is clear that IfllK = Ifl~ (f E CoCK)), and so CoCK) is a commutative C*-algebra. (ii) Let 8 be a non-empty set, and let X be a non-empty topological space. Then (£00(8); I· Is) and (Cb(X); I· Ix) are commutative C*-algebras: the involution is * : f 1-+ ] . (iii) Let H be a Hilbert space, and consider l3(H); the involution is the map T*, where T* is the Hilbert-space adjoint of T E l3(H), as in Example 6.1(iv). That the C*-condition holds for l3(H) was verified in Proposition 2.57.
T
1-+
(iv) Each C *-subalgebra of a C *-algebra is itself a C *-algebra. In particular, each closed *-subalgebra of l3( H) for a Hilbert space H is a C *-algebra. This includes the example lC(H). D We shall see in the following section that every C *-algebra is (isometrically *isomorphic to) a C *-subalgebra of l3( H) for some Hilbert space H. In the present section, we shall prove the easier theorem that every commutative C *-algebra has the form CoCK) for some non-empty, locally compact Hausdorff space K.
270
Introduction to Banach Spaces and Algebras
Lemma 6.17 Let a be a normal element of a C*-algebm A. Then p(a)
= Iiali.
Proof First, suppose that a E A sa , so that a = a*. Then IIa 2 11 = IIal1 2 by the C*-condition; a simple induction then gives Ila 2 " II = Ila11 2 " (n EN), and so
Iiall = Ila 2 " liT"
--->
p(a)
as
n
---> 00
by the spectral radius formula, Theorem 4.23. Thus p(a) = Iiali. For each a E A, a*a is self-adjoint. Suppose that a is normal. Then, by Corollary 4.48(ii), p(a*a) p(a*)p(a), and so
s:
IIal1 2 = Ila*all = p(a*a)
s: p(a*)p(a) = p(a)2 s: lIa11 2 .
Thus p(a) = Iiall, as required.
o
Corollary 6.18 Every C*-algebm is semisimple. Proof Let A be a C *-algebra, and take a E J(A). By Lemma 6.3, a* E J(A), and so also b = (a + a*)/2 and c = (a - a*)/2i belong to J(A). But then band c are quasi-nilpotent as well as self-adjoint, and so Ilbll = p(b) = 0 and IJeII = o. Thus b = c = 0, and so a = o. Hence J(A) = {O}. 0 Corollary 6.19 Let A and B be C*-algebms, and let e : A ---> B be a *-homoII all (a E A). morphism. Then e is continuous and, moreover, Ileall
s:
Proof Let a E A. Then a*a, and hence (ea)*(ea), are normal, and so, by Lemma 6.17, we have
Ileal1 2 = II(ea)*(ea)11 = PB((ea)*(ea)) = PB(e(a*a))
s: PA(a*a) = Ila*all = Ila11 2 , o
giving the result.
It follows that a C *-algebra has a unique complete norm in a strong sense: if A is a C*-algebra with respect to 11·11 and 111·111, then II all = Illalll (a E A).
We can now establish the famous Gel'falld~Naimark theorem for commutative C *-algebras; we recall that the Gel'fand representation theorem for commutative Banach algebras was given in Theorem 4.59. We shall require the following preliminary results, in each of which A is a unital C *-algebra with identity 1. Lemma 6.20 Let f E A* with Ilfll = f(l). Then f(a) E IR whenever a E Asa· Proof Without loss of generality, we may suppose that Ilfll
= f(l) =
1.
271
Algebras with an involution
Take a E A sa , and set f(a) = 0: + i,6, with 0:,,6 E R Now define
u(t)
=
tl - ia
(t E JR).
Let t E R Then we have Ilu(t)11 2 = Ilu(t)u(t)*11 = IIt 21 + a 2 11 ::; t 2 + Ila11 2 . But also
t 2 + 0: 2 +,62 + 2,6t = It - io: +,612 = If(u(t))1 2
::;
Ilu(t)112 ,
so that 0: 2 + ,62 + 2,6t ::; Ila11 2. This holds for all t E JR, and so we must have ,6 = O. Thus f(a) E JR. D
Corollary 6.21 (i) 'P(a) E JR (a E A sa , 'P E
E
Asa. Then SPA a <;;; R
(ii) Let u E U(A). Then SPA a <;;; T.
Proof (i) By Lemma 6.6, there is a commutative, unital C *-subalgebra C of E
A containing a and such that Spca = SPAa. But Spca = {'P(a) : 'P 'P(a) E JR for 'P E
D
Proposition 6.23 Let A be a unital C * -algebra, and let B be a C * -subalgebra of A with lA E B. Then SPBb = SPAb for every b E B. Proof Of course we always have Sp Bb ~ SPA b for each b E B. For the reverse inclusion, it suffices to show that, if b E B is not invertible in B, then it is not invertible in A. Thus let bE B, and suppose that b rf. G(B). By Proposition 6.2, at least one of b*b and bb* is not invertible in B, say b*b rf. G(B). But b*b is self-adjoint, and so, by Corollary 6.22(i), SpAb*b <;;; R By Corollary 4.38, SpBb*b = SPAb*b, and so b*b rf. G(A). Thus b rf. G(A). D Let H be a Hilbert space, and let T E B(H); we write SpT to denote the spectrum of T regarded as a bounded linear operator on H. Clearly, SpT is identical with the spectrum of T when T is regarded as an element of the Banach algebra B(H). By Theorem 6.23, SpT = SPAT for any C*-subalgebra A of B(H) that contains T.
272
Introduction to Banach Spaces and Algebras
Theorem 6.24 (Commutative Gel'fand-Naimark) Let A be a commutative, unital C * -algebra. Then the Gel 'fand representation of A is an isometric *-isomorphism of A onto C(1)A). Proof Using Corollary 6.21(ii), we already have the formula
-;?(cp) = cp(a*) = cp(a) = a(cp)
(a
E
A, cp
E
1>A),
and so the Gel'fand transform g: a f---+ a, A ----+ C(1)A), is a *-homomorphism. But also each a E A is normal, and so lalA = p(a) = Iiall by Lemma 6.17; this implies that 9 is isometric. The image A = {a : a E A} is a closed subalgebra of C( 1> A) that contains the constants, separates the points of 1> A, and is closed under conjugation. By the Stone-Weierstrass theorem, Corollary 2.33, A = C(1)A). Thus 9 is an isometry. 0 In the case where A is a commutative, non-unital C *-algebra, we see that 1> A i=- 0 and that 9 : A ----+ C o ( 1> A) is an isometric *-isomorphism.
Example 6.25 Let S be a non-empty set, and write (3S for the character space of the commutative C *-algebra £OO(S). Then £OO(S) is isometrically *-isomorphic to C((3S) for a certain compact Hausdorff space (3S; (3S is the Stone-Cech compactincation of S. See also Exercise 6.6. 0 6.5 The continuous functional calculus. We described the holomorphic functional calculus for an element a in a Banach algebra A in Section 4.15; this was a map 8 a from OSPAa into A such that 8 a (Z) = a. We now show that, for a normal element a of a C* -algebra A, continuous functions, rather than just analytic functions, 'operate on' a. We write C*(a) for A(a, a*), the closed C *-subalgebra of A polynomially generated by the elements a and a*. Theorem 6.26 (Continuous functional calculus) Let A be a unital C *-algebra, and let a be a normal element of A. Then there is a unique unztal *-homomorphism 8 a : C(Sp a) ----+ A such that 8 a (Z) = a. Moreover:
(i) 8 a is isometric; (ii) im 8 a
=
C*(a);
(iii) a8 a (g) = 8 a (g)a and Sp(8 a (g)) = g(Spa) for every g E C(Spa). Proof Set C = C* (a), a commutative, unital C *-subalgebra of A. By Proposition 6.23, we have Sp a = Spca. Note first that, by the Stone-Weierstrass theorem, Corollary 2.35(ii), polynomials in Z and Z* are dense in C(Spa), so that each (necessarily continuous) *-homomorphism B: C(Spa) ----+ A with B(Z) = a has imB ~ C and is uniquely defined. The mapping
273
Algebras with an involution
*-isomorphism from C onto C(Spa). The mapping 8 a that we seek is then just the inverse of this Gel'fand representation. The required properties of 8 a are immediate from Theorem 6.24. 0
Remarks: (i) The mapping 8 a is called the continuous functional calculus map for the normal element a of A. Frequently, 8 a (g) is written as g(a) whenever gEC(Spa). (ii) The continuous functional calculus extends the holomorphic functional calculus 8 a in the following sense: if f E OSPA a , then f I Spa E C(Spa) and 8 a (f) agrees with its previous definition. This fact is an easy consequence of the uniqueness property of the holomorphic functional calculus. (iii) An important application of the continuous functional calculus is obtained by taking a = T, a bounded normal operator on a Hilbert space H. Then
Ilg(T)11
=
Igl spT
(g
E
C(SpT)).
Corollary 6.27 Let A be a unital C * -algebra, and let a be normal in A. Let f E C(Spa). Then f(a) is normal and (gof)(a) = g(f(a)) (g E C(Sp(f(a)))). Proof This follows from Theorem 6.26 in the same way as Theorem 4.95 followed from Theorem 4.89. 0 Corollary 6.28 Let H be a Hilbert space, and let T E 13(H) be a normal operator. Suppose that A is an isolated point ofSpT. Then A is an eigenvalue ofT, and there is an orthogonal projection P : H ----> E(A). Proof By Corollary 4.97, there is a non-zero projection P E 13(H) such that TP = PT and SP13(p(H))(T I P(H)) = {A}. Since P is the image of a real-valued function in C(Sp T), it follows that P = P*, and hence that PT* = T* P and TP is normal. Thus (T - >.IH)P is normal in A, and p((T - >.IH)P) = O. By Lemma 6.17, TP = AP, and so Tx = AX (x E P(H)). Thus A is an eigenvalue ofT.
0
Let H be a Hilbert space, and let T E 13(H) be both compact and normal. By Theorem 4.34, each non-zero A E Sp T is an isolated point of Sp T and an eigenvalue, and the corresponding eigenspace E(A) is finite-dimensional. By Corollary 6.28, there is an orthogonal projection P).. = P; : H ----> E(A). Again, set T).. = T I E(A), so that T).. = >.IE()..) and Sp 13(E()..))T).. = {A}. It is clear that T).. is a normal operator (or matrix) on E(A), and it is standard linear algebra that there is an orthonormal basis of E(A) (use the Gram-Schmidt procedure, for example) with respect to which T).. is a diagonal matrix. The union of all these orthonormal bases, together with an orthonormal basis for ker T, is an orthonormal basis for H. Thus we may regard T as the direct sum of the zero operator on a closed subspace of H and a diagonal operator Ef){.\P).. : A E SpT}. Let us be a little more precise about the formula T = Ef){.\P).. : A E Sp T}, given above. Consider the case where Sp T is infinite. Let us enumerate Sp T as
274
Introduction to Banach Spaces and Algebras
(A n )n>l, where the An are distinct and IAn+11 :::; IAnl (n EN), and write En for E(An). For each n E N, choose an orthonormal basis {C mn _ 1 +1, .... 'c mn } of En (where mo = 1), so that (Ck k::: 1 is an orthonormal basis of a closed linear subspace of H, and set Ji,k = An for k = m n-1 + 1, ... , m n . Then we have CXl
Tx = LJi,k(X, Ck)Ck k=l
(x
E
H).
Now take n E N, and set fn(z) = Z (izi :::; IAnl) and fn(z) = 0 so that fn E C(SpT). Then T - ~~=1 AJP;,,) = 8 T (fn), and so
(Izl > IAnl),
n
T-LAJP;"JII=lfnISpT~O as n~oo, J=l where we are using the isometry condition of Theorem 6.26(i). This shows that ~{AP;" : A E SpT} converges to T in (8(H); 11·11). For kEN and x E H, set (Ck ® Ck)(X) = (x, Ck)Ck. Then we have shown that CXl
T = L Ji,kCk ® Ck k=l
in
8(H).
Now consider the formula
f(T)
=
L{f(A)P;" : A E SpT}
(*)
for f E C(SpT). The above argument shows that the sum on the right-hand side of (*) does converge to f(T) in (8(H); 11·11) in the case where, additionally, f(O) = o. Further, in the case where f is the identity function 1 on Sp T, we have
f(T)x = x = L{P;,,(x): A E SpT}
(x
E
H)
by Theorem 2.63, (a) =} (c); since each f E C(SpT) has the form f where g E C(SpT) and g(O) = 0, we see that
f(T)x
=
L {f (A) P;., (x) : A E SpT}
(x
E
=
f(O)l +g,
H)
for each f E C(SpT). Thus (*) is a correct formula when we take convergence in the strong-operator topology. We shall return to this in Section 7.3. The following result complements Corollary 6.19. Proposition 6.29 Let A and B be unital C * -algebras, let () : A ~ B be a unital *-homomorphism, and take a to be normal in A. Then: (i) f(()a) = ()(f(a)) (f E C(SpAa)); (ii) in the case where () is injective, Sp B()a = SPA a, the map () is isometric, and ()(A) is a C*-subalgebra of B.
Algebras with an involution
275
Proof (i) Certainly SPBOa <:;; SPAa. Set C = C(SpAa). Then 0 0 Sa and Sea I C are two *-homomorphisms from C into B which agree on Z, and so, by the remark on uniqueness in Theorem 6.26, 0 0 Sa = Sea I C. It follows that f(Oa) = O(f(a)).
(ii) Assume towards a contradiction that there exists A E SPA a \ Sp BOa. Take f E C(SPAa) such that f(A) = 1 and f I SPBOa = O. Then f(a) -=I- 0, but O(f(a)) = 0, a contradiction of the fact that 0 is injective. Thus SPBOa = SPAa. For each b E A, the element b*b is normal, and so it follows from the above that PB((Ob)*(Ob)) = PA(b*b). Now it follows as in Corollary 6.19 that 0 is an isometry. It is then immediate that O(A) is a C*-subalgebra of B. D An element a of a C* -algebra A is positive, written a :::: 0, if a E Asa and pa C JR+. The collection of positive elements is denoted by Apos. We write a :::: b for a, b E Asa if a - b :::: o. Theorem 6.30 Let A be a C * -algebra, and let a E Asa. Then a :::: 0 if and only if a = b2 = b*b for some b E Asa. In the case where a :::: 0, we may choose b so that also b:::: 0, in which case b zs uniquely specified. Proof Suppose that a = b2 for some b E Asa. Then Sp a = {A2 : ). E Sp b} C JR+ because Sp b c JR, and so a :::: o. Conversely, suppose that a :::: O. The non-negative square root function
Zl/2 : t
f---+
tl/ 2
is well defined and continuous on JR+, and hence Zl/2 E C(Spa). By the continuous functional calculus, the element b := Zl/2(a) E A(a) satisfies b E Asa and b2 = a. Note also that Spb = Zl/2(Spa) C JR+, so that b:::: o. Now let c E Asa be such that c2 = a. Then c commutes with a, and so also with b because b E A(a). Set C = A(b,c). Then C is a commutative C*sub algebra of A containing band c (and therefore a). By Theorem 6.23, we have Spcx = Sp AX (x E C). We now apply the commutative Gel'fand-Naimark theorem to the algebra C to see that and b are both square roots of a in C(c). If both b,c E Apos, so that b,CE C(o)+, then b = C, and then b = c. This establishes the uniqueness of b. D
c
Corollary 6.31 Let A and B be unital C * -algebras, and let 0 : A ---; B be a unital *-isomorphism. Then O(Asa) = Bsa and 0 I Asa : Asa ---; Bsa is an orderpreserving, isometric, real-linear isomorphism. In the case where A is commutative, 0 I Asa : Asa ---; Bsa is a real-algebra isomorphism. D
Let H be a Hilbert space, and take T E B(H). Recall from page 77 that T is a positive operator if and only if (Tx, x) 2: 0 (x E H). It is important that this usage does not conflict with the notion of positivity just introduced. That this is indeed the case is established in the next result.
276
Introduction to Banach Spaces and Algebras
Proposition 6.32 Let H be a Hilbert space. An operator T E B(H) is a positive operator if and only if it zs positive as an element of the C * -algebra B(H). Proof Suppose that T is positive as an element of B(H). Then T E B(H)sa and there is 8 E B(H)sa with 8 2 = T. Then, for every x E H, we have
(Tx, x)
= (8 2 x, x) = (8x,8x) = II 8xll 2
~ 0,
so that T is a positive operator. Conversely, suppose that T is a positive operator on H. Then we have
(Tx, x)
= (x, Tx) = (T*x, x) (x
E H) ,
and so, by Corollary 2.56, T = T*. Since T is normal, for each A E Sp T, there is a sequence (Xn)n~l of unit vectors in H for which (Txn, x n ) ---+ A (see Theorem 4.30). It follows that A ~ 0, and T is therefore positive as an element of the C* -algebra B(H). 0 Let T E B(H). Then 8
= T*T
is positive because
(8x, x) = (T*Tx, x) = (Tx, Tx)
~
0
(x E H) .
In fact the same result holds for an element of any C* -algebra, but that is decidedly less obvious; it will be proved in Theorem 6.38. Proposition 6.33 Let A be a unital C*-algebra, and let a Then there exists b E Asa such that eb = a. Proof We have Sp a
c
~+ •.
~
0 with a E G(A).
Set
f(t)
= logt (t > 0),
so that f E C]R(Spa) and b:= f(a) E Asa. Set g ~+., and so eb = a by Corollary 6.27.
= exp on R Then go f = Z on 0
Proposition 6.34 Each element of a unital C * -algebra is a linear combination of four unitary elements. Proof Let A be a C * -algebra. It is sufficient to show that each a E (Asa)[lJ is a linear combination of two unitary elements. For such an a, we have Sp a ~ [-1, 1]. Define f(t) = t + (t E [-1,1]),
iJ1=t2
so that f E C[-l, l], Z = (f + 7)/2, and f7 = 7f = 1. Set u a = (u + u*)/2 and uu* = u*u = lA, giving the result.
= f(a). Then 0
The following result is included mainly for the elegance of its proof; also, it will be used in Theorem 7.19.
277
Algebras with an involution
Theorem 6.35 (Fuglede's theorem) Let a be a normal element of a C* -algebra A, and let b E A satisfy ba = abo Then also ba* = a*b. Proof We may suppose that A is unital. We have anb = ban for each n EN), and it follows that, for every z E
fez) = exp(za* - za)bexp( -(za* - za)) = u(z) bU(Z)-l
(z E
q,
so that Ilf(z)11 ::; Ilbll (z E q, and hence f is bounded. But f is an A-valued entire function, and so is constant by Liouville's theorem, Theorem 3.12. In particular, fez) = f(O) = b (z E q, and so exp(za*)b = bexp(za*). Expanding the exponentials and equating the coefficients of z in the two sides of the equation gives a*b = ba*. D 6.6 Representation of general C*-algebras. Our main objective in this section is the Gel'fand-Naimark theorem, Theorem 6.47, which gives the canonical form of an arbitrary C* -algebra; en route to this, we shall show in Theorem 6.38 the key fact that a*a is always positive in a C *-algebra. We shall utilize the class of positive linear functionals on a C *-algebras A and the state space SeA). We require two further lemmas towards Theorem 6.38.
Lemma 6.36 Let A be a unital C *-algebra. (i) Let a E Asa. Then a ::::: 0 if and only if Iltl- all::; t for each (respectively, some) t::::: Iiali. (ii) Suppose that a, bE Apos. Then a + b E Apos. Proof (i) We shall use Lemma 6.17. Suppose that a ::::: 0 and t ::::: Iiali. Then Sp a <:: [0, II all], and so we have Sp(tl- a) <:: [t -ilall ,t] C JR+. Thus Iltl- all = p(tl- a)::; t. Conversely, take t E JR+ with Iltl- all::; t. Then Sp (tl- a) <:: [-t, t], so that Sp(a - tl) <:: [-t,t] and Spa <:: [0,2t]. In particular, a::::: O. (ii) Certainly a + bE Asa. Set s = Iiall and t = Ilbll, so that Ilsl- all::; sand Iitl - bll ::; t by (i). But then s + t ::::: Iia + bll and lI(s + t)1 - (a
so that a + b ::::: 0 by (i).
+ b)11
::; Iisl - all
+ Iitl -
bll ::; s
+ t, D
278
Introduction to Banach Spaces and Algebras
Lemma 6.37 Let A be a unital C*-algebra, and take a E A with -a*a Then a = o. Proof Set a = b + ie, with b, c E
Asa.
~
o.
By Proposition 4.16(ii),
Sp (-aa*) ~ Sp (-a*a) U {O} ~ lR.+ , and so -aa* ~
o.
But a*a
+ aa* = 2(b 2 + c2), a* a
=
so that
2(b 2 + c2) - aa* ~ 0
by the last lemma. Since also -a*a
~
0, we have Sp (a*a)
=
{O}, and hence
IIal1 2 = Ila*all = p(a*a) = o. Thus a
=
o.
o
Theorem 6.38 Let A be a C*-algebra. Then a*a
~
0 for every a EA.
Proof We may suppose that A is unital. Set b = a*a and B = A(b), so that B is a commutative C*-subalgebra of A. Using the commutative Gel'fand-Naimark theorem, Theorem 6.24, for B, we may write b = b+ - b-, where b+, b- ~ 0 in Band b+b- = o. Let c = ab-. Then
c*c
= b-a*ab- = b-(b+ - b-)b- = _(b-)3.
Since b- ~ 0 and Sp((b-)3) = {t 3 : t E Spb-}, we see that -c*c By Lemma 6.37, c = 0, so that (b-)3 = 0, and then, since b- E b- = O. Hence a*a = b+ ~ o. Corollary 6.39 Let A be a and A+ = Apos.
C
* -algebra. Then the subset
Apos
=
(b-)3 ~ o. we have
Asa,
0
is closed in A,
Proof We may suppose that A is unital. By Lemma 6.36(i), we have Apos
= {a
E Asa :
IIlIall . 1 - all::::; lIall},
and so it is clear that Apos is closed. By Theorem 6.30, Apos ~ A+. By Theorem 6.38 and Lemma 6.36(ii), we also have A+ ~ Apos. Hence A+ = Apos. 0 Corollary 6.40 (Polar decomposition of an invertible element) Let A be a unital = bu with
C * -algebra, and let a E G (A). Then there is a unique decomposition a b 2: 0 and u E U(A). Moreover, b = (aa*)1/2.
279
Algebras with an involution
Proof Since a E G(A), we have aa* E G(A) (by Proposition 6.2) and aa* ~ 0 (by Theorem 6.38). By Theorem 6.30, there exists b ~ 0 with b2 = aa*j clearly bE G(A). Let u = b-1a. Then u E G(A) and uu* = b-1aa*b- 1 = b- 1b2 b- 1 = 1. Thus 1 u- = u* and u E U(A). If also a = cv with c ~ 0 and v E U(A), then aa* = cvv*c = C2 j by the uniqueness statement in Theorem 4.99, c = b, and then v = u. 0 We summarize some results obtained so far. Theorem 6.41 Let A be a C * -algebra, and let a E A. Then the following are equivalent:
(a) a E A+ ; (b) a ~ 0, so that a
Apos; (c) a = b2 for some b E Asa ; (d) a = b*b for some bE A; E
(e) a E Asa and Iitl - all ::; t for each t E lR with t ~ Iiali.
o
Corollary 6.42 Each C * -algebra is *-semisimple. Proof Let A be a C*-algebra, and take a E J*(A). Since, by Corollary 6.39, A+ is closed, it follows from Theorem 6.13 that -a*a E A+. By Theorem 6.41, 0 (a)=}(b), we have -a*a ~ O. By Lemma 6.37, a = O. Thus J*(A) = {O}. Lemma 6.43 Let A be a unital C * -algebra, and take f E A *. Then f is positive if and only if Ilfll = f(I). Proof Suppose that f is positive. Then Ilfll = f(l) by Proposition 6.8(v). For the converse, suppose that f E A* with Ilfll = f(I). The result is trivial when f = 0, and so we may suppose that IIfll = f(l) = 1. By Proposition 6.20, f(a) E lR whenever a E Asa. Next, we shall show that f(a) ~ 0 whenever a ~ O. We may suppose that pea) = Iiall = 1, so that Spa <;;; 1I. Then Sp (1 - a) <;;; 1I and, by Lemma 6.17, 111 - all = p(1 - a) ::; 1. Thus 11 - f(a)1 = If(1 - a)1 ::; 1, so that f(a) ~ O. Now, for each a E A, we have a*a ~ 0 by Theorem 6.38, so that f(a*a) ~ 0 and f is a positive linear functional. 0 Let A be a unital Banach algebra. Recall from page 205 that
KA = {J E A* ; Ilfll = f(l) = I}. Corollary 6.44 Let A be a unital C * -algebra. Then SeA)
= {J E A* ; Ilfll = f(l) = I} = K A
.
Proof This is immediate from the definitions and Lemma 6.43.
o
280
Introduction to Banach Spaces and Algebras
It is simple to see that SeA) is a convex, weak-*-compact subset of A*. We now define the numerical range, WA(a) = W(a), of an element a of a unital Banach algebra A: it is the subset
WA(a) = W(a) = {I(a) : f
E
KA}.
The numerical radius of a is WA(a) = w(a) = sup{I).1 : ). E W(a)}, so that w(a) ::; Iiali. Proposition 6.45 Let A be a unital Banach algebra, and let a, b E A. Then:
(i) W(a) is a non-empty, compact, convex subset ofe; (ii) W(a1 + f3a) = a + f3W(a) (a, f3 E
q
and W(a + b)
~
W(a) + Web);
(iii) Spa ~ W(a); (iv) WB(a) = WA(a) for each Banach subalgebra B of A with 1,a E B; (v) pea) ::; w(a) ::; Ilall· Proof Clause (i) follows because W(a) is the image of KA under the continuous linear mapping f f--+ f(a) from (A*;O"(A*,A)) into e, and (ii) is clear.
(iii) Let). E Spa. For (, TJ E C, we have (A + TJ E Sp ((a + TJ1), and so I(A+TJI ::; II(a+TJ111· Thus the map f: (a+TJ1 f--+ ( ) . + TJ, lin{l,a} ----+ e, is a linear functional on lin{l,a} with f(l) = Ilfll = 1. By the Hahn~Banach theorem, f has a norm-preserving extension to an element, also called f, of KA· Thus). = f(a) E W(a). (iv) The restriction map f
f--+
fiB maps KA onto K B .
o
(v) This follows immediately. Proposition 6.46 Let A be a unital C * -algebra, and let a E A. Then:
(i) Ilall/2 ::; w(a) ::; Iiall ; (ii) a E Asa if and only if W(a) c JR; (iii) a 2 0 if and only ifW(a) c JR+; (iv) there exists f E SeA) with f(a*a)
= Ila11 2.
Proof (i) First, take b E Asa. Then pCb) = Ilbll by Lemma 6.17, and so we also have web) = Ilbll from Proposition 6.45(v). Now write a = b+ic, with b, c E Asa. Then either Ilbll 2 Ilall/2 or Ilcll 2 Ilall/2, say Ilbll 2 Ilall/2. There exists f E SeA) such that If(b)1 = Ilbll, and then
If(a)12 = f(b)2 + f(c)2 2 f(b)2 = IIbl1 2 2 (1Iall/2)2. Hence w(a) 2: Ilall/2. (ii) Suppose that a E Asa. Then, by Lemma 6.20, f(a) E JR for each f E SeA), and so W(a) C R
281
Algebras with an involution
Conversely, suppose that W(a) c R Write a = b + ic, with b, c E Asa. Then, for each f E S(A), we have f(a) = feb) + if(c) with f(a), f(b), f(c) E R Thus f(c) = 0, and so W(c) = {O} and w(c) = o. By Proposition 6.45(v), c = 0, and so a = b E Asa. (iii) Suppose that a ~ o. Then, by Theorem 6.30, a = b*b for some b E A, and so f(a) = f(b*b) ~ 0 (J E SeA)). Hence W(a) C JR+. Conversely, suppose that W(a) C JR+. By (ii), a E Asa, and then it follows that a ~ 0 because Spa ~ W(a). (iv) By Proposition 6.45(v), p(a*a) ::; w(a*a) ::; Ila*all. But p(a*a) = Ila*all by Lemma 6.17, and so w(a*a) = Ila*all. The result follows because we have
SeA) = K A .
0
We now have all the ingredients to complete the proof of the famous noncommutative Gel'fand-Naimark theorem. Theorem 6.47 (Gel'fand-Naimark theorem) Let A be a C* -algebra. Then there is a Hilbert space H and an isometric *-isomorphism of A onto a C* -subalgebra
ofl3(H). Proof We may suppose that A is unital. By Corollary 6.42, the C* -algebra A is *-semisimple. By Theorem 6.15, there is a Hilbert space H and a faithful, universal *-representation T of A on H with IITII ::; l. By Proposition 6.46(iv), for each a E A, there exists f E SeA) such that f(a*a) = Ila11 2. Thus IITfal12 ~ II(Tfa)(7rf(l))II~ = f(a*a) = IIal1 2, and this implies that IITal1 ~ IITfal1 ~ Iiali. It follows that IITal1 = lIall, and so T is an isometry; its image is a C*-subalgebra of l3(H). 0 6.7 Weak-operator convergence in l3(H). Let H be a Hilbert space. We shall later need to use the notion of weak-operator convergence in l3(H); the weak-operator topology, written wo, on l3(H) was introduced in Section 3.6. In the present case, it is the locally convex topology defined by the collection of seminorms {Px : x E H}, where
Px(T) = I(Tx,x)1
(T E l3(H)).
That this is exactly the weak-operator topology, as defined earlier, follows from the polarization identity of Proposition 2.46. The weak-operator topology is strictly weaker than the norm topology on l3(H) whenever H is infinite-dimensional. For example, let H = £2, let (en)n> 1 be the standard orthonormal basis of H, and let Pn be the operator of projection onto the nth coordinate for each n E N. Thus, for any x = (Xn)n~l in H, we have Pn(x) = xnen and (Pnx,x) = Ix n l2 ----+ 0 as n ----+ 00. Thus Pn~O. However, IIPnl1 = 1 (n EN), so certainly Pn f+ 0 in (l3(H); 11·11). The following result is a slight variation of Proposition 3.24.
Introduction to Banach Spaces and Algebras
282
Proposition 6.48 Let H be a Hzlbert space. .) (1
wo
Suppose that Tn ----7 T. Then
wo
T~ ----7 T*
.
(ii) Suppose that Tn~T and S E l3(H). Then STn~ST and TnS~TS. In particular, ifTnS = STn (n EN), then TS = ST. (iii) For any subset E of l3(H), its commutant EC zs closed under the weak0 operator convergence of sequences. Indeed, a commutant E C is always weak-operator closed. Let H be a Hilbert space, and take T E l3(H)sa. Recall that T :::: 0 (i.e. T is positive) if and only if (Tx, x) :::: 0 (x E H) (cf. Proposition 6.32). Then (l3(H)sa; ~) is a poset.
Theorem 6.49 Let H be a Hzlbert space, and let (Tn)n2:1 be an increasmg sequence in (l3(H)sa;~) with Tn ~ kI (n E N) for some k :::: o. Then (Tn)n2:1 is strong-operator convergent to T E l3(H)sa. Moreover, T = sUPn2:1 Tn. Proof First, we remark that the convergence of the sequence (Tn)n2:1 is equivalent to the convergence of (Tn - Tdn2:1' so that we may suppose that we have o ~ Tn ~ kI (n EN). For each x E H, the sequence ((Tnx, x) )n2:1 is increasing in [0, kllxl1 2 ], and so it converges in JR.+ to a limit in [0, kllxI1 2 ]. By the polarization identity, ((Tnx, y) )n2:1 converges to a limit, say 'ljJ(x, y) for each x, y E H. Clearly, 'ljJ is a sesquilinear form on H, and also IITnl1 =sup{I(Tnx,x)l: Ilxll = I} ~ k because each Tn is self-adjoint. Let x,y E H. Then I(Tnx,y)1 ~ kllxllllyll, and so I'ljJ (x, y)1 ~ kllxlillyll, i.e. 1/; is a bounded sesquilinear form. By Theorem 2.55, there is a unique T E l3(H) such that (Tx, y)
= 'ljJ(x, y) = n->oo lim (Tnx, y)
(x, y E H),
i.e. Tn ~T. Since (Tx, x) = limn->oo(Tnx, x) :::: 0 (x E H), T is positive. For every x E H, the sequence ((Tn x,X))n2:1 is increasing in JR., so that Tn ~ T, and T is an upper bound for the set {Tn: n EN}. If S is also such an upper bound, then clearly (Tnx, x) ~ (Sx, x) (x E H), and so, passing to the limit, (Tx,x) ~ (Sx, x) (x E H), i.e. T ~ S. Thus T = sUPn> 1 Tn· Let n E N. Since 0 ~ Tn ~ T, we have liT - Tnll ::; IITII .;; k, and so II(T - Tn)x11 2 ::; k II(T - Tn)1/2XI12 = k((T - Tn)x,x) for each x E H. Thus Tn ~T.
-->
0
as
n
--> 00
0
Algebras with an involution
283
Notes There is huge number of texts on C·-algebras. For an introductory account, see [125], [47, Section 3.2], and [56, Chapter 1], which continues with many interesting examples. For a major account, which nevertheless begins at a rather elementary level, see the four volumes of Kadison and Ringrose, beginning with [102, 103]. For more advanced texts, see [154] and later volumes, and also [126, 127]. As one of the many texts relating operator algebras to a branch of physics, see the volumes of Bratteli and Robinson, beginning with [34]. An important analogue of C· -algebras in the context of topological algebras is the class of algebras called 'b· -algebras' in [5, 59], 'pro-C·algebras' in [17], and 'locally C·-algebras' in [78]. We regret that we have not discussed the important topic of bounded approximate identities in C·-algebras. From such a discussion, it follows that AI I is a C·-algebra whenever I is a closed ideal in a C·-algebra A [47, Theorem 3.2.21]. Given this, it follows from Proposition 6.29 that the range of a *-homomorphism between C· -algebras is always a C· -subalgebra. See also [102, Theorem 4.1.9]. A C· -algebra A is a von Neumann algebra if, as a Banach space, A is isometrically isomorphic to a dual of another Banach space. For example, the C· -algebra £ is a von Neumann algebra; for each Hilbert space H, the C·-algebra B(H) is a von Neumann algebra. It is a theorem of Sakai that each von Neumann algebra can be represented as a unital C· -subalgebra A of B( H) for some Hilbert space H such that A CC = A (see [103, 10.5.87] and [154, III.3.5]). For a substantial theory of von Neumann algebras, see [103] and [127, Section 9.3]. Considerations of the numerical range of an element of a unital Banach algebra lead to a striking geometric characterization of C· -algebras in the Vidav-Palmer theorem; see [31, Section 38] and [127, Theorem 9.5.9]. For more on the strong-operator and weak-operator topologies on B(H) for a Hilbert space H, see [102, Section 5.1]; in fact, there are several other interesting topologies on B(H). A unital C· -subalgebra of B(H) is a von Neumann algebra if and only if it is closed in the weak-operator topology on B(H). Let A be a C· -algebra. Then every derivation from A into a Banach A-bimodule is automatically continuous [47, Corollary 5.3.7]; this is a theorem of Ringrose [141]. In Section 5.11, we discussed the 'dense range problem', which asks if each homomorphism 8 : A ---> B, where A and B are Banach algebras with 8(A) = Band B semisimple, is necessarily continuous. This problem is open even when A is a C·algebra, and even when both A and B are C· -algebras; various partial solutions are given in [3] and [70]. It is also not known if every epimorphism from a C· -algebra onto a Banach algebra is automatically continuous. There is a vast theory on when a C· -algebra is amenable, and on the cohomological properties of C· -algebras and of von Neumann algebras; see [149], for example. (Xl
Exercise 6.6 A topological space X is completely regular if, for each closed subset F of X and each x E X \ F, there exists f E Cb(X) such that f I F = 0 and f(x) = 1. Let X be a completely regular space, and let (3X denote the character space of the commutative C·-algebra Cb(X). Show that X is homeomorphic to a dense subset of (3X, and that every function in Cb(X) extends to a continuous function on {3X. The space {3X is the Stone-Cech compactincation of X. Show that {3N is a compact, extremely disconnected space (in the sense that every open subset of (3N has an open closure) and that I{3NI = 2'. For more on {3X and especially (3N, see [83, 95]. Exercise 6.7 Let T be a compact, normal operator on a Hilbert space H. We have shown that Tx = L),xpA(x) : ,x E SpT}
Introduction to Banach Spaces and Algebras
284
in H for each x E H. However, the series is not necessarily absolutely convergent. Indeed, take H = £2 and Tx = (anx n ), where an -> 0 in IR+-. Then Tis compact and self-adjoint, the eigenvalues of l' are just the numbers an, and Pn is the projection onto the nth coordinate. So the above sum becomes 1'x = L:~= I anxnen for x E £ 2. Show that there exist (x n ) E £2 and (an) E Co such that absolute convergence of the series fails. Exercise 6.8 Let T be a compact operator on an infinite-dimensional Hilbert space H. Then 1'*1' is compact and self-adjoint, and so there is an orthonormal sequence (ukh~l in K := ker(T*1')1- such that, for each kEN, we have T*Tuk = fi,kUk for some fi,k E
a
=
1 -2' 7rl
1 11
dz exp(za)-2; Z
(v) II all /e::; w(a) ::; lIall. Exercise 6.13 Consider the left and right shift operators Land R on the Hilbert space H = £2, as in Exercise 2.11. Note that L* = R. Show that L n -> 0 in (8(H);so), but that R n ft 0 in (8(H); so). Deduce that the map T f--+ T* on 8(H) is not strongoperator continuous, and hence that the strong-operator and weak-operator topologies do not coincide on 8(H).
7
The Borel functional calculus
Our aim in this chapter is to generalize the continuous functional calculus of the previous chapter to give a 'Borel functional calculus' and a 'spectral theorem'. To do this we require a preliminary discussion of an integral; this discussion is given in the first section, and then our main theory will be developed in the second section of this chapter. Our development of the theory eschews any knowledge of measure theory.
The Daniell integral In the theory of the Daniell integral, we show how to 'integrate' rather a lot of functions without previously introducing any measure theory, or the notion of a Lebesgue integral. 7.1 Baire functions and monotone classes. Let X be a non-empty set. As in Example 1.4, the space (JR x ; :::;) is a poset with the ordering: f :::; 9 if and only if f(x) :::; g(x) (x E X). For E C;;;; JRx, we write E+ for the set of functions fEE with f(X) C;;;; JR+. We shall consider a sequence (ft) == (ft)t>1 in JRx and an individual function f in JR x . We shall write ft l' f (or ft l' f as i ----+ (0) to mean that: (i) (ft) is an increasing sequence in (JR x ; :::;) ; (ii) ft(x) ----+ f(x) as i ----+ 00 for every x E X. The notation ft 1 f is defined analogously. We say that, in either case, f is the pointwise monotone limit of the sequence (ft). Suppose that ft l' f and gt l' 9 and that >., j1 E JR+. Then it is easy to see that >'ft + J19t l' >.f + j1g. Now let M be a subset of JR x . Then M is a monotone class on X if it is closed under pointwise-monotone limits: i.e. if (ft) is contained in M and if either ft l' f or ft 1 f, then also f E M. It is clear that JRx itself is a monotone class, and also that the intersection of any collection of monotone classes on X is a monotone class. Hence, for each subset E of JRx, there is a unique smallest monotone class M on X with M :2 E. This smallest M is called the monotone class generated by E. Let K be a non-empty, locally compact Hausdorff space, and set L = CO,IR (K), the space of continuous, real-valued functions that vanish at infinity, so that L C;;;; JRK. Then the set of (real-valued) Baire functions on K is defined to be the monotone class generated by L. A real-valued Baire function on K is simply a member of this set; a complex-valued function on K is a Baire function if and only if its real and imaginary parts are both real-valued Baire functions.
286
Introduction to Banach Spaces and Algebras
In all interesting cases (e.g. K = II), it is elementary to see that Cffi. (K) is not itself a monotone class, so that the set of real-valued Baire functions is strictly larger than Cffi.(K). For example, the characteristic function of any non-empty, open subset of II is the pointwise limit of an increasing sequence of continuous functions, and so a real-valued Baire function. Also, be warned that there is generally no useful description of an arbitrary Baire function on K; this fact has a profound effect on methods of proof. Because our interest will be entirely in bounded functions, we have a slightly modified notion of 'monotone class'. Let X be a non-empty set. A subset M of IR x is a bounded-monotone class if it is closed under bounded monotone limits, in the sense that, if (f,) is an increasing sequence in M with f, ::; c (i E N) for some c > and if f, f, then f E M, and similarly for decreasing sequences. For each subset E of IR x , there is a unique smallest bounded-monotone class containing E; it is the bounded-monotone class generated by E. For the remainder of this section, we take K to be a non-empty, locally compact Hausdorff space, and set L = Co,ffi.(K). We write Bffi.(K) and B(K)for the sets of bounded, real-valued Baire functions on K and of bounded Baire functions on K, respectively. Clearly, Bffi.(K) and B(K) are real- and complexvector spaces, respectively.
°
r
Lemma 7.1 (i) The set Bffi.(K) is the bounded-monotone class generated by L.
(ii) The set Bffi.(K)+ is the bounded-monotone class generated by L+. Proof (i) From the definition of the Baire functions, it is clear that Bffi.(K) is a bounded-monotone class on K that contains L. Now suppose that M is any such bounded-monotone class, and define E to be the set of all those real-valued functions f on K such that, for every c > 0, we have (f /\c) V (-c) E M. It is evident that E is a monotone class that includes L, and therefore it includes the set of all real-valued Baire functions on K. But then, if f E Bffi.(K), we have fEE. However, for sufficiently large c > 0, necessarily f = (f /\ c) V (-c) E M.
(ii) This is similar, and is left as a simple exercise.
o
Theorem 7.2 Let K be a non-empty, locally compact Hausdorff space. Then B(K) is a C * -subalgebra of (ROO(K); I·IK)' Also, Bffi.(K) is a sublattice of Rli{'(K). Proof Let f E Bffi. (K), and then take M (f) to be the set of all those functions 9 E Bffi.(K) such that each of f + g, f V g, and f /\ 9 is in Bffi.(K). It is clear that M(f) is a bounded-monotone class and that M(f) ::2 L for each f E L. Thus, by Lemma 7.1(i), M(f) is the whole of Bffi.(K). However, 9 E M(f) if and only if f E M(g), so that M(g) ::2 L for every 9 E Bffi.(K). Thus, for every f,g E Bffi.(K), we see that f + g, f /\g, and fV g are all in Bffi.(K). Also, )..f E Bffi.(K) for each f E Bffi.(K) and)" E IR. In particular, Bffi.(K) is a sublattice of £li{'(K).
287
The Borel functional calculus
A similar argument (first considering non-negative functions, and then using Lemma 7.1(ii)) shows that BIlt(K) is closed under the pointwise product. Now suppose that (fn)n?1 is a sequence in BIlt(K) with fn --+ f uniformly. By replacing each fn by fn -If - fnlK . 1, we may suppose that fn ~ f (n EN). By replacing each fn by II V··· V fn, we may suppose that fn+! :::: fn (n EN). Now fn i f, and so f E BIlt(K). Thus BIlt(K) is closed in (£IR'(K); I·I K )· Since B(K) = BIlt(K) EEl iBIlt(K), the extension to complex-valued functions, by combining real and imaginary parts, is now immediate. D
Corollary 7.3 Let K be a non-empty, locally compact Hausdorff space. Then
SPB(K)f
=
f(K)
(f
E
B(K)).
Proof Let f E B(K). For each x E K, the map ex : f ~ f(x) is a character on B(K), and so f(K) <;;; SPB(K)f· Since SPB(K)f is compact, f(K) <;;; SPB(K)f. Now take A E
(i) X
E
I:;
(ii) if E E I:, then X \ E E I:; (iii) if (En)n?1 is any sequence of sets in I:, then also
Un?1 En
I:. Let I: be a O'-field on X. Then, from (i) and (ii), 0 E I:, so that I: is also closed under finite unions. Also, by (ii) and (iii), I: is closed under countable intersections. Clearly, P(X) is itself a O'-field on X, and the intersection of any collection of O'-fields is a O'-field. Thus, given a family E in P(X), there is a smallest O'-field on X that includes E; this is the O'-field generated by E. Let I: be a O'-field on a set X. A complex measure on I: is a function fJ : I: --+
of sets in I: with U~1 E J = E. Now let (X; T) be a topological space. Then the Borel O'-field on X is the O'-field generated by T; it is denoted by B(X). A Borel subset of X is simply a member of B(X). A measure defined on B(X) is a Borel measure. For technical reasons, it is sometimes necessary to use a (potentially) smaller O'-field on X than B(X): this is the Baire O'-field on X. It is the O'-field generated by the collection of all the compact, Go-subsets of X, and it is denoted by Ba(X). It.is important to note that, when X is a compact metric space (in particular, a compact subset of
288
Introduction to Banach Spaces and Algebras
Consider a function I : X ---., C. Then I is a Borel function (respectively, a Baire function) if and only if, for every open subset U of C, the set 1-1 (U) is a Borel set (respectively, a Baire set) in X. Clearly, every function in C(X) is a Baire function. Suppose that I is a Borel function. Then it is easy to see that E B(X) for each B in B(C), and so g 0 I is a Borel function whenever g : C ---., C is a Borel function. It is not difficult to show that this notion of a Baire function on a locally compact Hausdorff space K is identical with that previously defined in terms of monotone classes on K. Note that, for us, a Baire subset of X is just a subset E such that XE is a Baire function; again, this is equivalent to the definition in terms of O"-fields. The following easy remark gives some examples of Baire functions.
1- 1 (B)
Proposition 7.4 Let K be a non-empty, compact metric space, let a E K, and let I E eOO(K) be continuous on K \ {a}. Then I E B(K). Proof Since eOO(K) = lineOO(K)+, it suffices to suppose that I E eOO(K)+. Set U = K \ {a}. Then there is an increasing sequence (h t ) in C~(K)+ with h t i Xu, and with supph t <:;;; U (i EN). So each htl is continuous on K, and h,f i xul, which is therefore a Baire function on K. But also X{a} is a Baire function, so that I = I(a)x{a} + Xu I is a Baire function. 0
7.2 The Daniell extension process. Throughout this section, K will be a non-empty, compact Hausdorff space, and we again set L = C~(K). (We require K to be compact, not just locally compact, to ensure that 1 E K; more general situations are indicated in the notes.) An integral on L is defined to be a real-linear functional 1 : L ---., IR which is positive, in the sense that 1(f) 2': 0 whenever I E L +. For example, set K = il, and let 1(f) =
10
1
I(x) dx
(f E L),
the Riemann integral. Then 1 is an integral on L in our new sense.
Lemma 7.5 Let 1 be an integral on L. Then:
(i) 11(f)1 ::; 1(1) IIIK (f E L); (ii) il (ft) in L and
IE
L, and il either It
1I
or It
i I, then 1(f,) ---., 1(f).
Proof (i) This follows from the positivity of 1 since -IIIK ::; I ::; I/IK (f E L). (ii) By Dini's theorem, Corollary 5.38, I, ---., I uniformly on K, so that 1(f,) ---., 1(f) by (i). 0 The Daniell process will give a way to extend 1 to an integral on B~ (K). In the case of the above example, with K = IT, this would lead to the Lebesgue integral restricted to the bounded, Borel functions, rather than on the slightly
289
The Borel functional calculus
more general space of all Lebesgue integrable functions. But it should be noted that, in general, the definition of an 'integrable function' would depend on the particular integral 1 being extended. The definition of B(K), in contrast, depends only on L, and provides a convenient common domain to which all integrals on L can be extended. This will be important for us in Theorem 7.16. We shall suppose, for the rest of this section, that we have a specified integral 1 on L = C]R(K).
The first extension step. Let U be the subset of JRK consisting of those f E £;0 for which there is some increasing sequence (ft) in L with ft i f. It is immediately clear that U :2 L and that U is closed under addition, multiplication by non-negative scalars, and the lattice operations. It is also clear that f 9 E U+ whenever f, 9 E U+. Let fEU, and take (ft) in L with ft i f. Since 1 is positive, the sequence (J(ft) )t>l is increasing and bounded above in JR, and so limH(X) 1(ft) exists in R We would like to define 1 on U by setting 1(f) = lim 1(ft). t-+(X)
(*)
We must first check that 1(f) is well defined.
Lemma 7.6 Let (ft) and (gJ) be increasing sequences in L, each uniformly bounded above, with limJ-+(X) gJ ~ limt-+(X) ft. Then lim 1(gJ) ~ lim 1(ft). J~CX>
1,----+00
Proof Suppose first that h ELand that limH(X) ft :2: h. Then ft :2: ft 1\ hand ft 1\ hi h, so that limH(X) 1(ft) :2: limH(X) 1(ft 1\ h) = 1(h) by Lemma 7.5(ii) and the order-preserving property of 1. Taking h = 9J' and then passing to the limit as j ----+ 00, gives the result. D Corollary 7.7 (i) The mapping 1 is well defined on U by (*). (ii) 1 I L
= 1.
(iii) 1(f + g)
= 1(f) + 1(g) (f,g E U) and 1(>.f) = >.1(f) (>. :2: 0, fEU).
(iv) 1(f) ~ 1(g) whenever f,g E U with f ~ g.
Proof (i) and (iv) follow immediately from Lemma 7.6, and (iii) is a simple exercise. For (ii), we may take ft = f for all i E N. D E £;o(K), and suppose that there exists (ft) in U with ft i f· Then fEU and 1(ft) ----+ 1(f).
Lemma 7.8 Let f
Proof For each i E N, choose (g;)J?l in L with!!? i ft as j ----+ 00, and then define h j = gi v··· V g~ (j EN). Then (hJ)J?l is an increasing sequence in L with ~ h J ~ fJ for i = 1, ... ,j.
g;
290
Introduction to Banach Spaces and Algebras
First, let j ----> 00 to deduce that f, :::; limJ--->CXl hJ :::; f, and then let i ----> 00 to see that f :::; limJ--->CXl h J :::; f· Thus hJ i f, so that fEU. Analogously, from the inequality I(gn :::; I(h J ) :::; J(fJ) (i = 1, ... ,j), we see that lim HCXl J(f,) :::; J(f) :::; limJ--->CXl J(fJ) , so that limJ--->CXl J(f,) = J(f). 0 Now let -U = {f : - fEU}. For f E -U, we define l(f) = -J( - 1). (In fact, -U is then precisely the set of f E CiR such that f, 1 f for some decreasing sequence in L, and the definition of 1 is equivalent to setting l(f) = lim,--->CXl I(f,) for any such sequence.) Suppose that f E (-U) n U. Then l(f) = J(f) because f + (-1) = 0 and J is additive on U. (This applies in particular to all f E L.) The set -U, with the mapping L has properties exactly analogous to the properties of U and J given in Corollary 7.7. Thus we can conclude that L <;;; -U and that II L = I. Also, -U is closed under bounded, decreasing limits, addition, multiplication by non-negative scalars, and the lattice operations. The mapping 1 is additive and positive-homogeneous on -U. Further, if f, 9 E -U with f :::; g, then l(f) :::; [(g). We also have the following simple lemma. Lemma 7.9 Take 9 E -U and h E U with 9 :::; h. Then leg) :::; J(h) and h - 9 E U.
Proof Using Corollary 7.7, we have h - 9 J(h) -leg)
= h + (-g)
E U and
= J(h) + J( -g) = J(h - g)
~
o. o
This implies the result.
The second extension step. It remains the case that we are taking K to be a non-empty, compact Hausdorff space and that L = CJR(K). Let f E CiR(K). Then f is I-summable (or just sllmmable, when I is clear) provided that: sup{l(g) : 9 E -U, 9 :::; f}
= inf{I(h)
: h E U, h ~ f}.
In considering this definition, it should be noted that, by Lemma 7.9, [(g) :::; J(h) when 9 :::; f :::; h in U. We denote the set of I-summable functions in IRK by L~(K). (We use the notation 'L~', rather than 'L~', because we are considering just bounded functions on K.) Let f E L~(K). Then we define J(f) to be the common value of the above sup and inf, i.e. we define J(f)
= sup{l(g) : 9
E
-U, 9 :::; f}
= inf{J(h) : hE U, h ~ f}.
We now aim to show that BJR(K) <:::; L'fi'(K); it will then be the case that is the extension of I that we are seeking. Before proceeding, it is useful to note two alternative forms of the definition of summable functions.
J I BJR
The Borel functional calculus
291
Lemma 7.10 Let f E fYr(K). Then:
(i) f is summable if and only if, for every c > 0, there are functions 9 E -U and h E U with 9 :s: f :s: hand I(h) -[(g) < C; (ii) f is summable if and only if there is an increasing sequence (g')'"21 in -U and a decreasing sequence (h,),>l in U with g, :s: f :s: h, (i E N) and I (h,) -[(g,) ----* 0 as 1, ----* 00; moreover~ in this case, for any such pair of sequences (g')'"21 and (h')'"21, we have IU) = lim [(g,) = lim I(h,) . 1,----+00
,/----+00
Proof This is an easy exercise.
D
Note that, in clause (ii) of Lemma 7.10 it is NOT claimed that the sequences (h')'"21 and (g')'"21 converge pointwise to f·
Corollary 7.11 With the above notation: (i) UU(-U) <:;; Lif(K); (ii) II U = I; (iii) II (-U) = Ii (iv) L = I.
Yi
Proof Let fEU. Then there exists (g,) in L <:;; (-U) with g, i f and with [(g,) = I(g,) ----* IU), and we may take h, = f (i EN). Now use Lemma 7.1O(ii) to establish (i) and (ii); (iii) is similar, and then (iv) follows. D The next lemma is the easy part of what must be proved.
Lemma 7.12 The set Lif (K) is a real vector subspace and sublattice ofF. K , and a closed subalgebm of (fYr(K); I·I K ). Further, I is a positive linear functional on Lif(K) such that IIU)I:S: IfI K l(l) U E Lif(K)). Proof This follows easily from either of the alternative formulations of the definition of summability. That IIU)I :s: IflK 1(1) U E Lif(K)) follows as in the proof of Lemma 7.5(i). D Let E be a bounded-monotone class with L <:;; E <:;; fYr(K). Then a linear functional J on E has the monotone-convergence property if JU,) ----* JU) whenever U,) in E with f, i f·
Theorem 7.13 (Monotone-convergence theorem) Let K be a non-empty, compact Hausdorff space. Then the Banach space Lif(K) (with the norm I·I K ) is a bounded-monotone class, and the functional I on Lif(K) has the monotoneconvergence property. Proof Suppose that I E fYr(K) is such that f, i f for a sequence U,) in Lif(K). Note first that, since I is bounded, there is c > 0 with I, :s: I :s: c (i EN), and so (IU,)).>l is increasing with JU.) :::; cI(l) (i EN), and hence limt---+oo JU.) exists. By ~onsidering the sequence U, - f)'"21, we may suppose that 1=0.
292
Introduction to Banach Spaces and Algebras
Take c
> O. For each i
o :s:
f, - f,-l
E
N, choose h,
:s:
h,
and
E
-
U with -
I(h,) < IU, - f,-d
c
+ 2'
.
Let n E N. Then fn :s: 2::~=1 h, := H n , say, and Hn E U+. Further, Hn /\ c E U+ and Hn /\ c;::::: fn, and (Hn /\ C)n21 is an increasing sequence, bounded above by c. Set H = limn--->oo Hn /\ c, so that H ;::::: f and Hn /\ c i H. By Lemma 7.8, we have H E U+, and n
I(H) = lim I(Hn /\ c) n----+CX)
:s: lim sup LI(ht):S: lim IUn) + c < 00. n----+oo n----+oo ,=1
Now choose N so that I(H) < IUm) + 2c (m;::::: N). Since fN E L'fR(K), there exists 9 E -U with g:S: fN and with [(g) > IUN) - c. We now see that 9 :s: fN :s: fm :s: f :s: H (m;::::: N), and I(H) < [(g) + 3c. Thus f E L'fR(K), so that L'fR(K) is a bounded-monotone class, and, further, IU) = limn--->oo IUn) , so that the space L'fR(K) has the monotone-convergence property. 0 Corollary 7.14 We have B'R,(K) <:;; L'fR(K). Proof By the theorem, L'fR(K) is a bounded-monotone class with L <:;; L'fR(K). But B'R,(K) is the smallest bounded-monotone class containing L, and so we have B'R,(K) <:;; L'fR(K). 0 For simplicity, we now write I for II B'R,(K). Thus I is a real-linear functional on B'R,(K). Theorem 7.15 The functional I is the unique extension of I to a positive linear
functional on B'R,(K) satisfying the monotone-convergence property. ~roof
Certainly I is a positive linear functional on B'R,(K) that extends I, and I has the monotone-convergence property. Now let J be any positive linear functional on B'R,(K) with J I L = I such that J has the monotone-convergence property. Then J must agree with I on U and with [ on -U, and hence J agrees with Ion U U (-U). But then, since B'R,(K) <:;; L'fR(K), the definition of I-summability makes it clear that J = Ion the whole of B'R,(K). 0 We conclude by extending the above to the complex-valued case. An integral on C(K) is now defined to be a complex-linear functional I : C(K) ~
IU so that
+ ig)
f: B(K)
+ iI(g)
U
+ ig E
=
IU)
B'R,(K) EB iBlR(K)
~
=
B(K)) ,
The Borel functional calculus
293
In a similar way, we can define LOO(K) to be the complexification of L';:(K); specifically, LOO(K) = {f + ig: f,g E LOO(K)}. It is easily seen that LOO(K) is a C*-subalgebra of (1;OO(K); I·I K ). Notes An approach to the Daniell integral via measure-theoretic considerations is given in [110], and there is a more extensive account in [128, Section 6.1]; we have modified this theory to concentrate on bounded functions. One can run the extension process described above starting with any non-empty set S (rather than K) and any vector space L ~ ]Rs such that L is closed under V and 1\ (rather than just L = CII(K)) and any positive linear functional J on L such that JU,) ---> 0 whenever j, 1 O. For example, we could take K to be a non-empty, locally compact Hausdorff space, and L to be C o,II(K), the space of continuous, realvalued functions on K that vanish at infinity. The results in Section 7.1 still apply, but in Section 7.2 a problem arises because it is no longer true that 1 E L (in the case where K is not compact). In this case, it is natural to modify our approach by dealing with monotone classes, rather than bounded-monotone classes; in this modified approach, one finishes with the Lebesgue measurable functions. See [110, Section 9.4] and [117, Section 12]. The notation L'{'(K) is not standard. In the case where K = IT and the initial integral on CII(IT) is the Riemann integral, functions in L'{'(IT) are the 'bounded, Lebesguemeasurable functions' of measure theory; this gives some justification for the notation. Let K be a non-empty, compact Hausdorff space. Then we can define a 'measurable set' as a subset A of K such that XA E L'{'(K); we obtain a O'-field of measurable subsets of K. A subset A ofIT is 'Lebesgue measurable' if and only if XA E L'{'(IT). Let K be a non-empty, compact Hausdorff space. The Baire functions of order a are the functions in C(K). Given a definition of the Baire class of order {3 for each (3 < n, the class of order n is the space of bounded functions on K which are pointwise limits of sequences of functions in the union of the earlier classes; the construction terminates at n = WI. The Baire functions on K are the members of this final class. Each Baire class is itself a Banach algebra which is a closed subalgebra of (B(K); I·I K ). Some properties of the C* -algebra B( K) and of its character space are given in [53]. Exercise 7.1 Let K be a non-empty, compact, metrizable space. Show that the space U is precisely the set of bounded, lower semi-continuous, real-valued functions on K. Exercise 7.2 Let X be a non-empty, discrete space (so that X is a locally compact, metric space). Show that B(X) = P(X) and that Ba(X) consists of subsets Y of X such that either Y or X \ Y is countable. Thus, in the case where X is uncountable, B(X) =I Ba(X). Exercise 7.3 Let K be an uncountable, compact, metrizable space (such as K = IT). Prove that IKI = c, that IBII(K) I = c, and that 1l'nr'(K) I = 2', so that BII(K) <;; l'nr'(K). Exercise 7.4 Let C be the Cantor set, introduced in Exercise 1.7. Show that the characteristic function of every subset of C belongs to L'{' (IT). Deduce that IL'{' (IT) I = 2', and hence that BII(IT) <;; LR'(IT). Exercise 7.5 The following exercise requires a subset E of IT which is not Lebesgue measurable; such a set is 'constructed' in [143, Example 2.22], for example. (The construction of such a set E requires the axiom of choice.) Then XE E l'nr'(][), but XE f/- L'{'(][), and so LR'(][) <;; l'nr'(][).
294
Introduction to Banach Spaces and Algebras
Exercise 7.6 This exercise requires some knowledge of the Stone-Cech compactification j'3N of the discrete space N; this compact space was introduced in Exercise 6.6. The Baire subsets of j'3N are determined by the space C(j'3N), and IC(j'3N)I = c, so that IBa(j'3N)I = c. However, each singleton in j'3N is a Borel set, and so IB(j'3N)I ;::: Ij'3N I = 2'. This shows that Ba(j'3N) <;;; B(j'3N). [In fact, IB(j'3N)I = 2'.J
The Borel functional calculus and the spectral theorem 7.3 Introduction to the spectral theorem. We shall now establish a Borel functional calculus for normal operators on a Hilbert space, en route to a spectral theorem. We first think what such a calculus might consist of. Let T be a normal operator on the finite-dimensional Hilbert space en. Then the spectral theorem for T is another name for the diagonalization theorem of linear algebra: en has an orthonormal basis consisting of eigenvectors of T. The next stage of generality is for T to be a compact, normal operator on an infinite-dimensional Hilbert space H. In this case, the spectrum Sp T of T consists of {a}, together with a finite or countable sequence of non-zero eigenvalues, each having a finite-dimensional eigenspace. In the case where the sequence of eigenvalues (An) is infinite, An ---+ a as n ---+ 00. As we remarked on page 273, H has an orthonormal basis that consists of eigenvectors of T. These first two cases may be stated together as follows. Let T be a compact, normal operator on a Hilbert space H. For each eigenvalue A ofT, with eigenspace E(A) and corresponding orthogonal projection P).., the 'spectral theorem' for T is equivalent to the statement:
T = with convergence in (l3(H); T gives:
f ~ f(T) =
L {>. P).. : A E Sp T} ,
(*)
II . II). Further, the continuous functional calculus for
L {J(A) P).. : A E SpT},
C(SpT)
---+
l3(H) ,
with convergence in the strong-operator, and hence weak-operator, topology on l3(H); see Section 4.6. Now suppose that T is an arbitrary normal operator in l3(H). Then T need not have any eigenvalues-and even when it does, Sp T may have many points that are not eigenvalues. There is still a good analogue of (*), but it is less elementary, in that the infinite (or finite) sum in (*) is replaced by a kind of integral with respect to a 'projection-valued measure', so that, symbolically,
T =
r
A dE).. .
JSpT
In terms of this representation, the continuous functional calculus of Theorem 6.26 appears as:
The Borel functional calculus
f(T) =
295
r
f().) dE)..
(fEC(SpT)).
(**)
iSpT
However, a very important point about the spectral theorem is that we may make good sense of the above formula (**) for functions f that are more general than continuous functions-namely for bounded Borel (or Baire) functions. In this case, the map f f--+ f(T) becomes the Borel functional calculus (precise definitions will be given in the next section). We shall reverse the more usual order by developing the Borel functional calculus without assuming knowledge of integration theory (we shall prove what we need). Indeed, we shall extend the continuous functional calculus map to a certain *-homomorphism from the commutative C *-algebra B(Sp T) into B(H) (where B(SpT) is the algebra of bounded Borel functions on SpT). This extended homomorphism is precisely the Borel functional calculus. We shall give a few applications of this calculus and, finally, explain briefly and informally how the integral representation may be obtained from the Borel functional calculus. 7.4 The Borel functional calculus. Let H be a Hilbert space, and let T be a normal operator in B(H). Then, from Theorem 6.26, the continuous functional calculus BT is an isometric *-isomorphism from C(Sp T) onto the commutative C*-subalgebra C*(T) of B(H) generated by T. As has already been indicated, the Borel functional calculus is a certain extension of BT to a *-homomorphism
f3T : B(Sp T)
--->
B(H) .
Note that, in view of Corollary 6.19, f3T will be automatically continuous, with IIf3T(f) II :::; Ifl spT (f E B(SpT)). The homomorphism f3T will be uniquely determined, as an extension of BT , subject to one very natural extra condition involving monotone sequences.
Remarks: (i) Unlike the continuous functional calculus, the Borel functional calculus is definitely about homomorphisms into B(H), and not just into some abstract C* -algebra. This arises because, although the continuous functional calculus maps into (indeed, onto) A = C*(T), the extended map f3T (in all interesting cases) takes values outside A. In fact, as will be seen, f3T maps into (but not, generally, onto) the algebra A ee, the bi-commutant of A (which we know to be a commutative C*-subalgebra of B(H)). (ii) Also, unlike the continuous functional calculus, the Borel functional calculus is not generally injective; this gives significance to clause (ii), below. Theorem 7.16 (Extension of homomorphisms) Let K be a non-empty, compact Hausdorff space, let H be a Hilbert space, and let B : C(K) ---> A be an isometric *-isomorphism onto a C*-subalgebra A of B(H). Then B can be extended to a *-homomorphism f3 : B(K) ---> B(H), and the extension f3 is unique subject to the following monotone-convergence property: if (f.) in BIR(K) , if
f
E
BIR(K) , and if f. i f,
then
f3(f.)~f3(f).
296
Introduction to Banach Spaces and Algebras
Moreover: (i) (3(B(K)) <;;; Ace; (ii) (3(XU) -=I- 0 for each non-empty, open subset U of K ; (iii) SPB(H)(3(f) <;;; f(K) (f E B(K)) . Proof Again, we set L = CJR(K). By Corollary 6.31, 0 ILL order-preserving, isometric, real-linear algebra isomorphism. For each x E H, define Ix : L ----f lR by
Ix(f) = (O(f)x,x)
----f
Asa is an
(f E L).
Then Ix is an integral on L,~nd so, by Theorem 7.15, it has a unique extension to a positive linear functional Ix on BJR(K) that satisfies the monotone-convergence property. Next, for each f E BJR(K), we define a mapping IJ f : H x H ----f C by the formula:
IJf(x, y) = For every
f
E
~ (Ix+y(f) -Ix_y(f) + iIx+iy(f) -
iIx-iy(f))
(x, y
E
H).
L, the polarization identity shows that
IJf(x,y) = (O(f)x,y)
(x,y
E
H).
In particular, IJ f is a bounded, self-adjoint, sesquilinear form on H such that IlJf(x,y)l:::; Ilfllllxllllyll (using the fact that 110(f)11 = IfI K )· Let M be the set of all those f E BJR(K) such that IJ f is a self-adjoint, sesquilinear form on H. We have just seen that M ::2 L, so that, to prove that M = BJR(K), it suffices to show that M is a bounded-monotone class. But this follows easily from the fact that if, say, f. l' f with f., f E M, then Iz(f.) ----f IAf) for every Z E H, and thus IJf,(x,y) ----f IJf(x,y) for all X,y E H. We shall now show that IJ f is a bounded form for every f E BJR (K). Since we know that IJ f is sesquilinear, it suffices to show that it is a bounded function on the set {(x, y) : x, y E H[lJ}. Indeed, take x, y E H[lJ. By Lemma 7.12, we have
IIIzl1 = I z (1) = IIzI12
(z E H), and so the formula for IJf shows that
IlJf(x, y)1
:::;
IflK
(11x11 2+ IIYI12) :::; 21flK
,
so that the boundedness of IJ f is proved. (We shall later see that the '2' is unnecessary, but that seems to be unclear at this point.) By Theorem 2.55, it follows that, for each f E BJR(K), there is a unique operator (3(f) E 8(H)sa such that IJf(x,y) = ((3(f)x,y) (x,y E H)---or, more simply, (3(f) is unique subject to the condition that
((3(f)x, x) = Ix (f) It is then clear that the mapping (3 : BJR(K)
(x
E
H).
8(H)sa is a real-linear mapping and that, if either f. i f or f. ! f in BJR(K), then (3(f.)~(3(f). ----f
297
The Borel functional calculus
We now extend (3 to be a complex-linear mapping from B(K) into 8(H) by combining the real and imaginary parts of functions in the obvious way. Trivially, this extended mapping satisfies (3(l) = (3(J)* (J E B(K)). To show that (3 is multiplicative, it suffices to consider just non-negative functions f, 9 E B(K)+. So, for 9 E B(K)+, set
M(g) = {J
E
B(K)+ : (3(Jg) = (3(J)(3(g)}.
Then M(g) is a monotone class (using Proposition 6.48(iii)). Also, if 9 E L+, then M(g) :2 L+, so that M(g) = B(K)+. But clearly f E M(g) if and only if 9 E M(J), so that, for every 9 E B(K)+, we have M(g):2 L+, and so, again, M(g) = B(K)+. It has thus been shown that (3(Jg) = (3(1)(3(g) (1, 9 E B(K)+), and hence (3 is multiplicative. Combining real and imaginary parts, it then follows that (3 : B(K) ---+ 8(H) is a *-homomorphism. So (3 is automatically norm-decreasing (whence it may be noted that 100f(x,y)1 :::; IflK Ilx1111Y11 (x,y E H), without the extra '2'). (i) To show that we have (3(BJR(K)) <:;;; A CC, we just note that, by the monotone-convergence property of (3 and the fact that A cc is closed under weakoperator convergence of sequences, it is clear that {J E BJR(K) : (3(J) E A CC} is a bounded-monotone class containing L, and thus equal to the whole of BJR(K). The result follows immediately. (ii) Let U be a non-empty, open subset of K. Then, by Urysohn's lemma, there is a non-zero function h E L with 0 :::; h :::; 1 on K, while h(x) = 0 for all x E K \ U. Then 0 :::; h :::; Xu, so that (3(Xu) ::::: (3(h) = e(h) > 0 because e is isometric and order-preserving. (iii) Let f E B(K). By Corollary 7.3, Sp B(K)f follows because SPI3(H)(3(J) <:;;; SPB(K)f.
=
f(K), and so the result D
Theorem 7.17 (Borel functional calculus) Let T be a normal operator on a Hilbert space H, and let eT : C(SpT) ---+ 8(H) be the continuous functional
calculus for T. Then eT extends to a (norm-decreasing) *-homomorphism (3T: B(SpT)
---+
{T,T*Yc
<:;;;
8(H).
Further, (3T has the monotone-convergence property, and (3T is unique subject to satisfying this extra condition. Let U be a non-empty, open subset of Sp T. Then (3T(XU) =J: o. Proof This is a special case of Theorem 7.16.
D
Remark on terminology. Let K be a compact metric space (such as SpT). As explained above, the families of Baire and Borel subsets coincide, and the space B(K) of bounded Baire functions on K is the same as the space of bounded Borel functions on K. This is why Theorem 7.17 is almost always called the Borel functional calculus, although 'Baire' would be just as good.
Introduction to Banach Spaces and Algebras
298
As with the continuous functional calculus, it is quite common to write f(T) as a notation for /3r(f) when f E B(Sp T). We shall use this notation in the applications to follow. Let f E B(SpT). By Theorem 7.16(iii), SPl3(H)f(T) <:;;; f(SpT), and so we have go
f
E
B(Sp T) whenever 9 E B(f(Sp T)).
Corollary 7.18 Let T be a normal operator on a Hilbert space H. Then
(gof)(T)=g(f(T))
(gEB(f(SpT)),JEB(SpT)).
(*)
Proof Equation (*) certainly holds whenever 9 is a polynomial in Z and Z and f E B(SpT). It follows by uniform approximation on f(SpT) that (*) holds whenever 9 E C(f(SpT)) and f E B(SpT) (and this is sufficient for later applications). The two maps 9 f---+ g(f(T)) and 9 f---+ (g 0 f)(T) from B(f(SpT)) to B(H) both have the monotone-convergence property, and so the result follows 0 from the uniqueness clause in the theorem. 7.5
Applications of the Borel functional calculus.
Invariant subspaces. Let E be a Banach space, and let T E B(E). We have defined a proper invariant subspace for T in Section 4.6. As we explained, we are hoping to show that, when dim E ~ 2, many such T have such a proper invariant subspace. We have noted that this is true for operators T with an eigenvalue; also, by Corollary 4.97, it is true whenever Sp T is a disconnected subset of C Thus we are now mainly concerned with the case where Sp T is connected. We shall show, using the Borel functional calculus, that normal operators on Hilbert spaces have proper hyper-invariant subspaces. Recall that a closed subspace F of a Hilbert space H is a hyper-invariant subspace for T E B(H) if S(F) <:;;; F for each S E B(H) with ST = TS.
Theorem 7.19 Let H be a Hilbert space with dim H ~ 2, and let T be a normal operator on H with T tJ. CIH. Then there is a proper hyper-invariant subspace forT. Proof Suppose that Sp T has an isolated point. Then, by Corollary 6.28, T has an eigenvalue -\, and, as in Section 4.6, the corresponding eigenspace, E(-\), is a closed subspace of H such that E(-\) is invariant for each S E B(H) with ST = TS. Assume that E(-\) = H. Then T = AIH, a contradiction. Thus E(-\) is a proper hyper-invariant subspace for T. We may suppose, therefore, that Sp T has no isolated points. (In fact, this implies that Sp T is uncountable-but the rest of the argument works just on the hypothesis that Sp T has at least two points.) In this case we may find an open subset U of C such that both Un K and K \ U are non-empty. Then P := f3T(XU) is a self-adjoint operator such that p 2 = P. By the last sentence of Theorem 7.17, P =1= 0 and P =1= I H .
299
The Borel functional calculus
Let F = P(H). Then F is a closed subspace of H, with F =1= {O} and F =1= H. We have P E {T, T*} CC. Now suppose that S E B(H) with ST = TS. Then, by Fuglede's theorem, Theorem 6.35, ST* = T* S, and so S E {T, T*} C. Thus SP = PS, and so S(F) <;;; F. This shows that F is a proper hyper-invariant subspace for T. D
Polar decomposition. In Corollary 6.40, we gave a polar decomposition of an invertible element in a unital C * -algebra. We shall now give a similar decomposition of an arbitrary normal operator. Theorem 7.20 (Polar decomposition) Let H be a Hilbert space, and let T be a normal operator on H. Then there exist R E B(H)+ and U E U(B(H)) such that T = RU, and R, U, and T are pairwise commuting. Proof Define bounded Borel functions rand u on Sp T by
r(A) = IAI,
U(A) = AlIAI (A
=1=
0),
u(O) = 1
for A E C. Notice that r E C(SpT) and that u E B(K) by Proposition 7.4. Let R = r(T) and U = u(T). Then R E B(H)+ and U E U(B(H)). Since Z = ru, we have T = RU = UR. Further, RT = R 2U = RUR = TR and UT = U RU = TU, and so R, U, and T are pairwise commuting. D
Unitary operators as exponentials. In Section 4.15, we defined e a = expa for an element a of a unital Banach algebra A. Proposition 7.21 Let H be a Hilbert space, and let U E U(B(H)). Then there exists Q E B(H)sa such that U = eiQ . Proof Since U is unitary, Sp U <;;; 'JI'. There exists a bounded, real-valued function f on 'JI' with eij(z) = Z (z E 'JI') and such that f is continuous on 'JI' save at Z = 1. By Proposition 7.4, f E BJR.('JI'). Set Q = f(U) E B(H)sa. Then e iQ = U by Corollary 7.18. D
The group of units of B(H). Let A be a unital Banach algebra. Then Go(A) is the principal component of the group of units of A; by Theorem 4.105(ii), Go(A) consists of the finite products of exponentials. Theorem 7.22 Let H be a Hilbert space. Then the group of all invertible operators in B(H) is connected, and every invertible operator is a product of two exponentials. Proof Let T E G(H). By Corollary 6.40, T = RU, where R = (TT*)1/2 is an invertible positive operator and U is unitary. Since Sp R c ~+., we can apply Proposition 6.33 to see that R = e S for some S E B(H)sa. Also, by the previous application, U = e iQ for some Q E B(H)sa. Thus T = e S e iQ , a product of two exponentials. The mapping t f-+ e tS e itQ , IT -+ B(H), is a path in G(H) from IH to T, so that G(H) is connected. D
300
Introduction to Banach Spaces and Algebras
Let H be a Hilbert
7.6 The spectral theorem for normal operators. space, and let T be a normal operator in B(H); as above, (3T : B(SpT)
---->
A = {T} cc
is the Borel functional calculus for T. For every Borel subset E of Sp T, the characteristic function XE is an idempotent element of BIR(SpT), so that
P(E)
:= (3T(XE)
is a self-adjoint, and hence orthogonal, projection contained in A. Evidently P(0) = 0 and P(SpT) = I H ; also, P(U) #- 0 for each non-empty, open subset of U of SpT. Let (Ekh>l be a sequence of pairwise-disjoint Borel subsets of SpT, and set E = U Ek~ Then the partial sums of the series 2::r=1 XE k increase pointwisemonotonically to the Borel set XE. By Theorem 6.49 and the fact that (3T satisfies the monotone-convergence property, (2::~=1 P(Ek) )n21 is an increasing sequence of projections converging in the strong-operator topology of B(H) to
P(E) = P
(U
Ek) = sup
k=l
nEN
t
P(Ek).
k=l
(For m #- n, P(Em ) ~ P(En ) because XEm . XEn = 0, so that P(Em)P(En) = 0.) The correspondence E f---+ P(E) is then a kind of 'projection-valued Borel measure' on Sp T. We shall not formalize this concept, but shall instead note that it follows immediately from the remarks just made that: for every x E H, the mapping
E
f---+
JLx(E)
:=
(P(E)x, x),
B(SpT)
---->
1R,
is a non-negative, Borel measure on Sp T. We wish to show that, for every .1: E H, we have
(Tx, x) =
r
Z d/Lx,
JSpT
where Z here denotes the coordinate function ). f---+ ). on Sp T. This is often written symbolically (or literally, if the appropriate integral has been defined), as
T =
r
Z dP =
JSpT
r ). dP-x .
(** )
JSpT
The proof of (**) is easier than you might expect. The secret is to be more ambitious, and to prove that, for every 9 E B(SpT), we have ((3T(g)X, x)
=
r
iSpT
9 d/Lx
(x
E
H)
(***) .
301
The Borel functional calculus
For this, consider, first, the very special case in which 9 = XE, where E is a Borel subset of SpT. Then, by definition, f3r(g) = f3T(XE) = peE), and so
(f3T(g)X, x) = (P(E)x, x) = fLx(E) =
r
IE dfLx .
JSpT
By the linearity of f3T, it then follows that (***) holds whenever 9 is any simple Borel function on Sp T. But every 9 E B(Sp T) is the uniform limit of simple Borel functions, and so the general result follows. The special case (**) follows by taking 9 = Z, since f3T(Z) = T. Notes For a related approach to the Borel functional calculus, see [128, Section 4.5J; for a different approach, see [102, Section 5.2J. There are many accounts of the spectral theorem for normal operators; see, for example, [64, Section X.2J, [88], [102, Section 5.2], [144, Chapter 12J. Our account of this calculus has been functional-analytic, and has avoided measure theory; most other accounts depend on the theory of regular Borel measures. There has been a substantial industry of obtaining results analogous to the spectral theorem for operators on a Hilbert space that are not necessarily normal and for operators on Banach spaces that are not Hilbert spaces; there are many problems to be overcome. A very important thread is the study of spectral operators; see [65, 114], where many applications of spectral theory can be found. Despite vast effort by many outstanding mathematicians, it is still an open question whether a general operator l' E B(H) has a non-trivial, closed invariant subspace. (Here, H is an infinite-dimensional, separable Hilbert space.) This is the famous invariant subspace problem for Hilbert spaces. Counter examples are known in the analogous case where l' is an operator in B(E) for a Banach space E that is not a Hilbert case: these examples, which are very sophisticated, are due to Enflo [68J and Read [138, 139J. There is a huge number of partial results on the above invariant subspace problem. For accounts, see [39, 71, 131], for example. Here is one striking theorem, from [36J. Let T E B(H) satisfy 111'11 ::; 1, and suppose that 'f <;;; SpT. Then l' has a proper invariant subspace. A generalization of this theorem is given in [14J. Exercise 7.7 Let H be a Hilbert space, and let l' be a normal operator in B(H). (i) Use the polarization identity to show that, for each x, y E H, there exists a non-negative, Borel measure /lx,y on on SpT such that
(f3r(g)x,y)
=
r
gd/lx,y
(x,y E H)
iSpT
for every g E B(Sp T). (ii) Show that, for each E > 0, there is a finite set {PI"'" Pn } of pairwise orthogonal projections in B(H) with '£;=1 PJ = IH and >\1, ... , An E C such that
IIT-'£;=IAJPJII ::; E. Indeed, for j = 1, ... ,n, take PJ = XJ(T), where XJ is the characteristic function of a suitable 'half-open' square in C. (iii) Show that the linear span of the orthogonal projections is II . II-dense in B(H). Exercise 7.8 Let H be a Hilbert space, and take T E B(H) with 0 ::; T ::; 1. Find a sequence (Pn )n2';1 of pairwise commuting projections such that T = '£~=1 2-n P n .
s
seJqa~le 4:l eue pue salqe!JeA xaldwo:l leJaAas
III lJed
8
Introduction to several complex variables
In Section 4.15, we developed the holomorphic functional calculus for a single element of a unital Banach algebra. In the final part of this work, we shall develop the important analogous result: a 'several-variable holomorphic functional calculus' for finitely many elements of a unital (commutative) Banach algebra. The single-variable functional calculus used a number of results from the theory of analytic functions of one complex variable; these results are covered in undergraduate courses, and were assumed to be known. Naturally, the several-variable functional calculus will rely on results from the theory of analytic functions of several complex variables. These results are considerably harder, and are rarely covered in undergraduate courses. Thus we shall first give an account of this theory that is sufficient for our purposes-and indeed goes a little beyond what is strictly necessary.
Differentiable functions in the plane Before looking at functions of several complex variables, we shall need to look at some properties of smooth functions in the plane. Again, we shall do a little more than is strictly needed for the applications to follow because the results seem to be of considerable interest in themselves. 8.1 Theorems of Green and Cauchy. Let U be a non-empty, open subset of the complex plane C. Then the space of functions j : U ----+ C such that both the partial derivatives oj lox and oj loy exist on U is denoted by P D l(U). It is clear that P D 1 (U) is an algebra of functions on U. For j E P D 1 (U), we define the following extremely convenient notation:
oj == oj == ~ (OJ _ i OJ) , oz 2 ox oy
8j== oj Oz
==~(Oj +i Oj ), 2
ox
oy
where z = x + iy E U. Clearly, a and 8 act as derivations from P D 1 (U) into CU. We also introduce the space C 1 (U) of functions in P D 1 (U) for which oj I ox and oj loy are both continuous on U. A function j : U ----+ C is real-differentiable at a E U if and only if there is a linear function La : ]R2 ----+ ]R2 and a function c : U ----+ C such that
j(z) = j(a)
+ La(z - a) + c(z)lz - al (z
E U),
(*)
306
Introduction to Banach Spaces and Algebras
where E( z) ----> E( a) = 0 as Z ----> a in U. The space of real-differentiable functions on U is denoted by D l(U), so that D l(U) is an algebra of functions on U. Clearly, Di(U) ~ PDi(U), and
La(z - a) = (z - a)af(a)
+ (z -
a) 8 f(a)
(z E U)
for fED 1 (U). It is a standard exercise that C 1 (U) ~ D 1 (U) ~ C(U). However, there are real-differentiable functions whose partial derivatives are not continuous. The Cauchy-Riemann criterion for analyticity becomes: fED 1 (U) is analytic (i.e. complex-differentiable) on U if and only if 8 f = 0 on U. Moreover, in the case where f is analytic on U, we have of = 1', the ordinary complex derivative of f, on U. We now come to a form of Green's theorem. We shall give a proof under conditions that make clear its link with Cauchy's theorem, Theorem 1.32(i). The area (or two-dimensional Lebesgue measure) of a compact plane set K is denoted by rn(K). Theorem 8.1 (Cauchy-Green theorem for a rectangle) Let R be a closed rectangle in
hR f(z) dz = 2i f
l
8f
dx dy.
Remarks. 1) Under the slightly stronger hypothesis that f has continuous partial derivatives of jax and of joy on R, this is a special case of the usual Green's theorem in the plane; this latter version would actually be sufficient for our purposes. 2) The theorem is given as stated since it seems pleasant to include the usual Cauchy theorem for a rectangle (where the assumed condition is 8 f = 0 on R) as a special case. Proof For every subrectangle S of R, define
w(S) = hsf(Z)dZ-2i ffs8fdXdY. Let Iw(R)1 = k 2: 0; we need to show that k = O. We first quadrisect R into four subrectangles, say R i , ... , R 4 . Clearly, we have w(R) = L~=i w(RJ ), so we can choose one of the quarter-rectangles-call it R(l)-with Iw(R(1))1 2: kj4. Now iterate the quadrisection process, obtaining a decreasing sequence of closed rectangles
R = R(O) :2 R(i) :2 R(2) :2 ... :2 R(n) :2 ... , with Iw(R(n))1 2: kj4n (n EN). Clearly, nn>i R(n) is a singleton, say {a}, where a E R. Since f is real-differentiable at a, we may suppose that (*) holds; for n E N, define En = SUp{IE(Z)1 : z E R(n)}, so that En ----> 0 as n ----> 00.
307
Introduction to several complex variables
As in the usual proof of Cauchy's theorem for a rectangle, we do simple direct calculations to show that:
r (J(a) JaR(n)
+ (z -
a)f)J(a)) dz = 0,
r (z - a) dz = 2i m(R(n») JaR(n)
=
2ibc/4n ,
say, where R has sides of length band c. It follows that
Iw(R(n»)1 ::; 2 Jr r 18 J(z) - 8 J(a)1 dx dy JWn)
+ 2cn (b + c)(b 2 + C2 )1/2 4n
(n
E
N).
Fix ry > 0. Since 8 J is continuous at a, we see that there exists no E N such that 18 J(z) - 8 J(a)1 < ry (z E R(n») for each n ;::: no; also, we may suppose that Cn < ry (n;::: no). We then deduce that
Iw(R(n»)1 ::; 2ry bC 4n
+ 2ry (b + c)(b2 + c2 )1/2 4n
Cry
4n
(n;:::no),
°: ;
say, where C is independent of nand ry. Thus k/4n ::; Cry/4 n (n;::: no), so that k ::; Cry. This holds for each ry > 0, and so k = 0, as required. D
°: ;
We now apply Theorem 8.1 to obtain a version of Cauchy's integral formula. First, recall that the function z f---' 1/ z is locally integrable on C with respect to two-dimensional Lebesgue measure, in the sense that
JJK-lzlr dxdy
<
00
for each compact subset K of C; to see this, use polar coordinates to evaluate the double integral over a disc, centre 0, that contains K. Corollary 8.2 Let R be a compact rectangle in C (with the sides oj R parallel to the axes). Let JED l(U), where U is an open neighbourhood oj R, and suppose that (8 J) IRE C(R). Let a E int R. Then
J(a)=~
r
J(z)
2m JaR z - a
dz-~Jrr 7r
8 J (z)dxdy. JR z - a
Proof For every sub rectangle S of R, define
w(S)
=
r J(z) dZ-2iJrr 8J(z) dxdy. Jasz-a Js z-a
For some 6 E (0,1]' the point a is at the centre of a square, say R a, with sides, of length 6, parallel to those of R and which is contained in the interior
308
Introduction to Banach Spaces and Algebras
of R. By extending the sides of Ro to meet those of R, we divide R into nine subrectangles, Ro and Rk (k = I, ... ,8), of which only Ro contains the point a.
~:
I R, I I
R8
IR
Then, by applying the above theorem to each of RI, ... , Rs for the function z f---> f(z)/(z - a), we see that w(RJ ) = 0 for j = 1, ... ,8, so that w(R) = w(Ro) for all sufficiently small S > o. U sing the analyticity of the function z f---> 1/ (z - a) on C \ {a}, the local integrability of z f---> 1/ z, and the continuity of f on R, it is simple to estimate that
a
flo a (!~~)
I
Iflo ~~;
dXdyl =
dXdyl
~ MIS,
say, where MI is a constant independent of S. Since f satisfies (*) on page 305, we estimate that, for each z E oRo, we have
- f(a) I ~ Jof(a)J +I(:2 - a) af(a) 1+ Jc:(z)J ~ M Ifez)z-a z-a
2 ,
say, where M2 is a constant independent of S. Thus
r
lim I fez) - f(a) dzl 0->0 JaR" z- a and so
.1
~
lim 4M2 S = 0, 0->0
1
fez) . f(a) . hm - dz = hm - dz = 27rlf(a). 0->0 aR" z - a 0->0 aR" z - a Hence
w(R) = w(Ro) =
r
fez) dz JaR" z - a
2iJrJR"r a(f(Z) ) dxdy z- a
---->
27rif(a)
o
as S ----> 0, from which the result follows. Corollary 8.3 Let fED on Co Then
I
f (w) = -
(q,
~ 7r
with supp f compact and with
J-JIJII.2r af (z) dx dy Z -
(w
E
af
continuous
q .
W
Proof Let w E C, and take a rectangle R such that supp f U { w} <:;;; int R. Since fez) = 0 (z E oR), Corollary 8.2 immediately gives the required result. 0
309
Introduction to several complex variables
8.2 Existence of smooth functions. Before going further, it is useful to describe briefly the existence of smooth functions (and these will be very much needed in the next chapter). Throughout this section, n is fixed in N. A complexvalued function f on a non-empty, open subset U of C n is called smooth provided that it has continuous partial derivatives of all orders with respect to the underlying real coordinates. Using the alaz and alOi notations described at the start of this chapter~but now using all the coordinates Zl, ... , Zn ~we are requiring that all the partial derivatives ar+s f Iar z) asz) exist and are continuous. The space of smooth functions on U is denoted by C=(U); clearly C=(U) is a subalgebra of C l(U). By well-known 'advanced calculus techniques', it follows that each of these partial derivatives is itself a smooth function, and that the various partial differentiation operations are mutually commuting. We shall sketch a few elementary facts about the existence of smooth functions. We successively define functions 0:, (3, "( : lR --+ H by setting:
o:(x) =
{~-l/X
if x> 0, if x ::; 0,
(3(x)
o:(x)
=
(xElR),
and then let
"((x) = (3(x
+ 2) (3(2 -
x)
(x
E
lR).
Then,,( is smooth on lR, "((lR) <:;; H, "((x) = 1 (Ixl ::; 1), and "((x) = 0 (Ixl ::::: 2). Now, for z E cn, define B(z) = "((llzl!), where 11·11 is the Euclidean norm on cn. Then B is smooth. B(z) = 1 (lizil ::; 1), and B(z) = 0 (11zll ::::: 2). Let f be a (real, complex or, more generally, vector)-valued function on a topological space X. Recall that the support of f, denoted by supp f, is the closure in X of {x EX: f(x) =I- a}. Lemma 8.4 Let U be an open neighbourhood of a non-empty, compact subset K of C n . Then there is a compactly supported, smooth function X on C n such that X I K = 1, supp XC U, and X(C n ) <:;; H. Proof By an elementary compactness argument, we can find and 1'1 ... ,rk > 0 such that
UB(z); 1')) )=1
K
k
k
K C
Zl ... ,Zk E
and
UB(z); 21')) C U, )=1
where B( w; 1') denotes the Euclidean open ball with centre wand radius l' in C n . For j = 1, ... ,k, set B) (z) = B( (z - z) )11')) (z E C n ). Then each B) is smooth on cn, B)(C n ) <:;; II, B)(z) = 1 for z E B(z);r)), and B)(z) = 0 for z tf. B(z);2r)). It is easily checked that the smooth function X := 1-(I-B1)(I-B2) ... (I-Bk) has the required properties. 0 We shall now state, without proof, the following more elaborate existence theorem for smooth partitions of unity.
310
Introduction to Banach Spaces and Algebras
Theorem 8.5 Let U be a non-empty, open subset of cn, and let {Un}nEA be an open covering of U. Then there is a family {en}nEA of smooth functions on U such that: (i) en(U) <;;; II (0: E A); (ii) on each compact subset of U, all but finitely many of the functions en are identically zero;
(iii) suppen
C
Un (0:
(iv) LnEA en(z)
=
E
1 (z
A); E
U).
0
It follows from (ii) that, for each z E U, en(z)
-I-
0 for only finitely many
0: E A, and so the sum in (iv) is well defined. The family {en}nEA is called a smooth partition of unity subordinate to the covering {Un}nEA. Lemma 8.6 (Inhomogeneous Cauchy-Riemann equations, with compact support) Let f E COO(q be such that supp f is compact, and set g(w)
1 = --
J~ 1Ft2
7r
Then 9 E COO(q and 8g =
fez) -
z-
dxdy
W
(W
Eq.
f on C
Proof Let wEe We use the change of variable z expression for 9 as g(W)
=
-~ 7r
f-+
z
+W
to write the
Jr r fez + w) dxdy. l1Ft2
Z
Then an easily justified 'differentiation under the integral sign' using the local integrability of 1/ z shows that 9 is smooth and that 8g(w) =
_~ 7r
and so 8 g( w) 8.3.
Jrr
l1Ft2
8f(z+w) Z dx dy ,
= f (w) by reversing the change of variables and applying Corollary 0
We remark that, on the open set C \ supp f, we have 8 9 = 0, so that 9 is analytic on that set. It is generally not possible to choose 9 to have compact support. Corollary 8.7 Let U be an open neighbourhood of a non-empty, compact subset K of C, and let f E COO(U). Then, for each open subset V with compact closure and such that K eVe V c U, there exists 9 E COO(V) with 8g = f on V. Suppose that the function f depends also on some additional complex variables, say on (Wi, ... ,Wk), and that f is smooth in (Z,Wi, ... ,Wk) and analytic in certain of the variables wJ • Then 9 may be chosen also to be smooth in (z, Wi, ... , Wk) and analytic in the corresponding variables Wj.
311
Introduction to several complex variables
Proof By Lemma 8.4, there is a smooth function e on C with compact support suppe C U and with e(z) = 1 (z E V). Then ef (defined to be a on C\suppf), is a compactly supported, smooth function on C. If we also allow extra variables (WI, ... ,Wk), as in the second part of the statement, then we define
1 g(Z,WI, ... ,Wk) = - 7f
J1
dc;d1] e(z)f(Z,WI, ... ,Wk)-;--
]R2
.., -
Z
(z
E
q,
where (= c; +i1] E C. By Lemma 8.6, 8g = e· f on C, so that 8g = f on V. The statements about smoothness and analyticity follow by differentiation under the integral sign. D 8.3 Cauchy-Green formula. Although not strictly needed, we should like to complete the 'Cauchy-Green' approach that we have been following by giving a version of Cauchy's integral formula that applies to real-differentiable functions; it does, of course, also provide a proof of the more familiar Cauchy integral formula of the winding-number variety. We shall use the notion of a contour as introduced in Section
1.11.
Theorem 8.8 (Cauchy-Green formula) Let U be a non-empty, open subset of C, and let"( be a contour in U such that nh; w) = 0 (w E C\ U). Let fED leU) with 8 f E C(U). Then
nh;w)f(w)=~l 2m
Proof Let
W
E U \
'Y
fez)
z- W
dZ-~Jrr 7f
Ju
n h ;z)8 f (z)dxdy
z- W
(wEU\b]).
b]' and define K =
bl U {z
E C : nh; z) =1= O} U {w}.
Then K is a compact subset of U. By Lemma 8.4, there is a compactly supported, smooth function e, with e == 1 on a neighbourhood of K and with supp e c U. Then, since f = e . f on a neighbourhood of K and since nh; z) = 0 on U \ K, we may replace f bye· f in the theorem. Thus, without loss of generality, we may suppose that f itself satisfies the condition that supp feU. Now, for every ( E bl and W E U \ b]' by applying Corollary 8.3 to f, we have fee) - few) = r 8 fez) dxdy (-w 7f((-w) Ju z-( z-w
-1 Jr
=
_~Jr{ 7f
Ju
(_1___1_)
8f(Z) dxdy. (z-()(z-w)
Next, we integrate round ['Yl with respect to (. By an elementary argument (and recalling that l/(z - () and l/(z - w) are locally integrable with respect
312
Introduction to Banach Spaces and Algebras
to m), we may use Fubini's theorem to interchange the path and area integrals, obtaining:
~ 2m
1 'Y
J(() - J(w) d( = ( -
W
-! Je r (~1~) 8 J(z) dxdy 7r JU 27rz 'Y Z - ( z - w
= !Jer n(r;z)8J(z) dxdy. 7r Ju z- w Thus
n(r;w)J(w)=~l 2m
'Y
J(z) dz-!Jer n(r;z)8 J (z)dxdy, z- w W 7r Ju
z-
o
as required. We may immediately deduce a form of Cauchy's integral formula.
Corollary 8.9 Let U be a non-empty, open subset oj C, and let, be a contour m U such that n(r;w) = 0 (w E C \ U). Let J E Dl(U) with 8J E C(U) be such that J is analytic on each component oj {z E U : '11,(,; z) -I- o}. Then
n(r; w)J(w) = _1 27ri
1 'Y
J(z) d z- W Z
(WEU\[r]).
Proof For every z E U \ [r], either '11,(,; z) = 0 or 8 J(z) = o. The integrand in the area integral of Theorem 8.8 is thus zero almost everywhere. 0
Notes Theorem 8.5 is also valid with any finite-dimensional, differentiable manifold in place of an open subset of en. A proof of this theorem may be found in many introductions to differentiable manifolds, or to distribution theory; a good reference is [33, Theorem 10.1]. The Cauchy-Green formula is sometimes called the 'Cauchy-Pompeiu formula'. Exercise 8.1 Let U be a non-empty, open subset of]Rn. Show that C leU) <;;; D leU). Exercise 8.2 Let U be a non-empty, open subset of e, let 'Y be a contour in U such that nCr, w) = 0 (w E e \ U), and let F be a closed subset of U \ b] such that F has no limit point in U. (i) Show that nCr, () = 0 for all but finitely many ( E F. (ii) (Residue theorem) Let 1 E O(U \ F). Show that
!
'Y
I(z) dz = 27ri
L
Res(f, ()nCr, ()
(EF
where Res(f, () is the residue of the function
1 at (.
,
313
Introduction to several complex variables
(iii) (Argument principle) Suppose, further, that f is non-constant on U, has a pole at each point of F, and that Z(J) c U \ b]. Show that n(g 0
/"
0)
=
L mj(()nb, () - L mj(()n(/" (), (EZ
(EF
where Z = Z(f) and mj(() denotes the multiplicity of fat ( for ( of the pole of fat ( for ( E F.
E
Z and the order
Functions of several variables 8.4 Holomorphic functions. We now consider analytic functions of several complex variables. We shall work in the space en, where n E No A generic point of en is
Z = (Zl,'" ,zn)
=
(Xl
+ iYI,'"
,Xn + iYn).
When we denote a point of en by a, say, we understand that a = (al,"" an) as an element of en. Suppose that 1 ::; m ::; n. Then we shall write Zm for the coordinate projection onto the mth coordinate and 7rm ,n for the projection of en onto the first m coordinates, so that
Zm
(ZI"",Zn)
f-+
Zm,
7r
m,n (ZI, ... ,Zm, ... ,zn)
f-+
(ZI"",Zm)'
Let a E en and let I = ('l"""n) E (JR+e)n, so that II, ... ,ln the open polydisc, with centre a and polyradius I, is
6.(a;l) = {z
E
> O. Then
en: IZk -akl < 'k (k = 1, ... ,n)}.
The closed polydisc is the closure of 6.(a; I), namely
6.(a;l)
= {z
E
en: IZk -akl::; 'k (k
= 1, ... ,n)}.
Evidently an open polydisc is a product of open discs, whilst a closed polydisc is a product of closed discs, and is a compact subset of en. In the case of a closed polydisc, we may allow Ik = 0 for some (or even for all) k. We shall also sometimes refer to the open polyball, defined for a E en and I> 0 by
B(a;l)
=
{z
E
en:
IZI -
al1 2 + ... + IZn -
an l 2 <
12}.
Let U be a non-empty, open subset of en. Then the space of those functions --> e such that all the partial derivatives af lax) and af lay) exist on U for k = 1, ... , n is denoted by P D I (U), generalizing the one-dimensional situation. It is clear that P D 1 (U) is an algebra of functions on U.
f: U
314
Introduction to Banach Spaces and Algebras
For a function
JE
P D 1 (U), it is again very convenient to set
i!!L),
ok! == oj == ~(oJ _ OZk 2 OXk 0Yk
ak! == oj == ~ (OJ Ozk 2 OXk
+i
oj ) 0Yk
for k = 1, ... , n. The operators Ok and ak are derivations from P D 1 (U) into
Cu. A function J E P D 1 (U) is real-differentiable if it is differentiable at each point of U as a map into JR2n; here, J is real-differentiable at a E U if and only if there is a function c : U ~ C such that n
J(z) = J(a)
+L
((Zk - ak)okJ(a)
+ (Zk
- (lk)akJ(a))
+ c(z) liz -
all
(*)
k=1
for z E U, where c(z) ~ c(a) = 0 as z ~ a in U and 11·11 denotes the Euclidean norm on JR 2n. The space of real-differentiable functions in P D 1 (U) is denoted by Dl(U); clearly, Dl(U) is a subalgebra of C(U). A function J E Dl(U) is analytic on U if it is complex-differentiable at each point of U, in the sense that, for each a E U, there exist continuous functions L 1 , ... ,Ln : U ~ C such that n
J(z) = J(a) + L(Zk - ak)Lk(z)
(z
E
U).
k=1
It is easily seen that the values of the functions L 1 , . .. ,Ln at the point a are uniquely determined and that, for an analytic function J and k = 1, ... ,17" ak! = 0 and ok! coincides with the ordinary kth complex partial derivative, defined as a limit, so that an analytic function on U is analytic 'in each variable separately'. A classical result of Hartogs gives the remarkable converse of this statement; we shall give only a weaker form of that result-under the unnecessary additional assumption of the continuity of J-in Theorem 8.11, (c)=?(a). The space of analytic functions on U is denoted by O(U), extending the notation of page 114. Clearly, O(U) is a unital subalgebra of C 1 (U). The composition of two analytic functions is analytic, with the usual 'chain rule' giving the derivative of the composition. For example, J E O(U) is invertible if and only if J(z) -I- 0 (z E U). Lemma 8.10 (Cauchy's integral formula for a polydisc) Let U be a non-empty, open subset oj en, and let J E C(U) be analytic in each variable separately.
Then, Jor every closed polydisc .6.(a; r) c U, we have J(Z) = (~) n 2m
Jor Z E .6. (a; r).
1 IA1-all=rl
... 1 IAn-a"l=r n
J(A) dAl ... dAn (AI - zd··· (An - Zn)
315
Introduction to several complex variables
Proof Since f is analytic in each variable separately, repeated application of Cauchy's integral formula in one variable establishes the required formula as an iterated integral. However, the integrand is continuous, and hence integrable, on the domain of integration (a product of circles), with respect to the product measure. The result then follows from Fubini's theorem. D The domain of integration in the above Cauchy integral is sometimes called the distinguished boundary of ~(a; r). As a product of n circles, it is a set ofreal dimension n; it is (for n > 1) a much smaller set than the topological boundary of ~(a; r), which has real dimension 2n - l. We now need some notation involving power-series expansions. Let X = (Xl"",Xn)' where we are thinking of Xl"",Xn as 'indeterminates', and take I = (i l , ... , in) E z+n. Then we set Xl
= X?l1
.. ·X?" n ,
etc' .,
also, III = L~=l ik and I! = i l !i2! ... in!. Then the standard formal power series f = f(Xl, ... ,Xn ) E In := C[[Xl, ... ,Xn]] may be written, using the contracted notation, as f(X)
=
LaIX I
=
L
{a(?l, ... ,?,,)X~'" ·X~n :
(i l , ... ,in) E z+n} ,
I
where aI = a
L
lallrI
<
00
and
f(z) =
L
al(z - a)l
(z E ~(a; r)).
316
Introduction to Banach Spaces and Algebras
We have the following fundamental equivalences. Theorem 8.11 Let U be a non-empty, open subset of en, and let f : U be a function. Then the following assertions are equivalent:
~
e
(a) f is analytic on U, so that f E O(U);
(b) f is real-differentiable on U and satisfies the Cauchy-Riemann equations Elkf = 0 on U for k = 1, ... , n; (c) fEe (U) and f is analytic in each variable separately;
(d) f has a power-series expansion about every point of u. Proof In fact, we shall just sketch the proof of this theorem. First, note that the implications (a) =}(b) =}(c) are all straightforward. Also, that (d) =}(a) follows easily by uniform convergence results. (c) =}(d). Let a E U. By termwise integration of a multi-variable geometric series, we obtain the power-series representation f(z) = I:I O:I(Z - a)I of f on ~(a; r), where
O:I
=
( l)nl 27fi 1)I1- a ll=rl
1 ...
f(A)dAI···dA n IAn-anl=rn (AI - al)""+l ... (An - an)"n+ l
o
for I = (i l , ... , in) E z+n.
The following four corollaries follow easily, essentially as in the one-variable case. Throughout, U is a non-empty, open subset of en. Corollary 8.12 Let f E O(U). Then f has complex partial derivatives of all orders, each being an analytic function on U. The coefficients in the power-series expansion f(z) = I:I O:I(Z - a)I of f about a E U are aI =
..!.. I!
ailif
azI (a)
(I E z+n).
o
Corollary 8.13 (Maximum modulus theorem) Let f E O(U), and take a E U. Suppose that U is connected and that If (z) I :::; If (a) I for all z in some neighbourhood of a. Then f is constant on U. 0 Corollary 8.14 (Uniqueness theorem) Let f,g E O(U). Suppose that U is connected and f(z) = g(z) for all z in some non-empty, open subset of U. Then f =g. 0 Corollary 8.15 Let (fn)n'21 be a sequence in O(U) such that fn on every compact subset of U. Then f E O(U).
~
f uniformly 0
317
Introduction to several complex variables
So far, the theory of analytic functions of several complex variables looks very much like the one-dimensional theory. However, there are striking differences between the two theories. We shall now show one way in which the n-variable theory (for n ~ 2) starts to differ from the one-variable case. In the next proof, we shall use the following easily checked topological fact. Let U be a connected, open neighbourhood of the frontier 811 of an open polydisc 11 c en. Then U n 11 is also connected.
Theorem 8.16 Let U be a connected, open neighbourhood oj 811 Jor an open polydisc 11 c en, where n ~ 2. Then, Jor every J E O(U), there zs a unique FE 0(11) such that F(z) = J(z) (z E Un 11). Proof We may suppose that 11 z = (Zl, ... , zn) E 11, define
F(z)
= _1
1
27ri 1(I=rn
11(0; r) for somer E
(~+.)n.
For each
J(Zl, ... , Zn-l, () Zn d( . (
-
Clearly, F E C(I1) and F is analytic in each variable separately. By Theorem 8.11, (c) =} (a), we have F E 0(11). By the compactness of 811, there exists c > 0 such that Z E U whenever Z E 11 with r1 - c < [Zl[ < '1'1. For such a z, the one-variable Cauchy integral formula gives F(z) = J(z). But then F and J agree on a non-empty, open subset of Unl1, and so, by Corollary 8.14 and the above topological remark, they agree on all of U n 11. 0 It follows from this theorem that, for n ~ 2, an analytic function of n variables cannot have an isolated singularity; neither can it have an isolated zero, since an isolated zero of J would be an isolated singularity for 1/J. Again, let U be a non-empty, open subset of en. We have already remarked that O(U) is an algebra of functions on U. It is clear that O(U) always contains some unbounded function, so that, by Example 4.20, O(U) cannot be given any Banach-algebra norm. There is, however, a very natural complete, metrizable topology on O(U) that we shall now describe. It specifies the topology of local uniform convergence (or uniform convergence on compacta), as on page 114. We can again write U = Um >l K m , where (Km)m~l is a sequence of compact subsets of U and Km C int Km+l (m EN). In particular, for each compact subset K of U, we have K <;;; Km for all sufficiently large m. For mEN, we define
Pm(f)
=
sup{[J(z)[ : z E Km}
(f
E
O(U)).
Then (Pm)m>l is a sequence of submultiplicative semi norms that defines a metrizable locally ;onvex topology on O(U); using Corollary 8.15, the topology is easily seen to be complete, so that O(U) is a F'rechet algebra, as in the case where n = 1.
318
Introduction to Banach Spaces and Algebras
8.5 Differential forms. We shall need a few ideas about smooth differential forms. Since we shall need such forms only on open subsets of en (rather than on more general manifolds), our approach can be quite simple-minded. The main point of the formalism to be introduced lies in the definition of the exterior which will provide an algebraically efficient means of studying derivative integrability conditions for the inhomogeneous Cauchy-Riemann equations. It will be seen that an equation between forms is just a concise way of writing a certain set of partial differential equations. Again, we fix n E N, and let U be a non-empty, open subset of en. A mapping JL : U --+ em is given by an m-tuple (JL1, ... , JLm) of complex-valued functions on U; we shall write JL = (JL1, ... , JLm) as a shorthand for the function specified by
a,
JL(z) = (JL1(Z), ... ,JLm(z))
(z
E U).
We say that JL is smooth (respectively, analytic) if each of JL1 ... , JL m is smooth (respectively, analytic) on U. We shall now try to introduce our ideas about forms. Let I E C=(U). Then we have seen in Theorem 8.11 that I is analytic if and only if ad = 0 (k = 1, ... , n) on U. With this in mind, we let "[l(U) be a direct sum of n copies of C=(U), and define a mapping
a: I
f--4
(ad, ... ,anI),
c=(U)
--+ "[ 1 (U).
a
Evidently, is a complex-linear mapping and kera = O(U). For k = 1, ... , n, we now write dZk
= (0, ... ,0,1,0, ... ,0)
-1
E [
(U),
where '1', the constant function of value 1, is in the kth place. We note that dZk is just a notation for something else, but, of course, it is chosen for a particular -1 reason. Thus we can regard [ (U) as a 'free C=(U)-module'; however is not a C = (U)-module homomorphism. With this notation, a generic element of"[ 1 (U) appears as W = 91 dZ l + ... + 9n dz n ,
a
where 91, ... ,9n E C=(U). For _
aI
I
n
=
E C=(U), we have _
LOkI dZk
=
k=l
n 01 L 0-
k=l
dZk;
Zk
this suggestive formula reveals one reason for the choice of the notation dZk for a certain vector in "[ 1 (U). -0 Set [ (U) = C=(U), and then note that we have the sequence
o
---4
O(U)
-0 ---4 [
(U)
a-I (U),
---4 [
where the second arrow is inclusion, and the sequence is an exact sequence, in the language of Section 3.17, for every open subset U of en.
319
Introduction to several complex variables
We remark that we are really telling only half the story: there is also a mapping 0 based on the operators ojoZk. However, we shall only use the part of the story that we are describing; the full treatment would be very similar. We shall find that we need to consider the so-called inhomogeneous CauchyRiemann equations: given w E £ 1(U), find an element f E £0 (U) with f = w. Notice that this formulation just gives a concise way of saying: given elements gl, ... ,gn E COO(U), find f E COO(U) with ad = gk (k = 1, ... ,n). Clearly, there is at least the following necessary conditions on the gk for this to hold: ogk oge (k,£= 1, ... ,n). Oze Ozk That this is so follows at once from the fact that, for f E COO(U), we have
a
o2f
o2f
OzeOZk
Ozk Oze
(k,£= 1, ... ,n).
(*)
(It should be clear that these relations for the 'symbolic partial derivatives' oj OZk and ojOzk, &c., follow at once from the corresponding well-known properties of the actual partial derivatives.) We shall also be forced to consider a sequence of higher-degree analogues of these inhomogeneous Cauchy-Riemann equations. These are handled very effectively by the formalism of exterior algebra, to which we shall give a brief introduction. It should be realized that, for our present purposes, the point of the algebra is just so that we can handle efficiently the consequences of the above equations (*). Fix n EN, and let X = {e1, ... , en} be a finite set of n distinct elements. For p = 1, ... , n, let Sp be the set of all ordered p-tuples of elements of X such that the p-tuple has suffixes in strictly increasing order, so that Sp = {e'l 1\ e'2 1\ ... 1\ e,p : 1 ::; i1
< i2 < ... < ip ::; n} ,
where we have written e'l 1\ e'2 1\ ... 1\ e,p instead of (e'l , e'2' ... , e,p). Evidently the number of elements in Sp is ISpl= (;)
=~
Of course we identify Sl with X itself. We also set So = {1}, where 1 is an additional symbol. Write S = U;=o Sp, so that lSI = 2n. Now let A == A(Cn ) be the free complex vector space on S as a basis, so that dim A = 2n; let AP == AP(C n ) be the subspace of A spanned by Sp, so that dimAP = ISpl, and set n
A= L:EBAP. p=o A1(C n ) ~ cn.
Note that It is also sometimes useful to define AP(Cn ) to be {O} for each p > n, and we shall do this.
320
Introduction to Banach Spaces and Algebras
For n E N, we write 6 n for the group of permutations of n distinct elements, so that 16 n l = n!. We shall now define a product operation, denoted by /\, on A, making A into a finite-dimensional, complex (associative) algebra. The product /\ is often called the exterior product. First, we define the product of any finite sequence (e", e'2' ... ,e,p) of elements of X, as follows. (i) Set e" /\ e'2 /\ ... /\ e,,, = 0 whenever ir = is for some 1 ~ r < s ~ p. (ii) In the case where all of il, ... , ip are distinct, let a E 6 p be the unique permutation with io-(l) < ... < ia(p) , so that e,"(,) /\ ... /\ e,"(p) ESp. Then we define e" /\ e'2 /\ ... /\ e,,, = c:( a )e'"(l) /\ ... /\ e,"(p) , where c:( a) (= ± 1) is the sign of the permutation a. (Thus our new definition coincides with the earlier symbolism in the case where (e", e'2' ... , e,,,) ESp.) Now suppose that a = e" /\ et2 /\ ... /\ e,p E Sp and b = eJ, /\ eJ2 /\ ... /\ eJ'I E Sq for some p, q E N. Then we set a /\
b = e" /\ e'2 /\ ... /\ e,p /\ eJ, /\ eJ2 /\ ... /\ eJ,/ '
as defined by (i) or (ii), above. The basis element 1 of S is to act as a multiplicative identity for A(C n ). Finally, we extend the product to all elements in A by bilinearity. It is now simple to verify that A(Cn ) is a complex algebra with an identity. Note also that
a/\b=(-l) pq b/\a
(aEAP,bEAq).
Now again take U to be a non-empty, open subset of C n . For p = 0,1, ... ,n, we define a (smooth, complex-conjugate) p-form on U to be a smooth mapping, say W : U ----+ AP(C n ); in this case, the degree of the form is p. Thus, for p :::: 1, a p-form may be written uniquely as:
L
W=
g[ e" /\ e'2 /\ ... /\ e,1' '
1$" <···
where I = (i 1 , ... , ip) and each g[ E COO(U). In accordance with the notation introduced before, we write' dZk' in place of 'ek'. Then the standard expression for a p-form becomes:
L
W=
g[ dz" /\ dZ'2 /\ ... /\ dz,,, .
1$" <···<',,$n
For
f
COO(U) and was above, we define
E
f· w =
L
(f . g[) dz" /\ dZ'2 /\ ... /\ dz,,, .
1$" <··-<',,$n -p
-p
[
Let [ (U)
(U) be the vector space of p-forms on U. For convenience, we also set p > n. It should be clear that, for p = 0 and p = 1, these
= {O} for each
321
Introduction to several complex variables
definitions are the same as those discussed more informally at the beginning of this section, and that in fact £p(U) is a CCXl(U)-module for the above module operation. Note also that we are identifying, for example, dZ 1 /\ dZ 2 E A2(C n ) with the constant 2 - form z f---> 1 . dz 1 /\ dz 2 , U ----+ A2 (C n ). From this viewpoint, we are using the notation
{dz'l /\ dZ'2 /\ ... /\ dz",: 1 ~ i1
< ... < ip
~
n}
to denote both a basis of AP (C n ) (as an (;) -dimensional vector space), and also for the corresponding basis of £p(U) as a free CCXl(U)-module of degree (;). Also, we write n
£(U)
= L EB£P(U) p=o
for the space (module) of all smooth mappings w : U ----+ A(C n ). The space £(U) becomes a complex algebra under the pointwise product of forms:
(w /\ ".,)(z) = w(z) /\ ".,(z)
(w,,,., E £(U), z E U)
The product /\ on £(U) is the exterior product. N ow recall the definition of the exterior derivative namely
8j
=
t :: k=1
dZ k
(! E £0(U))
8
£ 0 (U)
£ 1 (U),
----+
.
k
We next extend the definition of 8 to the whole of £(U). First, for each p 2:: 1, -
-p+1
-P
-P
we define 8 : [, (U) ----+ [, (U). A generic element of [, (U) is, uniquely, a sum of monomial forms, and these monomials have the form
w = g dZ'1 /\ dZ'2 1\ ... /\ dz,l' ' where 1
~
i1
< ... < ip -
8w
-
~
= (8g)
nand g E CCXl(U). For such a form, we define
_
_
_
-p+1
1\ dZ'1 /\ dZ'2 /\ ... /\ dz,p E [,
(U).
Then the map 8 is uniquely extended, by additivity, to a complex-linear mapping ----+ £P+l(U). This having being done for each p 2:: 1, the mapping 8 may then be again uniquely extended by additivity to a linear endomorphism of£(U). It is called the exterior derivative. We have, in particular, the sequence of spaces and linear mappings:
£p(U)
-
[,: {
0
-------+
O(U)
-0
-------+[,
a
-1
a
a
-P
(U) -------+ [, (U) -------+ .•• -------+ [, (U)
~ lP+l(u) ~ ... ~ ~(U)
-------+
o.
322
Introduction to Banach Spaces and Algebras
Example 8.17 We take n
-2 (U) dimE
=
(2) 2
= 1.
= 2,
-
-1
and then calculate 8 : E (U)
+ hdz2
For w = gdz 1
E
-1 E (U),
---->
-2
E (U), so that
where g, hE C=(U), we
have
-8w = (8 - 2 g) dz 2 !\ az1
+ (8- 1 h) dZ 1 !\ az2 = (8h Ozl
-
89 ) dz 1 !\ az2.
Oz2
It thus appears that the definition of 8 w has been so arranged that and only if 82 g = 8 1 h. In particular,
8 2 j = 0 (f
E
8w =
0 if
= 8 2 j /8Z 2 0z 1 .
D
EO(U) = C=(U));
note that this is just a re-writing of the equation 8 2 j / Ozl Oz2
Returning to the general case, for exactly the same reason we have the following lemma. -
-
Lemma 8.18 For the mapping 8: E(U)
---->
-
-2
E(U), we have 8
=
O.
D
In particular, therefore, the sequence E is a complex (so that, as in Section 3.17, the image of each mapping in E is included in the kernel of the next one). We now make the definitions, first, that a p-form w (for p ? 0) on U is 8-closed if and only ifaw = 0, and, secondly, that the p-form w (for p? 1) is 8 - exact if and only if there is some (p - I)-form 7] such that w = 7]. Let ZP(U) and BP(U)
a
be the spaces of all a-closed p-forms on U (for p ? 0) and of all 8-exact pforms (for p ? 1), respectively; for convenience, we also set BO(U) = {O}, so that BO(U) is the image of the mapping 0 ----> O(U) in the complex E. We have remarked that ZO(U) = O(U). The complex E is called the Dolbeault complex of U; the group (in fact, complex vector space)
HP(U)
=
ZP(U)/ BP(U)
(p? 0),
is called the p th Dolbeault cohomology group of U. For us, this will just be a name: we shall not be developing any formal cohomology theory. Our interest will be to prove just that, for suitably shaped open subsets U of we have HP(U) = {O} (or, equivalently, that BP(U) = ZP(U)) for each p ? 1. This says that the inhomogeneous Cauchy-Riemann equations and their higher-degree analogues are solvable on such open sets U. Towards the above programme, we shall need a few comments on the mapping properties of forms; we shall restrict our attention to analytic mappings.
en,
323
Introduction to several complex variables
Theorem 8.19 Let U and V be non-empty, open subsets of en and em, respectively, and let fJ : U ---4 V be an analytic mapping. Then there is a unique mapping fJ * : £ (V) ---4 £ (U) such that: -0
(i) for f E £ (V) == COO(V), we have fJ *(f)
=
f
0
fJ;
(ii) fJ * is a complex-linear mapping; (iii) fJ *(w 1\ "7) = fJ *w 1\ fJ *"7 (w, "7 E £(V)) ; (iv) fJ*(8w)
= 8(fJ*w) (w
E £(V)).
Proof Set fJ = (fJI, ... , fJm), and write (WI, ... , w m) for the coordinates in For a generic p-form won V, say
w
=L
em.
g[ dill'1 1\ dill'2 1\ ... 1\ dw,p ,
[
where g[ E coo(V), I
= (i l , ... ,ip), and 1 :::::
fJ *w = L(g[
0
fJ) 8ll'1
il
< ... < ip::::: m, define
1\ ... 1\ 8ll,p
,
[
and then extend fJ * by additivity to the whole of leV). The proof that fJ * has the required properties is a matter of simple verification. For example, (iv) is essentially just the 'chain rule' for differentiation: first, we see from (i) that
_
fJ*(af) = fJ*
(m L
af
Ow dw)
)=1)
) =
Lm(af Ow
0
)_
_
fJ all) = a(f OfJ)
)=1)
for f E COO(V), and then we extend this to all wE leV).
o
Remarks. (i) This 'dual mapping' fJ * also preserves the degree of forms, i.e. fJ * (£P (V)) ~
£P (U)
for each p ~ O. This is because
8f
is a I-form for each
f E coo(U). (ii) An important special case of the above theorem arises when U is an open subset of V and ~ : U ---4 V is the inclusion map. Then ~ * : £ (V) ---4 "£ (U) is just restriction of forms-i.e. restriction of each coefficient function. We then normally write w I U instead of ~ * (w). Before proceeding to show that HP(U) = {O} (p ~ 1) for certain open sets U in en, we give a simple consequence of such a result; it will be used later.
Theorem 8.20 Let U be a non-empty, open subset of en, and let V and W be open subsets of U with U = V U Wand V n W -1= 0. Take p 2: 0, and suppose that HP+1(U) = {O}. Then, for every w E ZP(V n W), there exist a E ZP(V) and (3 E ZP(W) such that w = a - (3 on V n W.
324
Introduction to Banach Spaces and Algebras
Proof It follows from Theorem 8.5 that there exists a function 0 E C=(U) with O(U) c IT and with supp 0 c V and supp(l - 0) ~ W. (Thus {O, 1 - O} is a smooth partition of unity on U, subordinate to the open covering {V, W}.) -p -p Now define 0:1 E [; (V) and (31 E [; (W) by:
O:I(Z) = {~l and (31(Z) =
-
(z E V n W), (z E V \ W);
O(z))w(z)
{~O(z)w(z) (Z
(z
E E
V n W), W \ V).
Then we do have w(z) = O:I(Z) - (31(Z) for Z E V n W, but, of course, the forms 0:1 and {3J will not generally be 8 -closed; we must modify them to achieve this. On V n W, we have 80:1 - 8 {31 = 8w = o. Thus there is a well-defined -MJ - form 17 E [; (U) specified by setting 17 1 V = a0:1 and 171 W = a{31. We then -2 -2 have a17 = 0 on U because a17 is locally equal to either a 0:1 or a (31. But, by -p hypothesis, 1{P+l(U) = {O}, and so there is some "y E [; (U) such that 17 = aT We then define 0:
= 0:1 -
("y
1
V)
on V,
{3 = {31 - ("y W) 1
on W.
On the intersection V n W, we still have 0: - {3 = 0:1 - (31 = w, whilst on V, for example, we have 80: = 80:1 - 8"Y = 80:1 -17 = o. Thus 0: E ZP(V) and, similarly, (3 This completes the proof.
E
ZP(W).
o
We record what is, for us, the most important special case of the above theorem, that in which p = O. Corollary 8.21 Let U be a non-empty, open subset of <en, and let V and W be open subsets of U wzth U = V U Wand V n Wi' 0. Suppose that 1{ I(U) = {O}. Then, for every f E O(V n W), there exist 9 E O(V) and h E O(W) wzth f=g-honVnW. 0 Notes For general introductions to the theory of analytic functions of several complex variables, see [85, 86, 96, 111, 133]' for example. It is a central fact that there are many key differences between the several-variable theory and the one-variable theory; for an attractive discussion of 10 key differences, see [111, Section 0.3]. For example, a non-empty, open set U in en is a domain of holomorphy if there do not exist non-empty, open sets V and W, with W connected, such that V <;;; W n U, such that W g U, and such that, for each f E O(U), there exists F E O(W) with F I V = f I V. Roughly, U is not a domain of holomorphy if there is a strictly larger open set such that each f E O( U) can be extended to be analytic on the larger set. Certainly, every non-empty, open set in e is a domain of holomorphy. In en where n 2: 2, some sets are indeed domains of holomorphy: this is trivially the case for polydiscs, and it is true for the polyball B(O; 1), but this is not so obvious. However,
325
Introduction to several complex variables
set Ur = {(Z1,Z2) E e 2 : IZJI < r (j = 1,2)} and U = U1 \ U 1/ 2. Then it is easy to see that U is not a domain of holomorphy. There are several deep theorems that characterize domains of holomorphy, for example involving 'pseudoconvex domains'. A compact subset K of U is holomorphically convex if, for each Z E U \ K, there exists IE O(U) with II(z)1 > IIIK; the set U is holomorphically convex if there is a sequence (Km)m~1 of compact, holomorphically convex subsets such that U = U Km. Then U is holomorphically convex if and only if it is a domain of holomorphy. Here is another example of a key difference between e and en for n 2:: 2. The Riemann mapping theorem says that every proper, simply connected, open subset of e is biholomorphic to the open unit disc. However, there is no analogous result in en for n 2:: 2: for example, by a classical result of Poincare, there is no biholomorphic mapping from the unit polydisc onto the unit polyball [111, Appendix to Section 1.4]. Let U be a non-empty, open set in en. Then the Frechet algebra O(U) is functionally continuous, and the character space
Zn = g(Z1, ... , Zn~1). Exercise 8.5 Let U be a non-empty open set in en, and take p, q 2:
a(w 1\ T)) = (aw)
1\ T)
+ (-l)P w 1\ (aT))
(w E lP(u),
T)
o.
Show that
E lq(U)).
Exercise 8.6 Complete the details of the following sketch that there is a subset U of
e 2 such that 1-{ 1 (U) '" {O}. Set U = e 2\ {(O,O)}, an open set in e 2, and write r2 -1
= IZ112 + IZ212. Define win
£ (U) by setting
w =
a ( Z1Z2r2 )
(Z1 '" 0) ,
w
=-
a( z2r Z1
To show that w is well defined, note that (z2/z1r2)
Z1Z2 '" O. It is immediate that w is a-closed.
2
)
(Z2 '" 0) .
+ (Zl/Z2r2)
1/z1z2 when
326
Introduction to Banach Spaces and Algebras
Assume towards a contradiction that w is a-exact, so that there exists f E COO(U) with f = w. Define
a
g(Z1,Z2)
= z1/(Z1,Z2) -
Z2 2" r
((Z1,Z2)
E
U).
Then 9 E COO(U) and, for Z1 f= 0, we have (a g)/ Z1 = a(g/ zd = a f - w = o. It is easy to see that 9 extends to be an analytic function on all of (:2. However, for each Z2 E (: with Z2 f= 0, we have g(0,Z2) = -1/z2. This is a contradiction. In fact, it can be shown that H 1 (U) is not a finite-dimensional space. On the other hand, H 1(:n \ {(O, ... , O)}) = {O} for n :0:: 3. Exercise 8.7 Set U = {(z,w) E (:2:
Izl < 1,
1/2
< Iwl < I},
V = {(z,w) E (:2:
Izl < 1/2, Iwl < I}.
Show that each analytic function on U U V has a unique extension to an analytic function on the polydisc b.(0; (1, 1», so that U U V is not a domain of holomorphy.
Polynomial convexity
en,
Our current aim is to prove that, for suitably shaped open subsets U of we have H P (l (U») = {O} for each p ~ 1. In the first subsection below, we shall prove this when U is a polydisc; later, in Corollary 8.35, we shall extend this result to polynomial polyhedra.
8.6 The DolbeauIt complex of a polydisc. Again, throughout this section, n E N will be fixed. We begin with a preliminary proposition.
en
Proposition 8.22 Let G be a non-empty, open subset of with G = U:=1 K m , where Km is compact and Km C int Km+l for mEN. Suppose also that: (i) for each mEN, each open neighbourhood U of K m , and each wE ZP(U) -p-l
(where p ~ 1), there are an open set V with Km C V <:;; U and'f/ E [; (V) with 8'f/ = w on V; (ii) for each mEN, each open nezghbourhood U of K m , and each f E O(U), there is a sequence (hkh>1 in O(G) such that hk --> f uniformly on some neighbourhood of Km. Then HP(G) = {O} for all p ~ 1. Proof Take W E ZP(G), where p :0:: 1. The cases p differently. Case p ~ 2. For any given mEN, hypothesis (i) on a neighbourhood of Km with 8'f/m = w. Now take neighbourhood of Km+1 with a'f/:"+1 = w. Then, near
8('f/:"+1 - 'f/m) = w - w =
:0:: 2 and p = 1 proceed gives a (p - 1) -form 'f/m 'f/:"+1 to be defined on a K m , we have
o.
So, again by (i), there is a (p - 2) -form, say ~m' defined on a neighbourhood of Km with 8~m = 17:"+1 - 17m near Km· Indeed, if we multiply ~m by a smooth
327
Introduction to several complex variables
function of compact support in G that is identically equal to 1 near Km (such a -p-2 smooth function exists by Lemma 8.4), we may even suppose that ~m E £ (G), with 8 ~m = '17~+ 1 - '17m near K m· Take 'l7m+1 = '17~+1 -8 ~m on a neighbourhood of K m+1. Then 'l7m+1 = '17m near Km and 8 '17m + 1 = 8'17~+1 = w near K m+1. In this way, we define inductively a sequence ('17m )m> 1 of smooth (p - 1)forms, with each '17m defined on a neighbourhood of K m , satisfying 8 '17m = w on the neighbourhood, and also with 'l7m+1 = '17m on a neighbourhood of Km. -p-1 We may then define '17 E £ (G) by '17 I Km = '17m I Km (m EN). We now have 8'17 = won G, and so the case where p 2': 2 is proved. Case p = 1. We start similarly, with, for each mEN, a smooth function (i.e. with a O-form), say fm' satisfying 8 fm = w on a neighbourhood of K m , and then take a smooth function gm+l such that 8 gm+1 = w on a neighbourhood of K m + 1 ; these smooth functions exist by the case p = 1 of hypothesis (i). Then 8(gm+1 - fm) = 0 near K m , so that gm+l - fm is an analytic function on a neighbourhood of Km. By hypothesis (ii), there is, for each mEN, a function hm+1 E O( G) such that
l(gm+1 - fm - hm+d(z)1 < T m for z in some neighbourhood of Km. Now set fm+1 = gm+1 - hm+1 near K m+1. Then 8fm+1 = 8g m+1 = w near Km+1' and Ifm+1(Z) - fm(z)1 < 2- m for z in some neighbourhood of Km. It is clear that the sequence (fm)m?l that we obtain converges locally uniformly on G, say to a function f. Fix N EN. For each z E K N, we have
f(z) = fN(Z)
+ m-+oo lim (Jm(Z)
- fN(Z)) ,
with uniform convergence on some neighbourhood of K N . But each fm - fN for m 2': N is analytic near KN, so that f(z) = fN(Z) + gN(Z), say, near K N , where gN is analytic. But then f E COO(G), while 8 f = 8 fN = w near K N · This holds for each N E N, so that 8 f = w on G. 0 Theorem 8.23 (The Dolbeault-Grothendieck lemma) Let ll. be a closed polydisc in en, let U be an open neighbourhood of ll., and let w E ZP(U), where p 2': 1. Then there exzst an open neighbourhood V of ll. with V <:;;; U and a form -p-1 '17 E £ (V) such that a '17 = w on v. Proof We may write w =
L WJ dZ
J1
1\ dZJ2 1\ ... 1\ dZJr> '
J
where 1 ::; j1 < ... < jp ::; n, J = (jl, ... ,jp), and each WJ E COO(U). Let v == v(w) be the greatest integer such that the I-form az" occurs explicitly in the expression for w, so that 0 ::; v ::; n. The proof is by induction on v.
328
Introduction to Banach Spaces and Algebras
Case 1: v(w) = o. Necessarily, w = 0, and so we may take V = U and T} = o. Case 2: v(w) ;::: 1, and we make the inductive hypothesis that the result holds for any form w' with v(w') < v := v(w). Then we may write w = -p-l
az
v 1\
a
+ (3,
-p
where a E [ (U), (3 E [ (U), and v(a), v((3) :S v - l. Now 0 = = + (3, and, from the definition of it follows v 1\ that, if rp is any coefficient function of either a or (3, then akrp = 0 on U for k = v + 1, ... , n, so that rp is analytic in (Zv+l, ... , zn). For each such rp, we may find (by Corollary 8.7, which applies because ~ is a polydisc) an open neighbourhood W of ~ with W <;;; U and '¢ E COO(W), with '¢ analytic in (zv+1, ... , zn) and v'¢ = rp on W. Thus, for each coefficient function rp of a, we can find a corresponding ,¢, and we may suppose that all these functions are defined on a common set W by taking a finite intersection of such sets. Next, define a new (p - 1) -form 'Y on W by replacing each coefficient function rp of a by its corresponding '¢. Then a'Y = dz v 1\ a + J, say, where J E [PeW) and v( J) :S v - 1. (We remark that no terms in azv+l, ... ,azn are introduced in this process because each '¢ is analytic in (Zv+l, ... , zn).)
aw -az aa a
a,
a
-
-p
-
-
-2
Let E = w-fh = (3-J E [ (W). Then VeE) :S v-I and oE = ow-a 'Y = o. By the induction hypothesis, there is an open neighbourhood V of ~ with -p-l V <;;; Wand a form E [ (V) such that = E on V. Thus, on V, we have -
-
e
oe
= o'Y + E = ob + e). We can thus take
= 'Y + e E
-p-l
[ (V). We have proved the result for the current value of v, and so the induction continues. 0
W
Corollary 8.24 Let
~
T}
be an open polydisc in
en.
Then
1iP(~) =
{O} for all
p;:::l.
Proof We simply take (Km)m>l to be an increasing sequence of compact polydiscs with Km C int K m+l (~E N) and U:=l Km = ~. Hypothesis (i) of Proposition 8.22 is supplied by the Dolbeault-Grothendieck lemma, whilst hypothesis (ii) follows by considering the power-series expansion of f in a neighbourhood of Km and letting (hk)k?l be the sequence of Taylor polynomials (see Lemma 8.11(d)). 0 We shall now give another application of Proposition 8.22: it provides an improvement to Lemma 8.6 and its corollary. We remark that, for an open subset U of the plane, it is trivial that 1iP(U) = {O} for p ;::: 2; the only question is for p = 1. Theorem 8.25 Let U be a non-empty, open subset of the complex plane C. Then
1i leU) = {O}. Proof We again apply Proposition 8.22. In this case, hypothesis (i) is immediate from Corollary 8.7, whilst hypothesis (ii) follows from Runge's theorem, Theorem 4.83. 0
329
Introduction to several complex variables
Corollary 8.26 Let V and W be open subsets of e with V n W -I=- 0. Then, for every f E O(V n W), there are g E O(V) and h E O(W) with f = g - h on VnW. Proof This follows at once from Theorem 8.25 and Corollary 8.21.
D
8.7 The Cousin problem. This section is not strictly necessary for the main results that we are aiming for, and so some details will be passed over quickly. It does, however, indicate a reason for the importance of the Dolbeault cohomology group 'H l(U). Take n E N, and let U be a non-empty, open set in en. Let us consider an open covering {UaJaEA of U. By Cousin data for this covering, we mean that, whenever a, (3 E A are such that Ua n U{3 -I=- 0, there is a specified function h a ,{3 E O(Ua n U(3), and that these functions satisfy the following conditions:
+ h(3,a = 0 whenever Ua n U(3 -I=- 0; h a ,(3 + h(3" + h"a = 0 whenever Ua n U(3 n U,
(i) h a ,(3
(ii) -I=- 0. The first Cousin problem for the given set of data is to find functions ha E O(Ua ) for each a E A such that ha - h(3 = h a ,(3 E O(Ua n U(3) whenever Ua n U(3 -I=- 0. The open set U is a Cousin domain if the first Cousin problem is solvable in U for each open covering {Ua}aEA of U and each set of Cousin data for the covering. Theorem 8.27 Let U be a non-empty, open set in en. Then U is a Cousin
domain if and only if'Hl(U)
= {a}.
Proof Let {Ua}aEA and {ha}aEA be as specified above. By Theorem 8.5, there are functions () a E Coo (Ua), defined for each a E A, that form a smooth partition of unity subordinate to the covering {Ua}aEA. For each a E A, define
fa =
2: {(),ha"
:I
E
A}
on
Ua ,
where h a " = 0 when Ua n U, = 0. It follows from Theorem 8.5 that, on each compact subset of Ua, fa is given by a finite sum, and so each fa is a smooth function on Ua' Further, in the case where a, (3 E A and Ua n U(3 -I=- 0, we have
fa - f!3 =
2: {(),(ha"
- h(3,,) : I E A} =
(2: {(), : I
E A}) ha,(3
= ha,(3,
and so the Cousin problem is solvable with smooth functions (for any non-empty, open set U in en). Now suppose that 'H l(U) = {a}. Then the proof that U is a Cousin domain is a slight extension of that of Corollary 8.21: we use the fact that 'H l(U) = {O} to modify the functions fa so that they become analytic. Conversely, suppose that U is a Cousin domain, and take w to be a I-form such that 73 w = O.
Introduction to Banach Spaces and Algebras
330 Let
{~a}aEA
en such that
be a family of open polydiscs in
:
U{~a a
E
A} = U.
By Corollary 8.24, for each a E A, there exists fa E COO(Ua ) such that a fa = w. For each a, (3 E A such that Ua n U{3 of. 0, define h a ,{3 E COO(Ua n U(3) by h a ,{3 = fa - f{3; on Ua n U{3, we have a h a ,{3 = a fa - a f{3 = 0, and so, in fact, h a ,{3 E O(Ua n U(3). It is now clear that we have Cousin data for the covering (~a : a E A). Since U is a Cousin domain, there exist ha E O(Ua ) for each a E A such that ha - h{3 = fa - f{3 whenever Ua n U{3 of. 0. We define f(z) = fa(z) - ha(z) (z E ~a) for each a E A. Clearly, f is well defined, f E coo(U), and a f = w. Thus w is a-exact. It follows that 1t I(U) = {O}. 0
8.8 Joint spectra. Before discussing some more general domains in en than polydiscs, we return briefly to Banach algebra theory, in order to motivate the direction that we shall take. This section also includes some simple results that will be needed in proving (and even in stating) results on the several-variable functional calculus. Let A be an algebra, and take n E N. In this section we shall write An for the n-fold Cartesian product of A with itself. We have discussed the spectrum of an element a in A in Section 4.5; we shall now give an n-variable version of the spectrum. We shall restrict ourselves to commutative Banach algebras. Let A be a commutative, unital Banach algebra, and consider an element a = (al,"" an) E An. Then the joint spectrum SPA a = SPA (al,"" an) of a is the subset of en defined by: SPAa = {('P(aI), 00. ,'P(an ))
:
'P
E <J>A}'
Of course, for n = 1, it follows from Corollary 4.47(ii) that we simply recover the spectrum SPAal of al. In fact, from Theorem 4.46, we also have SPAa =
{(>'1'
00.
,An)
E
en :
t
A(Ak 1 - ak)
of.
A} ,
k=l
which is an extension of the original definition of the spectrum. Note that, for a commutative, unital algebra A, a = (aI, ... ,an) E An, and p E qX] = qXI"'" Xn], we can define p(a) =P(al, ... ,an ); the map 8 a : p f---+ p(a), qX] --+ A, is a unital homomorphism such that 8 a (XJ ) = aJ (j = 1,oo.,n). Again, we wish to extend 8 a to have a larger domain. The next lemma summarizes some immediate consequences of the above definition of SPA a; it is an easy exercise.
Introduction to several complex variables
331
Lemma 8.28 Let A be a commutatzve, unital Banach algebra, let n E N, and let a E An. Then: (i) SPAa zs a non-empty, compact subset ofre n ;
(ii) Jrm,n(SPA(al, ... ,an )) =SPA(al, ... ,a m ) (m= 1, ... ,n); (iii) SPA(al, ... ,an ) ~ rr~=lSpA(ak); (iv) SpAP(a) = p(SpAa) (p E qX]).
o
8.9 Polynomial hulls. We now extend to the space ren a definition already made for the case where n ~ 1 on page 179. Let K be a compact subset of ren. Then the polynomial hull K of K is defined to be:
R = {z E ren : Ip(z)1
::; IplK for all polynomials p E qX]} .
The set K is defined to be polynomially convex if and only if K = K. Since the coordinate projections Zm (for m = 1, ... , n) are themselves polynomials, it is clear that R is always compact. Also, K ~ Rand R is polynomially convex. Suppose that Pl, ... , Pr E qX], and set
K = {z E
ren : Ip] (z) I ::; 1
(j = 1, ... , r)} .
If K is compact, then clearly K is polynomially convex. In particular, a closed polydisc is polynomially convex. By Corollary 4.40, a compact subset K of re is polynomially convex if and only if re \ K is connected, and so polynomially convex sets are specified topologically. For K c re n , we may still deduce from the maximum modulus principle, Corollary 8.13, that ren \ K is connected whenever K is polynomially convex. However, the converse assertion may fail when n 2: 2. For example, the compact set K = {(z, 0) E re 2 : Izl = I} is such that re 2 \ K is connected, but R = {(z,O) E re 2 : Izl ::; I} :2 K. This set K is homeomorphic to the set L = {(z, z) E re 2 : Izl = I}, but L is polynomially convex. Recall that a Banach algebra A is polynomially generated by al,"" an E A if A( al, ... , an) = A, in the notation of page 179. The reason why an approach to function theory via polynomial convexity is appropriate for applications to Banach algebras is contained in the following simple result. Theorem 8.29 Let A be a commutative, unital Banach algebra, and let a E An.
(i) S71ppose that A is polynomially generated by al, ... , an in A. Then Sp Aa is polynomially convex, and the mapping
a: r.p >--> (r.p(al)""
,r.p(an )),
A
is a homeomorphism. (ii) In the general case, SPA(al, ... ,an)(a) =
s;:a.
--+
SPAa,
Introduction to Banach Spaces and Algebras
332
Proof (i) Certainly a : q>A ----> SPAa is a continuous surjection. As in Proposition 4.64, is an injection, and hence a homeomorphism. Let A = (AI, ... , An) E en \ Sp A(a). Then there are bl , ... , bn E A with L:~=l bk(Ak 1 - ak) = 1. Since A(al,"" an) = A, there are PI,··· ,Pn E qX] such that
a
111-
~Pk(al, ... ,an)(Ak1-ak)11 < l.
Set q(Zl, ... , zn) = 1- L:~=l Pk(Z)(Ak - Zk) EqX]. Then q(Al, ... , An) = 1. By Lemma 8.28(iv), Iq(z)1 ::::; Ilq(a)11 < 1 (z E SPAa), and so A ~ This shows that SPA a is polynomially convex.
sp:a.
(ii) Trivially, Sp Aa t;;; Sp A(al, ... ,a n ) (a), and so, by (i),
sp:a.
sp:a t;;; Sp A(al, . ,an) (a).
Conversely, take A = (AI, ... , An) E en \ Then there exists P E qX] with Ip(A)1 = 1, but PA(p(a)) = sup{lp(z)1 : Z E SPA a} < 1. But the spectral radius is unchanged in passing to a closed sub algebra (by Theorem 4.23), so that also A ~ SPA(al, ... ,an)(a). 0 Example 8.30 In Example 4.3, we described some standard uniform algebras on a non-empty, compact subset K of C. Now suppose that K is a non-empty, compact subset of en. Then there are entirely analogous examples on K. Indeed, P(K), R(K), and O(K) are the uniform closures in C(K) of the restrictions to K of the algebras of all polynomials on en, of all rational functions p/q, where P and q are polynomials and 0 tI. q(K), and of the functions which are analytic on some neighbourhood of K, respectively, and the algebra A(K) is defined to be A(K) = {f E C(K) : flint K is analytic}.
Clearly, we again have P(K) t;;; R(K) t;;; O(K) t;;; A(K) t;;; C(K). It is easy to see that the character space of P(K) can be identified with the polynomial hull of K, and so 'up to homeomorphic identification' we have q>P(K)
=
SPP(K)(Zl,"" Zn)
= R.
To determine the character spaces of R(K), O(K), and A(K) is more challeng~g.
0
We shall now leave Banach algebras for a while and turn to the study of polynomially convex sets. 8.10 Polynomial polyhedra. Again, throughout this section, n E Ii will be fixed. So far, we have defined polynomial convexity only for compact subsets of en. We now say that an open subset U t;;; en is polynomially convex if and only if, for every compact subset K of U, then also R c u.
333
Introduction to several complex variables
A polynomial polyhedron in
P
{z
=
E ~ :
en
is an open subset P of
IPr (z) I < 1
(r
=
en of the form
1, ... , k)} ,
where ~ is some open polydisc and PI, ... ,Pk are polynomials. The closure of a polynomial polyhedron is compact. Evidently, an open polydisc is an example of a polynomial polyhedron. Lemma 8.31 (i) Every polynomial polyhedron is polynomially convex. (ii) Let U be an open neighbourhood of a compact, polynomially convex set K. Then there is a polynomial polyhedron P such that K cPS;:; U. (iii) Let U be a non-empty, open, polynomially convex subset of Then there exist compact, polynomially convex sets Km for mEN with U = U:=I Km and Km C int K m+ 1 (m EN).
en.
o
Proof This is an easy exercise.
We aim to prove that H PCP) = {O} for p 2': 1 and every polynomial polyhedron Peen. A key step towards this is contained in the following famous lemma. We now write D = {z E e : Izl < I} for the open unit disc in C. Lemma 8.32 (Oka's lemma) Let U be a non-empty, open subset of that p 2': 0 and that HP+1(U x D) = {O}. Take
U'" = {z
E
en . Suppose
U : 1
and define
P:Zf----+(z,
U",----.UxD.
Then, for each wE ZP(U",), there exists n E ZP(U x D) such that w
= p*n.
Proof Note first that p : U'" ----. U x D is an analytic mapping and that p* was defined in Theorem 8.19. Set A = U'" x D and B = {(z, w) E U x D : w #-
= (w -
E
A n B).
Then clearly T/ E Zp(AnB), and so, by Theorem 8.20, there are a-closed p-forms Ct and (3 on A and B, respectively, such that T/ = Ct - (3 on A n B. Hence, on An B, we have
w(z) - (w -
=
It follows that there is a well-defined form
nez, w)
-(w -
n E -p £ (U
x D) defined by:
= {w(Z) - (w -
((z,w)
E E
A), B).
Since w, Ct, and (3 are each a-closed on their respective domains and the map (z, w) f----+ W -
334
Introduction to Banach Spaces and Algebras
Corollary 8.33 Let U be a non-empty, open subset oj en.
(i) Suppose that H 1(U x D) = {a}. Then, Jar each f E O(U'P)' there exists O(U x D) such that J(z) = F (z, 'P(z)) (z E U'P)' (ii) Suppose that HP(U x D) = {a} (p;::: 1). Then HP(U'P) = {a} (p;::: 1).
FE
Proof (i) This is the case p = 0 of Oka's lemma. (ii) Fix p ;::: 1, and take W E ZP(U'P)' Since HP+I(U x D) = {O}, it follows from Lemma 8.32 that there exists D E ZP(U x D) with W = JL*D. But also -p-l HP(U x D) = {O}, so that there is a form 8 E [; (U x D) such that D = B8.
Setting e = JL*8 E lP-I(U'P)' we then have (using Theorem 8.19(iv)):
ae = a(JL*8) = JL*(a8) = JL*D = w.
o
Thus HP(U'P) = {O}. We can now extend Oka's lemma as follows.
Corollary 8.34 Let U be a non-empty, open subset oj en, and suppose that HP(U x Dk) = {O} (p;::: 1). Take 'PI, ... , 'Pk E O(U), set
U'Pl"",'Pk = {z E U: l'Pr(z)1 < 1
(r = 1, ... , k)},
and define JL: z
f--->
(Z,'Pl(Z), ... ,'Pk(Z)) ,
U'Pl"",'Pk
----*
U
X
Dk.
Then: (i) Jor each p;::: 0 and wE ZP(U'Pl, ",'Pk)' there exists D E ZP(U X Dk) such that W = JL*D ; (ii) Jar each f E O(U'Pl,. ','Pk)' there exists FE O(U X Dk) such that J = JL* F, so that J(z) = F (z, 'Pl(Z), ... , 'Pk(Z)) = (F
0
JL)(z)
(z E U'Pl"",'Pk);
(iii) HP(U'Pl, ",'Pk) = {O} (p;::: 1). Proof We shall indicate the proof of just clause (i); clauses (ii) and (iii) will then follow in the same way as Corollary 8.33 followed from Oka's lemma. The proof of (i) is by induction on k. The case where k = 1 is just Lemma 8.32. Thus, take k ;::: 2, and assume that the result is true for sets defined by smaller numbers of the functions 'PJ' Define ILo: (z, w)
f--->
(z, w, 'P2(Z), ... , 'Pk(Z)) ,
Since U'P2," ,'Pk x D = (U X inductive hypothesis gives
Dk. D)'P2, ... ,'Pk and since U x Dk = (U x D) X D k-
JLo(ZP(U x Dk)) for each p ;:::
o.
=
U'P2"",'Pk x D
ZP(U'P2"",'Pk
X
D)
----*
U
X
l ,
the
Introduction to several complex variables
335
As in the proof of Corollary 8.33(ii), HP(U'P2"",'Pk x D) = {O} (p ~ 1). But now Corollary 8.33(i) shows that HP(U'Pl, .. ,'Pk x D) = {O}. Define J.LI: z
>---*
(Z,'PI(Z)) ,
U'Pl"",'Pk
----+
U'P2, .. ,'Pk X D.
Then
J.L{(ZP(U'P2, .. ,'Pk
X
D))
= ZP(U'Pl, ... ,'Pk)
for P ~ O. Now J.L = J.Lo 0 J.LI, and so J.L* = J.Li 0 J.La. This implies the result by showing that J.L*(ZP(U x Dk)) = ZP(U'Pl"",'Pk) for every P ~ O. D Corollary 8.35 Let II be an open polydisc in
en,
let PI, ... ,Pk be polynomials,
and let P be the polynomial polyhedron P = llPl, .. ,Pk := {z Ell: IPr(z)1 < 1 (r = 1, ... , k)}.
Then HP(P) such that
= {O} (p
~ 1). Further, Jor J E O(P), there exists FE O(ll
J(z) = F(Z,PI(Z), ... ,Pk(Z))
(z
E
X
Dk)
P).
Proof Since II x Dk is an open polydisc in e nH , it follows from Theorem 8.23 that HP(ll x Dk) = {O} (p ~ 1). But now the result follows immediately from Corollary 8.34. D
en, and let V and W be open subsets of P with P = V U Wand V n W 1= 0. Then, Jor every J E O(V n W), there exist g E O(V) and h E O(W) with J = g - h on V n W. Corollary 8.36 Let P be a non-empty polynomial polyhedron in
Proof By Corollary 8.35, HI (P) = {O}, and so this is immediate from Corollary 8.21. D
For the following results, we recall that O(U) is a Frechet algebra for the topology of local uniform convergence; topological notions always refer to this topology. Corollary 8.37 Let Q be a polynomial polyhedron m
en,
let PI, ... ,Pk be poly-
nomials, and set P = QPl"",Pk := {z Then, Jor each J
E
Q : IPr(z)1 < 1 (r O(Q
X
Dk) such that
J(Z) = F(Z,PI(Z), ... ,Pk(Z)) = (F
0
J.L)(z)
E
O(P), there exists F
= 1, ... , k)}.
E
(z
E
P).
The map J.L* : F >---* F 0 J.L, O(Q X Dk) ----+ O(P), is a continuous, surjective homomorphism of Frechet spaces, and so there is a topological and algebraic isomorphism O(P) ~ O(Q x Dk)/ ker J.L* .
336
Introduction to Banach Spaces and Algebras
Proof For the existence of F, we note just that Q x Dk is a polynomial polyhedron in n + k , and then apply Corollaries 8.35 and 8.34. It is clear that /1* is a continuous, surjective homomorphism; the conclusion concerning the isomorphism then follows from the open mapping theorem for Frechet spaces, Theorem 3.58. 0
e
In Theorem 8.39, below, we shall identify the space ker /1* of the above corollary. Corollary 8.38 (Oka-Weil theorem) Let U be an open, polynomially convex subset oJ en. Then the set oJ polynomial Junctions on U is dense in 0 (U). Proof First, consider the case of a polynomial polyhedron, say, P = .6. P1 , ... ,Pk' where .6. is an open polydisc in en and PI, ... ,Pk are polynomials. By Corollary 8.35, for each J E O(P), there exists a function F E 0(.6. X Dk) such that J(z) = F(Z,PI(Z), ... ,Pk(Z)) (z E P). By Theorem 8.11, F has a power-series expansion on the polydisc .6. x Dk, say, 21 2,,)1 )k' F( Zl,···,Zn,WI,···,Wk ) -- ~ L....,a21 .. 2,,)1 ... )k ZI "'Zn WI '''W k '
with local uniform convergence on .6. x Dk. Hence
J(Z) =
L a21 ...
)k
Zl1 ... Z~" PI (z)11 ... pk(z)1k
(z
E
P),
with local uniform convergence on P. The partial sums of this series then provide the approximating polynomials to the function J. Now take U to be an arbitrary open, polynomially convex subset of en, and let K be a compact subset of U, so that R c U. By Lemma 8.31(ii), there is a polynomial polyhedron P, with K <:;;: ReP <:;;: U. The result therefore follows from the case already considered. 0 The following theorem uses some notation from Corollary 8.37. Theorem 8.39 Let Q be a polynomial polyhedron in en, let PI, ... ,Pk be polynomials, and set P = Qp1,. ,p<' Take F E O(Q X Dk) such that /1*F = O. Then
there are HI, ... ,Hk
E
O( Q
X
Dk) such that
k
F(z, WI,"" Wk) =
L Hr(z,
WI,""
Wk)(W r - Pr(Z))
((Z, W)
E
Q
X
Dk).
r=l
Proof The proof is again by induction on k. Case k = 1 (Note that this case makes no use of the vanishing cohomology of£(Q x Dk)). We have, say, FE O(Q x D) with
F(z,p(z)) = 0
(z
E
Qp == {z
E
Q: Ip(z)1 < I}).
We must find HE O(QxD) with F(z,w) = H(z,w)(w-p(z)) ((z,w) E QxD). Any such function H (if it exists at all) is clearly unique on each non-empty,
337
Introduction to several complex variables
open subset of Q x D, so that it is sufficient to define H on a neighbourhood of each point of Q x D. Thus, let (zo, wo) E Q x D, and suppose that Wo = p(zo)-since, otherwise, the definition of H is trivial. Now the mapping 8: (z,w) ~ (z,w - p(z)) is an analytic isomorphism oH:n onto itself, with 8(zo, wo) = (zo, 0), and hence F 0 8- 1 is analytic on a neighbourhood of (zo, 0) and (F 0 8- 1 )(z, w) = 0 whenever W = o. Thus, considering the power-series expansion of F 0 e- l about (zo,O), we have that, say, F(e- l (z,w)) = wG(z,w) in some open neighbourhood, say W, of (zo,O), where G E O(W). Thus, on the neighbourhood 8- 1 (W) of (zo, wo), we have F(z,w) = H(z,w)(w - p(z)), where H = Go e E O(8- 1(W)). This proves the result for the case where k = l.
Case k 2: 2, and, inductively, assume that the result is true for all sets defined by smaller numbers of the functions cpJ. We have F E O(Q X Dk), with F(Z,Pl(Z), ... ,Pk(Z)) = 0 (z E QPl, ... ,Pk). But QPl, ... ,Pk = (QPl)P2, ... ,Pk' so that, defining G E O(QPl x D k- l ) by
G(z, W2, . .. ,Wk) = F(z, PI (z), W2, ... ,Wk)
((z, w)
E
QPl X D k- l ) ,
we have G(Z,P2(Z), ... ,Pk(Z)) = 0 (z E (QPl)P2, . .,Pk). Then the induction hypothesis gives functions, say G 2 , •.. ,Gk E O(QPl x D k - l ), with k
G(z, W2, ... , Wk) =
L Gr(z, W2,···, Wk)(W r - Pr(z))
((z, w)
E
QPl X D k- l ).
r=2 By Corollary 8.33(i), there are functions H 2, ... , Hk in O(Q x Dk) such that = H r (Z,Pl(Z),W2, ... ,Wk) on QPl x Dk-l. Thus the function
G r (Z,W2, ... ,Wk)
k
F(z, WI,··· ,wd -
L
Hr(z, WI,···, Wk)(Wr - Pr(z))
r=2 is analytic on Q x Dk and vanishes whenever WI = PI (z). By the case k = 1 of the present theorem, proved above, this function has the form
H l (z,Wl, ... ,Wk)(WI-Pl(Z)) for some HI E O(Q X Dk). Thus the result holds for the present value of k, and so the induction continues. D Notes Theorem 8.23 and Corollary 8.24 are proved in [86, I, Section D] and [96, Theorem 2.3.3], for example. For an extension of this result and of Corollary 8.35, see [111, Section 6.3]. We have discussed the first Cousin problem in Section 8.7. For a more extensive discussion, see [111, Chapter 6] and [133, Chapter VI, Section 4]; in these sources, the second, or multiplicative, Cousin problem is also described. For example, it is shown in [111, Theorem 6.1.4] that a domain of holomorphy is a Cousin domain. For some applications of the Cousin problems, see [153, Section 2.1].
338
Introduction to Banach Spaces and Algebras
A recent book devoted to the study of polynomial convexity in several complex variables is that of Stout [153], where it becomes clear that the subject is complex and that it is hard to tell when sets in en are polynomially convex. For example, it is known that the union of three pairwise-disjoint, closed balls in en is always polynomially convex [153, Theorem 1.6.20], but it is apparently not known whether or not the union of four pairwise-disjoint, closed balls in en is necessarily polynomially convex. Let K be a non-empty, compact subspace of en. The character space of R( K) is discussed in Exercise 8.12, below. However, the character space of O(K) need not be a subspace of any space em [91]. Even if
(ii) Let K = B(O; 1) be the closure of the unit polyball B(O; 1) in en. Show that lo(P(K)) = r(P(K)) = 8K, the topological boundary of K. Exercise 8.10 For r > 0, let Kr = {(z, w) E e 2 : zw = 1, Izl = r}, a set homeomorphic to the circle. Show that each Kr is polynomially convex and that Kl n Kl/2 = 0, but that the polynomial hull of Kl U Kl/2 is the annulus
{(z,w) E e 2 : zw = 1,1::;
Iwl::; 2} .
Exercise 8.11 Let K J = {(z, w) E e 2 : z = jill, Iwl ::; 2} for j = 1,2. Then each of Kl and K2 is a compact, convex subset ofe 2 , and Kl nK2 = {(O,O)}. However, their union, K = Kl U K 2, is not polynomially convex. To see this, consider the map ~:(f--->(,I/(),
e\{0}->e 2
,
so that ~() E Kl when 1(1 = 1 and ~() E K2 when 1(1 = v'2. Take A to be the annulus {( E e : 1 ::; 1(1 ::; v'2}. Show that A ~ R, but that A g K, and so K is not polynomially convex. Exercise 8.12 Let K be a non-empty, compact subset of en. The rational hull of K consists of the points z E en such that If(z)1 ::; IflK (f E R(K)), and K is rationally convex if it is equal to its rational hull. (i) Show that
9
The holomorphic functional calculus in several variables
In this final chapter, we shall extend Theorem 4.89 to analytic functions of several complex variables. This several-variable result has many applications and is certainly the single most powerful method in the study of general commutative Banach algebras; some applications will be given in the second part of the chapter.
The main theorem Preliminary version. Again, we fix n E N in this section. Let A be a commutative, unital Banach algebra, let al, ... , an E A, and write a = (al, ... ,an) E An, in the notation of Section 8.8. Given an open neighbourhood U of SPA a in en, we wish to construct a continuous, unital homomorphism O(U) ----> A such that (again writing Zk for the kth coordinate projection, now restricted to U): 9.1
e:
(i) e(zJ) = aJ (j = 1, ... ,n); (ii) the spectral mapping property holds, i.e. for every
f
E O(U) and
we have
)).
Such a map is a 11Olomorphic functional calculus map. The question of the uniqueness of the holomorphic functional calculus is more complicated than it was in the one-variable theorem. The standard statement asserts a uniqueness, not for each individual but for the whole family of homomorphisms, as a ranges over all ordered n-tuples of elements of A, for all natural numbers n E N and, for each a, with U ranging over all open neighbourhoods of SPAa, subject to the appropriate conditions (i), and to certain compatibility conditions between the different homomorphisms. (In the most usual formulation, the mention of the neighbourhoods U is suppressed by expressing the result in terms of algebras of function-germs; see the notes for details.) The spectral mapping condition (ii) is not, in the classical result, required as a condition for uniqueness: it is an important additional property that is satisfied by each of the unique family of homomorphisms. There is a much nicer uniqueness formulation, due to W. R. Zame; this formulation asserts the uniqueness of each separate e : O(U) ----> A, subject to both clauses (i) and (ii), above, holding; i.e. given e, subject to (i) and (ii), then this e must be just the appropriate member of the unique family of homomorphisms that appears in the standard result. The only problem with Zame's formulation is
e,
340
Introduction to Banach Spaces and Algebras
that (at least so far) its proof requires a great deal of complex analysis in several variables-it would take a second course just to obtain that single improvement in the theorem (although the course would include a lot of good basic complex analysis along the way). In this book, we shall, with regret, dispense with proving any uniqueness statement, except in the simple special case of the next result that deals with polynomially convex neighbourhoods of the joint spectrum. Note that, in the proof of this result, we shall use most of the preliminary theorems on several complex variables that were established in the previous chapter. Lemma 9.1 (Preliminary functional calculus theorem). Let A be a commutative, unital Banach algebra, and let a = (al, . .. ,an) E An. Let U be an open, polynomially convex neighbourhood of Sp a. Then:
(i) there is a unique continuous, unital homomorphism 8 a such that 8 a (ZJ) = a J (j = 1, ... , n);
==
8~ : O(U) -+ A
(ii) for each f E O(U) and each
f (
Proof Notice first that the uniqueness of 8 a , subject to the condition that 8 a (ZJ) = aJ (j = 1 ... , n), follows at once from the density of the polynomials in O(U) given by the Oka-Weil theorem, Corollary 8.38, and the hypothesis of the continuity of the homomorphism. Likewise, the spectral mapping property (ii) also holds for the same reason because we have already established this for polynomials in Lemma 8.28(iv). It thus remains to prove the existence of 8 a . This is broken into several cases. Case (i) U is an open polydisc. By the compactness of Sp a, we may choose an open polydisc 6. = 6.(w; r) such that Sp a c 6. <;;; 6. c U. For each f E O(U) define 8 a (J) to be
( 2~i)n
1 -···1 _ IZ1 -W11- T 1
IZn -Wn I-Tn
f(z)(ztl-ad- l ... (Zn 1 - a n)-l dz l
···
dzn ·
It is clear that 8 a is a continuous linear mapping. Also, as in the one-variable case, _ ZJkJ( ZJ - a J )-ld ZJ = a JkJ (j=I, ... ,n)
1
whenever k J
E
1z.,-wJ I-T J Z+, and so we have
8 a (Zk1 1
n) ... Zk n
= a kl 1 ... ankn
for every monomial Z~1 ··.Z~n. Hence, by linearity, Sa(P) = p(all ... ,an ) for every polynomial p. In particular, therefore, Sa acts as a unital algebra homomorphism on the algebra qX] of polynomials. So, by continuity, Sa : O(U) -+ A is also a unital algebra homomorphism. This completes the proof of case (i).
341
Holomorphic functional calculus
[Note: It is also possible to prove case (i) by showing that it makes good sense to substitute al, ... , an, respectively, for Zl,"" Zn in the power-series expansion for f on ~; the convergence question is settled by using the spectral radius formula from Theorem 4.23.] Case (ii) U is an open polynomial polyhedron. We may suppose that U is the set P = ~Pl' ... 'Pk' in the notation of Corollary 8.35, where ~ is an open polydisc in <en and Pl, ... ,Pk are polynomials. Take r E {1, ... ,k}, and set br =Pr(al, ... ,a n ). Then Spbr = {Pr(z): Z E Spa}. Since Spa we have Spb r
C
C
P = {z E ~: IPr(z)1
D = {z E <e:
Sp (a, b)
=
Izi < I},
< 1 (r
1, ... ,k)},
=
and so, by Lemma 8.28(ii), we also have
Sp (al, ... , an, bl , ... , bk )
c P
x Dk <;;; ~
X
Dk.
Now ~x Dk is an open polydisc in <e n+k , so that, by case (i), there is a continuous, unital homomorphism, say B = 8(a,b) : O(~ x Dk) --'> A, such that, in an obvious notation, B(ZJ)=aJ
(j=I, ... ,n),
B(VVr)
= br
(r
= 1, ... ,k).
For F E O(~ x Dk), set (fJ* F)(z)
= F(Z,Pl(Z), ... ,Pk(Z))
(z E U),
as in Corollary 8.37. By Corollary 8.37, fJ* : O(~ x Dk) --'> O(P) is a continuous, surjective homomorphism, and 0 (P) ~ 0 (~ X Dk) / ker fJ *. There will thus be a (unique) continuous homomorphism, say 'ljJ : O(P) --'> A such that the diagram O(~
X
Dk)
M'i~ O(P)
'ljJ
• A
is commutative if and only if ker fJ* <;;; ker B. By Theorem 8.39, ker fJ* is generated as an ideal by the functions VVr - Pr : (z,w) ~ Wr - Pr(z),
~
X
Dk
--'>
<e,
for r = 1, ... , k. However, we do have B(VVr-Pr) =br-Pr(al, ... ,an)=O
(r=I, ... ,k),
and so indeed ker fJ* <;;; ker B. Thus the existence of 'ljJ is proved. For j = 1, ... , n, we have 7j;(ZJ) = (7j; 0 fJ*)(ZJ) = B(ZJ) = aJ • So, setting Sa = 7j;, case (ii) is established.
342
Introduction to Banach Spaces and Algebras
Case (iii) U is an arbitrary open, polynomially convex neighbourhood of Spa. Let K = Spa. Then R C U and, by Lemma 8.31(ii), there is a polynomial polyhedron P with K ~ ReP ~ U. Let R : O(U) --> O(P) be the restriction map, and let O(P) --> A be the homomorphism given by case (ii). Then let e a = e;: 0 R; the map e a has the required properties. 0
e;: :
9.2 General version. Our aim now is to remove the condition of polynomial convexity imposed on the neighbourhood U of Sp a in the last lemma; unfortunately, this process involves some arbitrary additional choices, and so the simple uniqueness statement of Lemma 9.1 is lost. The extension uses an ingenious idea of Arens and Calderon, which we give as a separate lemma. Lemma 9.2 (Arens-Calderon lemma). Let A be a commutative, umtal Banach algebra, let a = (a1, ... , an) E An, and let U be an open neighbourhood of Sp A a in en. Then there is a finite subset, say {a n+1, ... , am}, of A such that SPA a ~ SPA(a\, .,am)(a) cU. Further, there is a polynomial polyhedron P in em with SPA(a1, ... ,am) C P and 7rn ,m(P) ~ U.
with SPA a C ~. Take fl = (fl1, ... ,fln) E ~ \ U. Then fl tf- SPAa, and so there are elements b1 , .•• , bn E A such that
Proof There is an open polydisc
~
n
L br (flr1 -
ar )
=
l.
r=1
Set F(,L):= A(a1, ... ,an ,b 1, ... ,bn ). Then fl tf- SPF(/1)(a). Since this latter set is closed, there is an open neighbourhood V(fl) of IL with V(,L) n SPF(/1) (a) = 0. The set ~ \ U is compact, and so finitely many of the neighbourhoods V(fl) cover ~ \ U; for each such V(fl), there is a corresponding set {b 1 , ... , bn } of elements of A. Let {a n+1, ... , am} be the union of all these finite sets, and set F = A(a1, ... ,an ,an+1, ... ,am ). Then SpF(a) n (~ \ U) = 0. Certainly SPF(a) C ~ (for example, by using Theorem 8.29(ii)), and so it follows that SPF(a) C U. The final clause follows from Theorem 8.29(ii) and Lemma 8.31(ii). 0
Notation. Take 1 ~ n ~ m. For an open subset U of en, there is a continuous homomorphism 7r;:,m : O(U) --> O(U x e m - n ) defined by setting 7r;:,m(f)(Z1"", zm) = f(z1"", zn)
(J
E
O(U)).
We say that the mapping 7r;;',n is defined by 'ignoring coordinates' Zn+l,"" Zm· We can now give our statement of the holomorphic functional calculus in several variables for a fixed element of An.
343
Holomorphic functional calculus
Theorem 9.3 (Holomorphic functional calculus) Let A be a commutative, unital Banach algebra, let a = (al, ... , an) E An, and let U be an open neighbourhood of SPA a in C n . Then there is a continuous, unital homomorphism 8~ == 8 a : O(U) -+ A such that; (i)8 a (ZJ)=aJ (j=l, ... ,n)i (ii) cp(8 a (f)) = f(cp(at), ... , cp(a n )) (f E O(U), cp E
((ZI, ... , zm+n) E U x V).
Show that the function h is analytic on the neighbourhood U x V of Sp c, where c = (al, ... ,a m ,bl, ... ,bn ) E Am+n, and that 8 c = 8 a 8b. Exercise 9.2 Let K be a non-empty, compact subset of en such that K is rationally convex, in the sense of Exercise 8.12. Show that O(K) = R(K). Exercise 9.3 Formulate and prove a version of the composition theorem, Theorem 9.4, in which the second function of the composition is a function of m variables, rather than just one variable. Exercise 9.4 Let Kl
= {(z, w) E e 2: 1/2 ::; IzI2 + Iwl 2
L = B(O; 1) = {(z,w) E e 2
:
::;
I}, K
IzI2 + Iw12::;
= Kl U {(O, O)}, and
I}.
Set A = C(K), al = ZI K, and a2 = Z2 K, so that SPA(al,a2) = K. Take U to be an open neighbourhood of K. Construct two distinct continuous, unital homomorphisms from O(U) into A, both taking Zl and Z2 to al and a2, respectively. 1
1
345
Holomorphic functional calculus
Indeed, the first homomorphism is 8(al,a2)' For the second, use the following theorem: each function which is analytic on a neighbourhood of K 1 is the restriction of a function which is analytic on a neighbourhood of L (cf. Exercise 8.7).
Applications of the functional calculus 9.3 Shilov's idempotent theorem. As a first application of the holomorphic functional calculus, we shall prove the famous Shilov's idempotent theorem. In fact, the statement of this theorem makes no mention of analytic functions; they are simply used as a tool in its proof. In 1955, Shilov devised the functional calculus (in a weaker version than the one that we have given, and with a much more difficult proof) precisely to prove the idempotent theorem. It remains just possible that a quite different proof might be found, but this has evaded us for more than 50 years. As before, the characteristic function of a set K is denoted by XK. Theorem 9.5 (Shilov's idempotent theorem) Let A be a commutative, unital Banach algebra, and let K be subset of if? A which is both open and closed in if? A. Then there is a unique idempotent eK E A such that eK = XK· Proof We first deal with the question of uniqueness. Thus, let e and f be two idempotents in A with = [; we must show that e = f. Since A is commutative, we have
e
(e -
/)3
= e3
-
3e 2f
+ 3ef2 -
f3 = e - 3ef
+ 3ef -
f =e- f ,
and so (e - /)(1 - (e - /)2) = O. But, for every r.p E if? A, we have r.p(e) = r.p(f), and so r.p(1 - (e - /)2) = 1. By Corollary 4.47(i), 1 - (e - /)2 E G(A), and so e - f = O. Now we give the proof of the existence of e K. Clearly, if K = 0, then e K = 0, and, if K = if? A, then e K = 1. So, we may suppose that 0 =f=. K =f=. if? A; we write L = if? A \ K, and note that then if? A = K u L expresses the compact Hausdorff space if? A as a disjoint union of two non-empty, closed subsets. By the definition of the Gel'fand topology, it is elementary to find finitely many elements, say aI, ... ,an, of A such that a(K) and a(L) are disjoint, nonempty, compact subsets of en. Of course, SPAa = a(if?A) = a(K) u a(L). Now let V and W be disjoint, open neighbourhoods ofa(K) and a(L), respectively, in en, and set U = V u W. Then U is an open neighbourhood of Sp Aa. Take 8 a : O(U) -+ A to be the functional calculus homomorphism given by Theorem 9.3. Define h = Xv E O(U), so that h 2 = h in O(U), and define eK = 8 a (h), so that e'k = eK, and hence eK is an idempotent in A. Also, for each r.p E if? A, the spectral mapping property gives
r.p(eK) = r.p(8 a (h)) = h(r.p(al), ... , r.p(a n )) = Thus
eK
= XK, and so the result is proved.
{~
(r.p E K) , (r.pEL).
o
346
Introduction to Banach Spaces and Algebras
9.4 An implicit function theorem. The proof of Shilov's idempotent theorem amounted to solving the equation X 2 - X = 0 in a commutative Banach algebra A, with the solution having a specified Gel'fand transform. There is a considerable extension of this idea; this gives the result usually known as the implicit function theorem for Banach algebras. The statement of this theorem is as follows.
Theorem 9.6 Let A be a commutative, unital Banach algebra. Let 9 take a1, ... , an E A, and set L:
= {(a1('P), ... , an('P),g('P)) : 'P
E
C(cI>A),
E cI>A},
a compact subset of c n+1. Suppose that F is an analytic function on a neighbourhood of L: such that:
(i) F(Zl, ... , Zn+1) = 0 ((Zl, ... , zn+d E L:); (ii) 8F/8zn + 1 =1= 0 on L:. !hen there is a unique element bE A such that both F(a1, ... ,an ,b) = 0 and b= g.
0
The proof of the above result is rather complicated in detail, and we shall just give a special case, which still has interesting applications. Theorem 9.7 (Implicit function theorem) Let A be a commutative, unital Banach algebra. Let 9 E C (cI> A), and take a EA. Let f be an analytic function on an open neighbourhood of g( cI> A) in C, and suppose that:
(i) f(g('P)) = a('P) ('P E cI>A); (ii) f'(z) =1= 0 (z E g(cI>A)). Then there is a unique element b E A such that both f (b) = a and b = g. Proof Step 1: There is a jinzte subset {a1, . .. ,am} of A such that, whenever 'P1, 'P2 E cI> A have the property that ak ('Pd = ak ('P2) (k = 1, ... , m), then necessarily g( 'Pd = g( 'P2). Indeed, let 'Po E cI>A. Then f(g('Po)) = a('Po) and !'(g('Po)) =1= o. By the inverse function theorem, Proposition 1.36, there are neighbourhoods Vo and Wo of g( 'Po) and a( 'Po), respectively, in C such that f : Vo ---> Wo is bijective, with analytic inverse f- 1 : Wo ---> Vo. Now choose an open neighbourhood U = U('Po) of 'Po in cI> A such that both g( 'P) E Vo and a( 'P) E Wo when 'P E U. Since fog = a on cI> A, it follows that, for each 'P E U, the equation f (z) = a( 'P) has the solution z = g('P), and that this solution is the unique solution in Vo· Next, set W = U{U('Po) x U('Po) : 'Po E cI>A}. Then W is an open neighbourhood of the diagonal in cI> A xcI> A. If ('Pb 'P2) E W and ifa( 'Pd = a( 'P2), then, first, ('P1, 'P2) E U( 'Po) x U( 'Po) for some 'Po E cI> A, and
347
Holomorphic functional calculus
also g('Pd and g('P2) are both solutions of the equation J(z) = a('Pd = a('P2) with z in the corresponding Vo. It follows that g('Pd = g('P2). Suppose now that ('P~, 'Pg) E (cI> A X cI> A) \ W, so that 'P~ =1= 'Pg. Then there is an element C E A with 2('P~) =1= 2('Pg), and hence also 2('P1) =1= 2('P2) for all ('PI, 'P2) in a neighbourhood of ('P~ , 'Pg). Using the compactness of (cI> A X cI> A) \ W, we can thus find finitely many elements, say a2, ... ,am, of A such that, for each 'P1,'P2 E cI>A with the property that ak('Pd = ad'P2) (k = 2, ... ,m), we have ('P1,'P2) E W. We set al = a. We then see that g('Pt) = g('P2) whenever 'Pl,'P2 E cI>A and ak('Pt) = ak('P2) (k = 1, ... , m). This proves Step 1. Write a = (al,'" ,am) and a = (a1,'" ,am) : cI>A -+ em. Then, from Step 1, there is a unique function G : SPAa -+ e such that Goa = g. Elementary topology then shows that G is continuous. Step 2 : G extends to an analytic Junction on a neighbourhood oj Sp Aa.
Write ~ == SPA a. For each z = (Zl,"" zm) E ~, there is some 'P E cI> A with z = a('P)' Thus J(G(z)) = J(g('P)) = al('P) = Zl; also J'(G(z)) =1= O. By another use of the inverse function theorem, Proposition 1.36, each z E ~ has an open neighbourhood, say N(z), on which is defined an analytic function, say G z , which is unique subject to the conditions that: (a) Gz(z) = G(z);
(b) (f 0 Gz)(w) = WI (w E N(z)). By shrinking N(z), if necessary, we may then also suppose that: (c) Gz(w) = G(w) (w E N(z) n~). For simplicity, take N(z) to be an open ball N(z) = B(z; 30(z)). Define
N = U{B(z; o(z)) : z E ~}, so that N is an open neighbourhood of ~. If B(Zl; o(zt)) n B(Z2; O(Z2)) =1= 0, then, without loss of generality, we may suppose that o(zt) ::; O(Z2)' But then B(Zl;O(Zt)) ~ B(Z2;30(Z2))' so that G Z2 = G Z1 on the ball B(Zl;O(Zl)), and thus, in particular, on the set B(Zl;O(Zt)) nB(Z2;o(Z2))' We have shown that there is a well-defined function G E O(N) such that G I B(z; o(z)) = G z I B(z; o(z)) (z E ~). Then certainly G I ~ = G, and J(G(w)) = WI (W EN). This proves Step 2. Let 8 : O(N) -+ A be a continuous functional calculus homomorphism such that 8(ZJ) = a J (j = 1, ... ,m), so that, in particular, 8(Zt) = a1 = a. Define b = 8(G) EA. By the composition theorem, Theorem 9.4, we have f(b) = f(8(0)) = (f
Clearly,
b=
0
O)(a) = a1 = a.
g. Thus we have proved the existence of the required element b.
Introduction to Banach Spaces and Algebras
348
It remains to prove the uniqueness statement involving b. However, we know that the function J is analytic on a neighbourhood of g( A) = Sp b and that 1'(z) f. 0 (z E Spb)), and so this is immediate from Proposition 4.92. 0
Examples 9.8 Let A be a commutative, unital Banach algebra. (i) Let a E G(A), and suppose that (for some p ::::: 2) there exists 9 E C( A) with gP = Ii on A. Then there is a unique element b E A such that bP = a and b = g. [Take J(z) = zP in Theorem 9.7: clearly Jog = Ii and, since a is invertible, 1'(z) f. 0 (z E Spa).] (ii) Let a E G(A), and suppose that Ii has a continuous logarithm on A, in the sense that Ii E exp C( A)' Then there is a unique element b E A such that exp b = a and b = g. [This is just like (i), but with J(z) = eZ .] (iii) We remark that by taking a = 0 and 9 = XK (where K is an open-andclosed subset of A) and J(z) = z2 - z for z E C, we obtain an alternative proof of Shilov's idempotent theorem. 0
9.5 The Arens-Royden theorem. Another viewpoint on Shilov's idempotent theorem is to see it as the start of a programme to describe the topology of the character space A explicitly in terms of algebraic features of A: it sets up a bijection between the set of idempotent elements of A and the set of openand-closed subsets of A. The Arens-Royden theorem takes this a stage further: it says that Hl(A;Z) ~ G(A)/expA. Here, as usual, G(A) is the group of units of A and expA = {e a : a E A}. The notation H 1 ( A; Z) means the first integral cohomology group of A: if you know what that is, fine, but, if not, do not worry. Set C = C( A) (so that c = A)' Then what will be proved (in a slightly more explicit form) is that
G(A)/ exp A
~
G(C)/ exp C.
The fact that G(C)/ exp C, which is manifestly a topological invariant of A, can also be identified with HI ( A; Z) is then a matter of pure topology, and must be outside the scope of this book. Let A be a commutative, unital Banach algebra, with Gel'fand transform g. Then it is clear that 9(G(A)) <:;; G(C) and that g(expA) <:;; expC. Thus 9 certainly induces a group homomorphism G(A)/expA --> G(C)/expC.
g:
Theorem 9.9 (Arens-Royden) Let A be a commutative, unital Banach algebra. Then the homomorphism
g: G(A)/ exp A --> G(C)/ exp C, induced by the Gel 'fand representation oj A, is an isomorphism.
349
Holomorphic functional calculus
Proof The proof falls naturally into two parts. First, we note that 9 is mjective: i.e. if a E G(A) and if E expC, then a E expA. But this is an immediate consequence of Example 9.8(ii), above. It thus remains to prove that 9 is surjective: i.e. given f E G(C), there is some a E G(A) with f /a E expC. Thus, let f E G(C), so that f is a continuous function on the compact Hausdorff space A with Z(f) = 0. Set
a
c
= inf{lf(
E A}.
Then c > 0, and, for each g E C with Ig - fl
n
g
= Lakak+n. k=1
Define a: A
~
-t
== SPA(al, ... ,a2n) by
a(
c 2n
= (al(
E
A),
by n
G(ZI, ... ,Z2n)
=
LZkZk+n.
(*)
k=1
Then G E coo(C2n) and G(z) -:/= 0 on ~, so that it is also the case that G(z) -:/= 0 on some open neighbourhood U of ~ in C 2n. By the Arens-Calderon lemma, Lemma 9.2, there is a finite subset, say {a2n+l, . .. ,am}, of A and a polynomial polyhedron P such that
SPB(al, ... ,am) C P ~ U
X
cm- 2n ,
where B = A( aI, ... ,am). By a slight 'abuse of notation', we now regard G as being defined on P by the equation (*), above. Then G E COO(P) and G(z) -:/= 0 for all Z E P. Also, we have Go(3 = g, where (3: A - t SPA(al, ... ,am) is defined by (3('1') = (al(
E
-1
£ (P) by
w(z)
=
G(z)-1(8G)(z)
(z E P).
aw
A simple calculation shows that = O. By Corollary 8.35, 'H 1 (P) = {O}, and so there is some 'lj; E COO(P) such that w = a'lj;. Set h = exp( -'lj;)G E COO(P), so that Gh- 1 = e'I/J on P and 8h = exp(-'lj;) (8G - G8'lj;) = 0,
which shows that h
E
O(P).
350
Introduction to Banach Spaces and Algebras
Now, applying a functional calculus map O(P) a
----+
A, define
= h(al, ... ,am)'
Then, by the spectral mapping property (in fact only Lemma 9.1 is needed),
a = h 0 (3, so that
9 . a-I = (G 0 (3)(h 0 (3)-1 = e,po{3 E exp C.
Thus fa- 1 = (f g-1 ) (ga- 1) E exp C, and the proof is complete.
o
9.6 The local maximum modulus theorem. The final application of the functional calculus in several variables that we shall offer is Rossi's local maximum modulus theorem. Let A be a natural Banach function algebra on a compact space K. In Section 4.13, we defined the notion of a peak set for A. We now say that a subset E of K is a local peak set for A if there exists f E A and a neighbourhood U of E such that f(x)
= 1 (x
E E)
If(x)1 < 1
and
(x E U \ E);
the function f peaks locally on E. A point Xo E K such that {xo} is a local peak set is a local peak point. We make a preliminary topological remark. Let mEN, and let U1 and U2 be non-empty, open sets in em. Suppose that L is a compact subset of U U V. Then we claim that there exist compact sets Ll and L2 such that L1 CUI, L2 C U2 , and L = Ll U L 2 . Indeed, for j = 1,2, take compact subsets K),n of U) for n E N such that K),n C int K),n+l (n E N) and U{ K),n : n E N} = U). Then there exists no EN such that L C intK 1,no UintK2 ,no' Set L) = LnK),nn for J = 1,2. Theorem 9.10 (Rossi's local peak set theorem) Let A be a natural Banach function algebra on a compact space K, and let E be a local peak set for A. Then E is a peak set for A. Proof Choose f E A and a neighbourhood U of E such that f(x) = 1 (x E E) and If(x)1 < 1 (x E U \ E). Set h = f - I, so that Z(h) n U = E and Reh(x) < 0 (x E U \ E), and hence Reh(x) :::; 0 (x E U). Since E is compact, we may suppose that there are finitely many basic open sets U1 , ... , Ur of the form Uk
= {x
E
K : If)(x)1 < 1 (j
where 1 = no < ... < nr and that
= nk-l + 1, ... , nk)} (k = 1, ... , r),
12, ... , fnr
{x E K: If)(x)1 < 2 (j = nk-l
E
A, such that E <;;; U~=1 Uk, and such
+ l, ... ,nk)} <;;; U
(k = l, ... ,r).
351
Holomorphic functional calculus
Set n = n r , and then define Vk ={ZEe n :lz]I<1 (j=nk-l+1, ... ,nk)}
(k=1, ... ,r),
an open set in en. Next, set V = U~=l Vk and
W
{z E en: Rezl < O} U (en \ V),
=
so that V and Ware open sets in en. Clearly, V n W ~ {z E en : Rezl < O}. We recall that SPA (fl,'''' In) = {f(x) : X E K}, where f(x)
= (Il(x), ... ,In(x))
(x E K).
We claim that f(E) ~ V \ Wand f(K \ E) ~ W, and hence that V U W is an open neighbourhood of SPA (fl,"" In). Indeed, first, suppose that x E K \ U. Then, for each k = 1, ... , r, there exists j E {nk-l + 1, ... , nd such that II](x)1 ~ 2, and so f(x) tf- V k. Hence f(x) E en \ V ~ W. Next, suppose that x E U \ E. Then, by the choice of I, we have f(x) E {z E en: Rezl < O} ~ W. Thus I(K\E) ~ W. Finally, suppose that x E E. Then x E Uk for some k E {1, ... , r}, and then f(x) E Vk ~ V, and certainly f(x) 1- W. This establishes our claim. By the Arens-Calderon lemma, Lemma 9.2, there exist In+l,.'" 1m E A and a polynomial polyhedron P ~ em such that L:= SPA(fl, ... ,1m) c P and 7rn ,m(P) ~ V U W. Set P v = 7r~,~(V) and Pw = 7r~,~(W), so that P v and P w are open sets in em with P v U Pw ~ P::::> Land
Pv n Pw ~ {w E
em: Rewl < O}.
By the preliminary remark, there are compact sets Ll and L2 with Ll C Pw, with L2 c Pv , and with Ll U L2 = L. Set Lo = {(h(x), ... , Im(x)) : x E E}. Since f(E) C V, we see that Lo is a compact subset of L n P v , and so we may suppose that Lo ~ L 2. We can define a branch of the analytic function log on {WI E e : Re WI < O}, and then W ~ (log wI) jWl is a well-defined analytic function on Pv n Pw. By Corollary 8.36, there exist G E O(Pv ) and H E O(Pw) such that
(G - H)( w)
log WI
= - --
WI
(w
E
Pv n Pw ) .
Define F(w)
=
{WI exp(wlG(w)) exp(wlH(w))
(WEPv ), (WEPw).
Then F is well defined and analytic on P v U P w . Set g = F(h, ... , 1m) E A. If x E E, then f(x) E V and h(x) = 0, and so g(x) = 0. If x E K \ E, then f(x) E Wand so g(x) -10. Hence Z(g) = E.
352
Introduction to Banach Spaces and Algebras
The function 1/ F is bounded on compact subsets of Pw , and so there exists Ml > 0 such that Re(l/F(w)) ::; Ml (w ELI)' Define
0/'( )
=
exp( -WI G (W)) -1 = ~ wiGJ+l(W)
0/ W
6 J=o
WI
(
J
(
1)1
+ .
wE
P)
v ,
so that 1jJ E O( P v ). Thus there exists M2 > 0 such that Re 1jJ( w) ::; M2 (w E L2)' Suppose that W E P v \ Lo and that (WI, ... , Wn ) = f(x) for some x E K. Then x
F(w)
1
=
WI
+ 1jJ(w)
(w E Pv \ Lo),
we have Re(l/F(w))::; M2 (w E L2 \Lo). Set M = max{M1 ,M2 } + 1. Then Re(l/F(w)) < M (w E L \ L o), and so Re(1/g(x)) < M (x E K \ E). Suppose that ( E e with I( - EI < E. Then ( = E(l + re iO ) for some r < 1 and E [0,27rJ, and so
e
Re(l/()
1
= -. E
1 1 + 2r cos
1
e + r2
>-. 4E
This implies that the set g(K) omits the open disc {( E e : IE - (I < E} in e, where E = 1/4M, and hence that the function E(cl- g)-1 E A peaks exactly on the set E. Thus E is a peak set, as required. 0 Theorem 9.11 (Local maximum modulus theorem) Let A be a natural Banach
function algebra on a non-empty, compact Hausdorff space K, and let U be a non-empty, open set in K. Then Iflu
=
Ifl(r(A)nu)u8u
(f
E
A).
Proof Let E = {x E U: If(x)1 = Iflu}' We must show that either EnaU -=f. 0 or that En r(A) n U -=f. 0. Suppose that E n aU = 0. We may suppose that there exists x E U with f(x) = Iflu = 1. Set F = {y E U : f(y) = I} and 9 = (1 + 1)/2. Then F -=f. 0, g(y) = 1 (y E F), and Ig(y)1 < 1 (y E U \ F). By Theorem 9.10, F is a peak set for A, and so F n r( A) -=f. 0. Since F ~ E c U, the result follows. 0 Recall that it is a conjecture of Gel'fand that the only natural uniform algebra on IT is C(IT). Corollary 9.12 Let A be a natural Banach function algebra on the closed unit interval IT. Then r(A) = IT.
353
Holomorphic functional calculus
Proof Assume towards a contradiction that f(A) C;; lI. Then there exist a, bE II with a < b such that (a, b) n f(A) = 0. Clearly, there exists a function f E A with f(a) = f(b) = 0, but with If(c)1 = Ifl[a,b] > for some c E (a, b). This contradicts Theorem 9.11. D
°
Notes For the applications that we have given, and other applications, of the functional calculus, see [4], [31, Section 21], [32, Chapitre I, Section 4], [47, Section 2.4], [79, Chapter III, Section 6], [96, Chapter III], and [126, Section 3.5]. In particular, Theorem 9.6 is proved in [79, Chapter III, Theorem 6.1] and [126, Theorem 3.5.12], and the Arens-Royden theorem, Theorem 9.9 is proved in [4, Chapter 15] and [79, Chapter III, Theorem 7.2]. Let A be a commutative, unital Banach algebra, and let C = C( A). For n E N, set An = Mn(A) and Cn = Mn(C). Then an extension ofthe Arens-Royden theorem states that G(An)/Go(An) ~ G(Cn)/GO(Cn ). Here Go(An) and Go(Cn ) are the principal components of the identity in G(An) and G(Cn ), respectively. These results explore the relationship between properties of a Banach algebra and of its character space; for further results along these lines, see [132, 157]. The original proof of Theorem 9.10 is due to Rossi; a simplified version is in [79, Chapter III, Theorem 8.1] and [86, Chapter I, Theorem 19]. A more general form is [152, Theorem 9.3]. Further extensions are given in [9]. Exercise 9.5 Let A be a natural Banach function algebra on A. (i) In the case where A has an identity, A is compact and non-empty. Deduce the converse from Shilov's idempotent theorem. (ii) Suppose that A is totally disconnected. Prove that A is dense in C( A). (iii) Let K be a compact, totally disconnected subset of en, where n E N. Prove that O(K) = C(K). Is it true that necessarily R(K) = C(K)? (iv) Let K be a compact, polynomially convex, totally disconnected subset of en. Prove that P(K) = C(K). (A compact, totally disconnected subset of en is not necessarily polynomially convex; see [79, Theorem 2.5].) Exercise 9.6 Let A be a commutative, unital Banach algebra, and let ao, .. . ,an E A. Suppose that g E C(A) satisfies the equation 'L.;=o aJgJ = 0 in C(A), whilst the function 'L.;=1 jaJg J- 1 does not vanish on A. Prove that there is a unique element
b E A with
'L.;=o a b1 J
= 0 in A and with
b=
g.
Exercise 9.7 Let A be a natural Banach function algebra on a non-empty, compact Hausdorff space K, and let f, g E A. (i) Set M(f,g) = {x E K: If(x)1 :0:: Ig(x)I}. Let E be a component of M(f,g) such that En Z(g) = 0. Prove that M(f,g) n r(A) 1= 0. (ii) Let g E A. Suppose that there is a set E c:::; K such that g(x) = 1 (x E E) and Ig(x)1 > 1 (x E K \ E). Prove that E is a peak set for A. This is a local minimum modulus theorem. This exercise is taken from [9].
References [1] Ahlfors, L. V. (1966). Complex Analysis. An Introduction to the Theory of Analytic Functions of One Complex Variable. McGraw-Hill Kogakusha, Tokyo. [2] Albiac, F. and Kalton, N. (2000). Topics in Banach Space Theory. Graduate Texts in Mathematics, Volume 233, Springer, New York. [3] Albrecht, E. and Dales, H. G. (1983). Continuity of homomorphisms from C*algebras and other Banach algebras, in Radical Banach Algebras and Automatic Continuity, Proceedings, Long Beach, 1981. Lecture Notes in Mathematics, 975, Springer-Verlag, Berlin, 375~96. [4] Alexander, H. and Wermer, J. (1998). Several Complex Variables and Banach Algebras, 3rd edition. Graduate Texts in Mathematics, Volume 35, SpringerVerlag, New York. [5] Allan, G. R. (1967). On a class of locally convex algebras. Proc. London Mathematical Society (3), 17, 91~114. [6] Allan, G. R. (1969). A note on the holomorphic functional calculus in a Banach algebra. Proc. American Mathematical Society, 22, 77~81. [7] Allan, G. R. (1969). An extension of the Silov~Arens~Calderon theorem. J. London Mathematical Society, 44, 595~601. [8] Allan, G. R. (1970). On lifting analytic relations in commutative Banach algebras. J. Functional Analysis, 5, 37~43. [9] Allan, G. R. (1971). An extension of Rossi's local maximum modulus principle. J. London Mathematical Society (2), 3, 39~43. [10] Allan, G. R. (1972). Embedding the algebra of formal power series in a Banach algebra. Proc. London Mathematical Society (3), 25, 329~40. [11] Allan, G. R. (1998). Stable inverse-limit sequences, with application to Frechet algebras. Studia Mathematica, 121, 277~308. [12] Allan, G. R. (1998). Stable elements of Banach and Fn3chet algebras. Studia Mathematica, 129, 67~96. [13] Allan, G. R. (2000). Stable inverse-limit sequences and automatic continuity. Studia Mathematica, 141, 99~107. [14] Ambrozie, C. and Miiller, V. (2004). Invariant subspaces for polynomially bounded operators. J. Functional Analysis, 213, 321~45. [15] Arens, R. and Calderon, A. P. (1955). Analytic functions of several Banach algebra elements. Annals of Mathematics, 62, 204~16. [16] Argyros, S. A. and Haydon, R. (2010). A hereditarily indecomposable .coo-space that solves the scalar-plus-compact problem, Acta Mathematica, to appear.
[17] Arveson, W. (1982). The harmonic analysis of automorphism groups. Proc. Symposium Pure Mathematics, American Mathematical Society, 38, 199~269. [18] Aupetit, B. (1979). Proprietes Spectrales des Algebres de Banach. Lecture Notes in Mathematics, 735, Springer-Verlag, Berlin. [19] Aupetit, B. (1982). The uniqueness of complete norm topology in Banach algebras and Banach Jordan algebras. J. Functional Analysis, 47, 1~6. [20] Aupetit, B. (1991). A Primer on Spectral Theory. Universitext, Springer-Verlag, New York. [21] Bade, W. C. and Curtis, P. C. Jr., (1960). Homomorphisms of commutative Banach algebras. American J. Mathematics, 82, 589~608. [22] Bade, W. C. and Dales, H. C. (1989). Discontinuous derivations from Banach algebras of power series. Proc. London Mathematical Society (3), 59, 133~52. [23] Bade, W. C. and Dales, H. C. (1989). Continuity of derivations from radical convolution algebras. Studia Mathematica, 95, 59~91. [24] Baker, J. A. (1999). The Dirichlet problem for ellipsoids. American Mathematical Monthly, 106, 829~34. [25] Banach, S. (1932). Theorie des Operations Lineaires. Monografje Matematyczne, Volume 1, Polish Scientific Publishers, Warsaw. (Reprinted Chelsea, New York 1955.) [26] Basener, R. F. (1973). On rationally convex hulls. Trans. American Mathematical Society, 182, 353~81. [27] Beardon, A. F. (1979). Complex Analysis. The Argument Principle in Analysis and Topology. John Wiley, Chichester. [28] Berndtsson, B. and Ransford, T. J. (1986). Analytic multi functions, the 8equation and a proof of the corona theorem. Pacific J. Mathematics, 124, 57~72. [29] Beurling, A. (1938). Sur les integrals de Fourier absolument convergentes. Congres des Math. Scand., Helsingfors, 345~66. [30] Bollobas, B. (1992). Linear Analysis. An Introductory Course. Cambridge Mathematical Textbooks, Cambridge University Press. [31] Bonsall, F. F. and Duncan, J. (1973). Complete Normed Algebras. SpringerVerlag, New York. [32] Bourbaki, N. (1960). Elements de mathematiques: Theories spectrales. Hermann, Paris. [33] Bredon, C. E. (1993). 'Japology and geometry. Springer-Verlag, New York. [34] Bratteli, O. and Robinson, D. W. (1979). Operator Algebras and Quantum Statistical Mechanics, I. Springer-Verlag, New York. [35] Brosowski, B. and Deutsch, F. (1981). An elementary proofofthe Stone-Weierstrass theorem. Proc. American Mathematical Society, 81, 89~92. [36] Brown, S. Chevreau, B. and Percy, C. (1988). On the structure of contraction operators, II. J. Functional Analysis, 76, 30~55.
356
References
[37] Carleson, L. (1962). Interpolations by bounded analytic functions and the corona problem. Annals of Mathematics, 16, 547-59. [38] Carleson, L. (1966). On convergence and growth of partial sums of Fourier series. Acta Mathematica, 116, 135-57. [39] Chalendar, 1. and Esterle, J. R. (2000). Le probleme du sous-espace invariant, in Development of Mathematics, 1950-2000, ed. J.-P. Pier, Birkhiiuser, Basel. [40] Cohn, D. L. (1980). Measure Theory. Birkhiiuser, Boston. [41] Cole, B. (1968). One point parts and the peak point conjecture, Thesis, Yale University. [42] Conway, J. B. (1990). A Course in Functional Analysis, 2nd edition. SpringerVerlag, New York. [43] Dales, H. G. (1971). Boundaries and peak points for Banach function algebras. Proc. London Mathematical Society (3), 22, 121-36. [44] Dales, H. G. (1973). The uniqueness of the functional calculus. Proc. London Mathematical Society (3), 21, 638-48. [45] Dales, H. G. (1973). Exponentiation in Banach star algebras. Proc. Edinburgh Mathematical Society (Series II), 20, 163-5. [46] Dales, H. G. (1979). A discontinuous homomorphism from C(X). American J. Mathematics, 101, 647-734. [47] Dales, H. G. (2000). Banach Algebras and Automatic Continuity. London Mathematical Society Monographs, New Series, Volume 24, Clarendon Press, Oxford. [48] Dales, H. G. (2000). Banach algebras, in Introduction to Banach Algebras, Operators, and Harmonic Analysis. London Mathematical Society Student Texts, Volume 57, Cambridge University Press. [49] Dales, H. G. and Loy, R. J. (1986). Uniqueness of the norm topology for Banach algebras with finite-dimensional radical. Proc. London Mathematical Society (3), 14, 633-6l. [50] Dales, H. G. and McClure, J. P. (1973). Continuity of homomorphisms into certain commutative Banach algebras. Proc. London Mathematical Society (3), 26,69-8l. [51] Dales, H. G. and Woodin, W. H. (1987). An Introduction to Independence for Analysts. London Mathematical Society Lecture Note Series, Volume 115, Cambridge University Press. [52] Dales, H. G., Lau, A. T.-M., and Strauss, D. (2010). Banach algebras on semigroups and on their compactifications, Memoirs American Mathematical Society, Volume 205, 1-165. [53] Dales, H. G., Lau, A. T.-M., and Strauss, D. (2011). Second duals of measure algebras, Dissertationes Mathematicae (Rozprawy Matematyczne) , to appear. [54] Dales, H. G., Loy, R. J., and Willis, G. A. (1994). Homomorphisms and derivations from 8(E). J. Functional Analysis, 120, 201-19. [55] Dales, H. G., Patel, S. R., and Read, C. J. (2010). Fn§Chet algebras of power series, Banach Center Publications, Polish Academy Sci., Warsaw, to appear.
References
357
[56] Davidson, K. R. (1996). C*-algebras by Example. Fields Institute Monographs, Volume 6, American Mathematical Society, Providence, Rhode Island. [57] Diestel, J. (1984). Sequences and Series in Banach spaces. Graduate Texts in Mathematics, Volume 92, Springer-Verlag, New York. [58] Diestel, J. and Uhl, J. J., Jr. (1977). Vector Measures. American Mathematical Society Mathematical Surveys, Volume 15, Providence, Rhode Island. [59] Dixon, P. G. (1970). Generalized B*-algebras. Proc. London Mathematical Society (3), 21, 693-715. [60] Dixon, P. G. (1977). A Jacobson semi-simple Banach algebra with a dense nil subalgebra. Colloquium Mathematicum, 37, 81-2. [61] Dixon, P. G. and Esterle, J. R. (1986). Michael's problem and the Poincare-Bieberbach phenomenon. Bull. American Mathematical Society, 15, 127-87. [62] Douglas, R. G. (1972). Banach Algebra Techniques in Operator Theory. Academic Press, New York. [63] Dunford, N. and Schwartz, J. T. (1958). Linear Operators. Part I. General Theory. Interscience, New York and London. [64] Dunford, N. and Schwartz, J. T. (1963). Linear Operators. Part II. Spectral Theory. Interscience, New York and London. [65] Dunford, N. and Schwartz, J. T. (1971). Linear Operators. Part III. Spectral Operators. Wiley-Interscience, New York and London. [66] Edwards, R. E. (1967). Fourier Series, A Modern Introduction, Volume I. Holt, Rinehart and Winston, New York. [67] Eidelheit, M. (1940). On homomorphisms of rings of linear operators. Studia Mathematica, 9, 97-105. [68] Enflo, P. (1987). On the invariant subspace problem in Banach spaces. Acta Mathematica, 158, 213-313. [69] Engelking, R. (1977). General Topology. Monografje Matematyczne, Volume 60, Polish Scientific Publishers, Warsaw. [70] Ermert, O. (1996). Continuity of homomorphisms from AF- C* -algebras and other inductive limit C* -algebras. J. London Mathematical Society (2), 54, 369-86. [71] Eschmeier, J. (2000). Invariant subs paces , in Introduction to Banach Algebras, Operators, and Harmonic Analysis. London Mathematical Society Student Texts, Volume 57, Cambridge University Press. [72] Eschmeier, J. and Putinar, M. (1996). Spectral Decompositions and Analytic Sheaves. London Mathematical Society Monographs, New Series, Volume 10, Clarendon Press, Oxford. [73] Esterle, J. R. (1978). Sur I'existence d'un homomorphisme discontinue de C(K). Proc. London Mathematical Society (3), 36, 46-58. [74] Esterle, J. R. (1980). Theorems of Gelfand-Mazur type and continuity of epimorphisms from C(K). J. Functional Analysis, 36, 273-86.
~58
References
[75] Farin, G. F. (1988). Curves and Surfaces for Computer Aided Geometric Design. A Practical Guide. Academic Press, New York. [76] Feinstein, J. F. (1992). A non-trivial, strongly regular uniform algebra, J. London Mathematical Society (2), 45, 288-300. [77] Feldman, C. (1951). The Wedderburn principal theorem in Banach algebras, Proc. American Mathematical Society, 2, 771-7. [78] Fragoulopoulou, M. (2005). Topological Algebras with Involution. North-Holland Mathematical Studies, Volume 200, Elsevier, Amsterdam. [79] Gamelin, T. W. (1969). Uniform Algebras. Prentice Hall, Englewood Cliffs, New Jersey. [80] Garnett, J. B. (1981). Bounded Analytic Functions. Academic Press, New York. [81] Gel'fand, I. M. (1941). Normierte Ringe, Rec. Math. N. S. (Matern. Sbornik) , 9,3-24. [82] Gel'fand, I. M., Raikov, D., and Shilov, G. (1964). Commutative Normed Rings. Chelsea, New York. (Translated from the Russian edition of 1960.) [83] Gillman, L. and Jerison, M. (1960). Rings of Continuous Functions. van Nostrand, Princeton, New Jersey. [84] Grafakos, L. (2004). Classical and Modern Fourier Analysis. Pearson Education, Upper Saddle Rivers, New Jersey. [85] Grauert, H. and Fritzsche, K. (1976). Several Complex Variables. Graduate Texts in Mathematics, Volume 38, Springer-Verlag, New York. [86] Gunning, R. C. and Rossi, H. (1965). Analytic Functions of Several Complex Variables. Prentice Hall, Englewood Cliffs, New Jersey. [87] Halmos, P. R. (1960). Naive Set Theory. D. van Nostrand, Princeton, New Jersey. [88] Halmos, P. R. (1957). Introduction to Hilbert Space and the Theory of Spectral Multiplicity. Chelsea Publishing Company, New York. [89] Halmos, P. R. (1967). A Hilbert Space Problem Book. van Nostrand-Reinhold, New York. [90] Harris, L. A. and Kadison, R. V. (1996). Schurian algebras and spectral additivity. J. of Algebra, 180, 175-86. [91] Harvey, R. and Wells, R. 0., Jr. (1969). Compact holomorphically convex subsets of a Stein manifold. Trans. American Mathematical Society, 136, 509-16. [92] Helemskii, A. Ja. (1993). Banach and Locally Convex Algebras. Clarendon Press, Oxford. [93] Helemskii, A. Ja. (2006). Lectures and Exercises on Functional Analysis. American Mathematical Society, Providence, Rhode Island. [94] Hewitt, E. and Ross, K. A. (1979). Abstract Harmonic Analysis, Volume I: Structure of Topological Groups, Integration Theory, Group Representations, 2nd edition. Springer-Verlag, Berlin.
References
359
[95] Hindman, N. and Strauss, D. (1998). Algebra in the Stone-6ech compactification, Theory and Applications. Walter de Gruyter, Berlin. [96] Hormander, L. (1973). An Introduction to Complex Analysis in Several Variables. North-Holland Mathematical Library, Volume 7, North-Holland, Amsterdam. [97] Hungerford, T. W. (1980). Algebra. Springer-Verlag, Berlin. [98] Johnson, B. E. (1967). The uniqueness of the (complete) norm topology. Bull. American Mathematical Society, 73, 537-9. [99] Johnson, B. E. (1967). Continuity of homomorphisms of algebras of operators. J. London Mathematical Society, 42, 537-41. [100] Johnson, B. E. (1972). Cohomology in Banach algebras. Memoirs American Mathematical Society, 127, 1-96. [101] Johnson, B. E. and Sinclair, A. M. (1968). Continuity of derivations and a problem of Kaplansky. American J. Mathematics, 90, 1067-73. [102] Kadison, R. V. and Ringrose, J. R. (1983). Fundamentals of the Theory of Operator Algebras, Volume I, Elementary Theory. Academic Press, New York and London. [103] Kadison, R. V. and Ringrose, J. R. (1986). Fundamentals of the Theory of Operator Algebras, Volume II, Advanced theory. Academic Press, Orlando, Florida. [104] Kahane, J.-P. and Zelazko, W. (1968). A characterization of maximal ideals in commutative Banach algebras. Studia Mathematica, 29, 339-43. [105] Kantrowitz, R. and Neumann, M. M. (2008). Yet another proof of Minkowski's inequality. American Mathematical Monthly, 115, 445-7. [106] Katznelson, Y. (1976). An Introduction to Harmonic Analysis. Dover Publications, New York, [107] Kelley, J. L. (1955). General Topology. D. van Nostrand, Princeton, New Jersey. [108] Kelley, J. L. and Vaught, R. L. (1953). The positive cone in Banach algebras. Trans. American Mathematical Society, 74, 44-55. [109] Korner, T. (1988). Fourier Analysis. Cambridge University Press. [110] Kingman, J. F. C. and Taylor, S. J. (1966). Introduction to Measure and Probability. Cambridge University Press. [111] Krantz, S. G. (1982). Function Theory of Several Complex Variables. John Wiley, New York. [112] Lacey, M. and Thiele, C. (2000). A proof of bounded ness of the Carleson operator. Mathematical Research Letters, 7,361-70. [113] Lang, S. (1993). Real and Functional Analysis. Graduate Texts in Mathematics, Volume 142, Springer-Verlag, New York. [114] Laursen, K. B. and Neumann, M. M. (2000). An Introduction to Local Spectral Theory. London Mathematical Society Monographs, New Series, Volume 22, Clarendon Press, Oxford.
360
References
[115] Lindenstrauss, J. and Tzafriri, L. (1977). Classical Banadl Spaces I: Sequence Spaces. Springer-Verlag, Berlin. [116] Lomonosov, V. I. (1973). Invariant subspaces for operators commuting with compact operators. Functional Analysis and Applications, 7, 213-4. [117] Loomis, L. H. (1953). An Introduction to Abstract Harmonic Analysis. D. van Nostrand, Princeton, New Jersey. [118] Losert, V. (2008). The derivation problem for group algebras. Annals of Mathematics (2), 168, 221-46. [119] Mac Lane, S. (1963). Homology. Springer-Verlag, Berlin. [120] Mac Lane, S. (1971). Categories for the Working Mathematician. Graduate Texts in Mathematics, Volume 5, Springer-Verlag, New York. [121] Megginson, R. E. (1998). An Introduction to Banach Space Theory. SpringerVerlag, New York. [122] Michael, E. A. (1952). Locally multiplicatively-convex topological algebras. Memoirs American ll,fathematical Society, 11, 1-82. [123] Muller, V. (1976). On discontinuity of spectral radius in Banach algebras. Comment. Math Univ. Carolinae, 17, 591-8. [124] Munkres, J. R. (1975). Topology, a First Course. Prentice Hall, Englewood Cliffs, New Jersey. [125] Murphy, G. J. (1990). C* -algebras and Operator Theory. Academic Press, San Diego, California. [126] Palmer, T. W. (1994). Banach Algebras and the General Theory of*-Algebras, Volume I: Algebras and Banach Algebras. Encyclopedia of Mathematics and its Applications, Volume 49, Cambridge University Press. [127] Palmer, T. W. (2001). Banach Algebras and the General Theory of*-Algebras, Volume II: *-Algebras. Encyclopedia of Mathematics and its Applications, Volume 79, Cambridge University Press. [128] Pedersen, G. K. (1989) Analysis Now. Graduate Texts in Mathematics, Volume 118, Springer-Verlag, New York. [129] Pham, H. L. (2006). Automatic continuity for Banach algebras with finitedimensional radical. J. Australian Mathematical Society, 81, 279-95. [130] Pietsch, A. (2007), History of Banadl Spaces and Linear Operators. Birkhauser, Boston. [131] Radjavi, H. and Rosenthal, P. (2003). Invariant Subspaces. 2nd edition. Dover Publications, Mineola, New York. [132] Raeburn, I. (1977). The relationship between a commutative Banach algebra and its maximal ideal space. J. Functional Analysis, 25, 366-90. [133] Range, R. M. (1986). Holomorphic Functions and Integral Representations in Several Complex Variables. Graduate Texts in Mathematics, Volume 108, Springer-Verlag, New York. [134J Ransford, T. J. (1984). A short elementary proof of the Bishop-Stone-Weierstrass theorem. Math. Proc. Cambridge Philosophical Society, 96, 309-11.
References
361
[135] Ransford, T. J. (1984). A short proof of Johnson's uniqueness-of-norm theorem. Bull. London Mathematical Society, 21, 487-8. [136] Ransford, T. J. (1995). Potential Theory in the Complex Plane. London Mathematical Society Student Texts, Volume 28, Cambridge University Press. [137] Ransford, T. J. (2001). Almost-everywhere discontinuity of the spectral radius. Proc. American Mathematical Society, 129,749-51. [138] Read, C. J. (1984). A solution to the invariant subspace problem. Bull. London Mathematical Society, 16, 337-401. [139] Read, C. J. (1988). The invariant subspace problem for a class of Banach spaces. II. Hypercyclic operators. Israel J. Mathematics, 63, 1-40. [140] Rickart, C. E. (1960). General Theory of Banach Algebras. D. van Nostrand, Princeton, New Jersey. [141] Ringrose, J. R. (1972). Automatic continuity of derivations of operator algebras. J. London Mathematical Society (2), 5, 432-8. [142] Roitman, Y. and Sternfeld, Y. (1981). When is a linear functional multiplicative? Trans. American Mathematical Society, 267, 111-24. [143] Rudin, W. (1966). Real and Complex Analysis. McGraw-Hill, New York. [144] Rudin, W. (1973). Functional Analysis. McGraw-Hill, New York. [145] Rudin, W. (1990). Fourier Analysis on Groups. John Wiley, New York. [146] Runde, V. (2005). A Taste of Topology. Springer, New York. [147] Runde, V. (2010). B(lP) is never amenable. J. American Mathematical Society, 23, 1175-85. [148] Sinclair, A. M. (1969). Continuous derivations on Banach algebras. Proc. American Mathematical Society, 20, 166-70. [149] Sinclair, A. M. and Smith, R. R. (1995). Hochschild Cohomology of Von Neumann Algebras. London Mathematical Society Lecture Notes, Volume 203, Cambridge University Press. [150] Singer, I. M. and Wermer, J. (1955). Derivations on commutative normed algebras. Mathematische Annalen, 129, 260-4. [151] Stolzenberg, G. (1963). Polynomially and rationally convex sets. Acta Mathematica, 109, 259-89. [152] Stout, E. L. (1971). The Theory of Uniform Algebras. Bogden and Quigley, Tarrytown-on-Hudson, New York. [153] Stout, E. L. (2007). Polynomial Convexity. Progress in Mathematics, Volume 261, Birkhiiuser, Boston, Massachusetts. [154] Takesaki, M. (1979). Theory of Operator Algebras, I. Springer-Verlag, New York. [155] Taylor, J. L. (1970). A joint spectrum for several commuting operators. J. Functional Analysis, 6, 172-91. [156] Taylor, J. L. (1970). The analytic functional calculus for several commuting operators. Acta Mathematica, 125, 1-48.
362
References
[157J Taylor, J. L. (1976). Topological invariants of the maximal ideal space of a Banach algebra. Advances in Mathematics, 19, 149~206. [158J Thomas, M. P. (1978). Algebra homomorphisms and the functional calculus. Pacific J. Mathematics, 105, 251~69. [159J Thomas, M. P. (1988). The image of a derivation is contained in the radical. Annals of Mathematics (2), 128, 435~60. [160J Thomas, M. P. (2009). The canonical test case for the non-commutative Wermer conjecture. Studia Mathematica, 194, 43~63.
Singer~
[161J Truss, J. (1997). Foundations of Mathematical Analysis. Clarendon Press, Oxford. [162J Vincent-Smith, G. F. (1998). An elementary analytical proof of the spectral radius formula. Expositiones Mathematicae 16, 383~4. [163J Waelbroeck, L. (1954). Le calcul symbolique dans les algebres commutatives. J. Mathematique Pures Appl., 33, 147~86. [164J Weibel, C. A. (1995). An Introduction to Homological Algebra. Cambridge Studies in Advanced Mathematics, Volume 38, Cambridge University Press. [165J Weir, A. J. (1973). Lebesgue Integration and Measure. Cambridge University Press. [166J Zame, W. R. (1984). Existence, uniqueness, and continuity offunctional calculus homomorphisms. Proc. London Mathematical Society (3), 39, 73~92. [167J Young, N. J. (1988). An Introduction to Hilbert Space. Cambridge University Press. [168J Zelazko, W. (1965). Metric generalizations of Banach algebras. Dissertationes Mathematicae (Rozprawy Matematyczne), 41, 1~69. [169J Zemanek, J. (1977). A note on the radical of a Banach algebra. Manuscripta Mathematica, 20, 191~6. [170J Zemanek, J. (1982). Properties of the spectral radius in Banach algebras, in Spectral theory, Banach Center Publications, 8, 579~95, Polish Scientific Publishers, Warsaw. [171 J Zygmund, A. (1959). Trigonometric Series, Volumes I and II, 2nd edition. Cambridge University Press.
I ndex of terms
Abel-summable, 65 absolutely convergent power series, 210, 315 absolutely convex, 34 absolutely summable, 38 absorbing, 34 abstract harmonic analysis, 95 Alexander's subbase lemma, 27 algebra, 155 commutative, 1.56 division, 156 identity, 156 Opposite, 232 primitive, 231 radical, 194, 232 semisimple, 194, 232 simple, 156 unital, 156 analytic function, 31, 314 A-valued, 249 E-valued, 113 weakly, 113 analytic mapping, 318 analytic multi-function, 257 analytic semigroup, 225 analytic space, 149 analytically closed, 257 analytically closure, 257 annihilator, 141 approximate identity, 97, 210 approximation property, 56 Arens' theorem, 183, 184, 207 Arens-Calderon lemma, 342 Arens-Royden theorem, 348, 353 argument principle, 313 Argyos-Haydon theorem, 182 Arzela-Ascoli theorem, 22 Aupetit's lemma, 254 Aupetit's theorem, 257 automorphism, 157 axiom of choice, 8, 14 Bade-Curtis theorem, 246 Baire category theorem, 22, 23, 128 Baire function, 285, 288, 293 real-valued, 285 balanced, 34 Banach algebra, 157 amenable, 239
strongly decomposable, 248 Banach *-algebra, 260 Banach A-module, 229 Banach algebra of power series, 187, 206 Banach extension, 183 Banach function algebra, 189 natural, 190 Banach limit, 121 Banach operator algebra, 160 Banach sequence algebra, 192 Banach sequence space, 47 Banach space, 35 Banach's isomorphism theorem, 131, 150, 151 Banach-Alaoglu theorem, 118 Banach-Steinhaus theorem, 128 Basener's theorem, 208 Bergman space, 83 Bernstein polynomial, 64, 69 Bernstein's approximation theorem, 64 Bessel's inequality, 79 bi-commutant, 180 bidual,46 biholomorphic, 325 bilinear mapping, 59, 129 bimodule, 229 Banach,229 dual, 229 Bishop's theorem, 207 Bishop-Phelps theorem, 127 Borel function, 288 Borsuk's theorem, 149 boundary, 202 Choquet, 208 closed, 202 Shilov, 203 bounded below, 132 bounded operator, 43 Calkin algebra, 261 canonical embedding, 46, 108 Cantor's intersection lemma, 20 Cantor's set, 29 Cantor's theorem, 15, 16 cardinal, 15 Carleson's corona theorem, 209 Carleson's theorem, 96
364 Cauchy's integral formula, 31, 113, 307, 312, 314 Cauchy's integral theorem, 113 Cauchy's theorem, 31, 306 Cauchy-Green formula, 311 Cauchy-Green theorem, 306 Cauchy-Riemann criterion, 306 Cauchy-Riemann equations, inhomogeneous, 310, 318, 319 Cauchy-Schwarz inequality, 70, 263 Cech cohomology group, 225 centre, 180 chain, 10 character, 181, 207 evaluation, 189 character space, 189 Choquet's theorem, 208 closed graph theorem, 133, 151 codimension, 34 Cole's theorem, 208 commutant, 180 complemented, 110 completely regular, 283 completion, 17, 36, 50 complex 147 dual, 148 exact, 147 condItion (F), 48 conjugate index, 39 conjugate-linear map, 42 continuum hypothesis, 15, 247 contour, 30, 311 convergent, absolutely, 56 convergent, unconditionally, 57 convex, 34 convolution product, 159, 195, 196 coordinate functional, 7 coordinatewise convergence, 187 countable, 15 Cousin data, 329 Cousin domain, 329 Cousin problem, first, 329, 337 Cousin problem, second, 337 cyclic vector, 227 C • -algebra, 269 C' -property, 77 C' -subalgebra, 269 Dales' theorem, 247 de la Vallee Poussin kernel, 97 dense range problem, 255 derivation, 229 inner, 229 point, 230 on A, 230 diameter, 20 dimension, 33 Dini's lemma, 249 Dini's theorem, 250
Index of terms Dirichlet kernel, 92, 97, 150 Dirichlet problem, 65, 68 disc algebra, 61, 158, 186 distinguished boundary, 315 Dixon's example, 258 Dixon-Esterle theorem, 183 Dolbeault cohomology group, 322 Dolbeault complex, 322 Dolbeault-Grothendieck lemma, 327 domain of holomorphy, 324 dual Banach space, 45 dual group, 207 dual norm, 45 dual space, 45 Eidelheit's theorem, 243 eigenspace, 171 eigenvalue, 171 approximate, 172 eigenvector, 171 entire function, 31 envelope of holomorphy, 325 equicontinuous, 22 equipotent, 15 equivalent norms, 44 Esterle's theorem, 246, 247 exterior algebra, 319 exterior derivative, 318, 321 exterior product, 320, 321 extreme point, 34, 125 extreme subset, 125 Fejer kernel, 94 Fejer's theorem, 94 Feldman's example, 248 field, 156 ordered, 206 Ford's square root lemma, 262 Fourier coefficient, 66 generalized, 80 Fourier inversion formula, 104 Fourier series, 63, 85 absolutely convergent, 192 Fourier transform, 86, 89, 95, 100, 103, 104, 197, 200, 207 fractional integral semigroup, 225 Frechet algebra, 157, 208 Frechet space, 114 Fuglede's theorem, 277 functional calculus, 215 Borel, 295, 297 continuous, 272 holomorphic, 215, 224, 339, 343 functionally continuous, 182, 325 fundamental theorem of Banach algebras, 166 Gaussian semigroup, 225
365
Index of terms Gel'fand representation theorem, 191 Gel'fand topology, 188 Gel'fand transform, 191 Gel'fand's conjecture, 207, 352 Gel'fand-Mazur theorem, 167 Gel'fand-Naimark theorem, 270, 281 commutative, 272 germs, 212 GNS-construction, 266 Goldstine's theorem, 125 graph, 132 group algebra, 159 Haar measure, 206 Hahn-Banach theorem, 106, 107, 115, 121 harmonic, 65 Harris-Kadison theorem, 235 Hartogs theorem, 314 Hartogs-Rosenthal theorem, 207 Hausdorff metric, 29 Hausdorff-Young theorem, 96 Hermite function, 98 Hermite polynomial, 98 Hilbert space, 71, 73 Hilbertian sum, 267 Holder's inequality, 40 holomorphically convex, 325 homomorphism, 157 hyper-invariant subspace, 171, 298 ideal, 156 left, right 156 maximal, 12, 156, 185 modular, 239 prime, 240 primitIve, 231 proper, 156 quotient, 231 separating, 244 idempotent, 166 trivial, 166 implicit functIOn theorem, 325, 346 independent, 16 index group, 224 infimum, 10 inner product, 70 inner product space, 70 integral, 53, 288, 292 invariant subspace, 171, 298 proper, 171 invariant subspace problem, 301 inverse, 162 inverse function theorem, 32 invertible, 162 left, right, 162 involution, 260 hermitian, 268 isometric, 260 symmetric, 268
Isometric, 17 isometrically isomorphic, 44 isometry, 17, 44 isomorphism, 157 Jacobson radical, 193, 232 Jacobson's density theorem, 236 James space, 59, 121, 209 Johnson's theorem, 239, 243, 246, 257, 262 generalized, 255 Johnson-Sinclair theorem, 245 Kahane-Katznelson theorem, 96 Kahane-Zelazko theorem, 222 Kaplansky's question, 247 Kelley-Vaught theorem, 265 Kolmogorov's theorem, 96 Krein-Milman property, 127 Krein-Milman theorem, 125, 127 Lacey-Thiele theorem, 96 Laplace transform, 105, 211 lattice, 10 Laurent series, 210 LCA group, 207 Lebesgue constants, 97 Lebesgue spaces, 52 left A-module, 226 Banach, 227 lexicographic ordering, 9 linear homeomorphism, 44, 138 linear map, 42 linearly dependent, independent, 11 linearly homeomorphic, 44, 138 Liouville's theorem, 32, 113 real-part, 32 Lipschitz functions, 58 Lipschitz algebra, 209 LMC algebra, 157 local maximum modulus theorem, 352 local minimum modulus theorem, 353 local uniform convergence, 115, 317 locally compact group, 206 locally convex space, 114 metrizable, 114 Lomonosov's theorem, 177 Losert's theorem, 239 lower semi-continuous, 18 main bounded ness theorem, 246 maximal, maximum, 10 maximum modulus theorem, 31, 316 maximum principle for the spectral radius, 253 maximum set, 202 meagre, non-meagre, co-meagre, 22 measure, 287 Borel, 287
366 complex, 287 Mergelyan's theorem, 207 merom orphic function, 210 metric space, 16 complete, 16 Michael's problem, 183 Mi!utin's theorem, 149 mmimal, minimum, 10 Minkowski functional, 127 Minkowski's inequality, 40, 56 Mittag-Leffler theorem, 23 module homomorphism, 227 module isomorphism, 227 module 227 cyclic, 227 irreducible, 227 separating, 244 monomorphism, 157 monotone class, 285 bounded-, 286 monotone-convergence property, 291, 295 monotone-convergence theorem, 291 Morera's theorem, 31 Miiller's example, 258 multiplicative semigroup, 155 nest, 136 stabilizes, 136 lllipotent, 166 norm, 34 submultiplicative, 157 norm-preserving extension, 109 normal, 261 normed algebra, 157 normed space, 34 nowhere dense, 22 null sequence, 135 numerical radius, 280 numerical range, 280 n-transitive, 236 Oka's lemma, 333 Oka-Cartan theory, 344 Oka-Wei! theorem, 336 open mapping lemma, 129 open mapping theorem, 31, 131, 140, 151 operator norm, 44 operator, 43 adjoint, 76 approximable, 47, 160 bounded, 43 bounded below, 52 compact, 46, 160, 175 diagonal, 58 dual, 119 finite-rank, 47, 160 invertible, 51, 78, 165 left, right, 51 normal, 76, 173
Index of terms positive, 77, 275 rank-one, 47 self-adjoint, 76 spectral, 301 unitary, 76, 173, 299 ordinal, 15 limit, 15 orthogonal, 73 orthogonal family, 73 orthogonal projection, 75 orthonormal basis, 79 sequential, 79 orthonormal sequence, 79 orthonormal subset, 79 pairing, 116 non-degenerate, 116 parallelogram rule, 71 Parseval's equation, 80 partial ordering, 8 partially ordered set, 9 partition of unity, continuous, 111 partition of unity, smooth, 310 path integral, 30 closed, 30 track,30 peak point, 202 local,350 peak set, 202 local,350 peak-point conjecture, 208 Phillips-Sobczyk theorem, 121 Plancherel formula, 104 generalized, 104 Poincare's theorem, 325 point spectrum, 171 pointwise monotone limit, 285 POisson integral, 67 Poisson kernel, 67, 91 polar decomposition, 278, 299 polarization identity, 71 general,71 polyball, 313 polydisc, open, closed, 313 polynomial hull, 179, 331 polynomial polyhedron, 333 polynomially convex, 179,331, 332 polynomially generated, 179, 331 Pontryagin duality theorem, 207 poset, 9 positive, 275 positive linear functional, 263, 288 principal component, 223 principal square root, 221, 262 principle of mathematical induction, 12 principle of transfinite induction, 13 product space, 8, 26 product topology, 26 projection, 171
367
Index of terms Pythagoras's theorem, 73 p-form, 320
a-closed, 322 degree, 320 quasi-nilpotent, 166, 233 quotient algebra, 161 quotient module, 227 quotIent norm, 134 quotient semi norm, 133 radical, 193 Jacobson, 193, 232, 239 prime, 240 strong, 239 Radon-Nlkodym property, 127 Ransford's proof, 257 Ransford's theorem, 258 rational hull, 338 rationally convex, 338 real-differentiable function, 305, 314 reflexive, 109 representation, 226 faithful, 226 left-regular, 228 normed,226 continuous, 227 standard, 231 residue theorem, 312 resolvent set, 177 Riemann-Lebesgue lemma, 86, 96, 198, 207 Riesz's representation theorem, 75 Riesz's lemma, 55 Riesz-Fischer theorem, 89 Ringrose's theorem, 283 Rossi's local peak set theorem, 350 Runde's theorem, 239 Runge's theorem, 212, 213 Russo-Dye theorem, 284 Sakai's theorem, 283 Schauder's theorem, 144 Schwarz's lemma, 32 self-adjoint, 190, 261 semi-inner product, 70 semi-inner product space, 70 semigroup, 159 semigroup algebra, 159 weighted, 159, 188 seminorm, 34 seminormed space, 34 separable, 19 separating ideal, 244 separating module, 244 separating space, 135 separation theorem, 123 sesquilinear form, 75 bounded,75
pOSItive, 75 self-adjoint, 75 shIft, left, right, 58 Shilov's idempotent theorem, 345 short exact sequence, 147 a-field, 287 Baire, 287 Borel,287 simple functIOn, 52 Singer-Wermer theorem, 230, 239 smooth function, 309 smooth mapping, 318 spectral radius, 169 spectral radius formula, 169 spectral theorem, 294 spectral three-circles theorem, 252 spectrum, 165, 171 approximate point spectrum, 172 compression, 182 joint, 330 surjectivity, 182 stability theorem, 137 *-algebra, 260 *-closed, 260 *-homomorphism, 260 *-ideal, 260 *-radical, 264 *-representation, 261 faithful, 267 universal, 267 *-semisimple, 264 *-subalgebra, 260 state, 263 pure, 284 state space, 263 Stone-Weierstrass theorem, 60 complex, 61 Stone-Cech compactification, 272, 283, 294 strong-operator topology, 119 sub algebra, 60, 157 bi-commutant, 181 inverse-closed, 180 sublattice, 10 sublinear functional, 106 subspace topology, 26 sum mabie function, 290 support, 309 supremum, 10, 15 surrounds, 30 Swiss cheese, 209 Thomas's theorem, 239 Tietze extension theorem, 130 Titchmarsh's theorem, 206 topological algebra, 182 topological divisor, 183 topological group, 223 topological space, 18 normal,23
368 topological vector space, 121 topology, 18 base, 25 subbase, 25 weakest, 25 totally bounded, 20 totally ordered, 9 trace, 83 trigonometric polynomial, 62 Tychonoff's theorem, 27 uncountable, 15 uniform algebra, 158 uniform boundedness theorem, 128 uniform norm, 36 unique complete norm, 241 uniqueness-of-norm theorem, 243 unit, 162 unitary, 261 upper semi-continuous, 18 Urysohn's lemma, 24 vector space, 33 Vesentini's theorem, 257 Vidav-Palmer theorem, 283 Volterra algebra, 196 von Neumann algebra, 283
Index of terms weak spectral maximum principle, 250 weak topology, 116, 117 weak-* topology, 117 weak-*-continuous, 118 weak-operator topology, 120, 281 weakly bounded, 118 weakly compact, ll8 Wedderburn's principal theorem, 240 Wedderburn's theorem, 237 Wedderburn-Art in theorem, 240 Weierstrass' polynomial approximation theorem, 62 trigonometnc polynomials, 62 weight function, 187, 196 well ordered, 12 well-ordering principle, 12, 14 Wiener algebra, 192 Wiener's theorem, 193 winding number, 30 Woodin's theorem, 247 Zame's theorem, 339, 344 Zemanek's theorem, 251 Zermelo-Fraenkel axioms, 16 zero set, 23 Zorn's lemma, 10
Index of symbols
a· E, 226 a . 1;, 226 [a, b], 7, 235 A EB.l B, 73 A(~), 61, 95, 186, 260 A(JR), 198 A(Z), 86, 89, 197 A ·1;,227 A"", B, 157 A+, 161 A+,263 ii, 192 A.l, A.l.l, 73 A O P,232 Apos,275 Asa, 261 AC,16 A(E, F), 47, 56 A(E), 47, 160 bv, 209 B(a;r), 16, 138,313 B(K), BJJl.(K), 286 BP(U),322 B(E,F),44 B(E, F; G), 59 B(E), 45, 159 B(H),261 B(X), Ba(X), 287
co,38 co(Z), 38 cardinals No, Nl, 2 ND , 15 cony A, 34 C(K), 37, 60, 158 CJJl.(K),37 CoCK), 37, 158, 260 C oo(K),37 C(U),31 Cl(U),305 C~X), CJJl.(X), C(X)+, 18 C (X), 58 Cb(X), C~(X), 37 C(li),42 CR(li),29 CI(li),83 Ck(li), 208
Coo (li), 208, 246 C(T), 65, 82 Co(JR),37 C[a, b], 37 CE [a, b], 110 C*(a),272 CH, 16, 247 C, 7 C X ,9 C
XT, 7 c, 15 Card, 15 diam(E),20 dimE,33 dZk, 318 dZtt /\ dZ'2 /\ ... /\ dz,,,, 321 DI(U), 306, 314 DN,91 6" 39 ~, 7, 61 ~(a; r), 313 8,318 aj, 8J, 305 akj, akj, 314
exC, 34, 125 expA,220 e" /\ e'2 /\ ... /\ e,l" 319
E(>"), 171, 175, 220 E ~ F, 227 E*,45 E[r],35 (E;P),114 lO(u), 318 ll(U),318 lP(U),320 l(U),321
370 ex, 186, 189 en, 47
f * g, 195 f, i f, 285 f(k),66 34 F tB G, 34, 110, 151 (F), 48 F, 89,100 F(E,F),47 F(E), 47, 160 J' = IC[[X]], 187, 210 J'n = IC[[X1, ... ,Xn]], 315 A, 181, 189
F+ G,
Index of symbols !inS, 33 linX,35 lip",K,58 £1 (JR.), 194, 260 £1(JR.+,w), 196,211,247 £1 ('Jr), 194 £1(JR.+),196 £P[a, b], £P('Jr), £P(JR.), 53 Lip",K, 58, 209 AP, A, 319 £(E,F),42 £(E),42 £(f), 211
MA,185 J1. *w, 323
G(A),162 Go(A),223 G(E), 165 Gr(f), 132 Q, 191 [,],30 r(A), 202 I'o(A), 202 H H
MI m ,n,43 MIn, 43, 159 norms
·11,34 .11 1 , 37 ·11=,38
·ll w ' 188 ·ll p ' 35, 40
(K;Z),225
. 11 p,q '
1 (A;Z),348
'l s ,36
I
H=(U), 58, 209 1{,98 1{P(U),322 imT,42 inner product (-, . ), 70 int~, 7 Ie, 42 J(f), 289 1(f),290 ll,7
,l(A), 193, 232 ,l*(A),264 kerT,42 R, 179, 331 KA, 205,279 KN,93 K(E, F), 46, 56 K(E), 47, 160, 176 {! l(s), 38 {!1(Z+,w),188 {!2, 72 {!=,38 {! (S), {!R"(S), 36 e=(Z),38 {!P 39 40
=
ee.:
50
n(,; z), 30 N,7
O(U), 31, 114, 161, 188, 246, 314 w 1\ 'f/, 321 OK, 212 Ord,15 pairing (-, .), 116 P(K), R(K), O(K), A(K), 158, 332 PD leU), 305, 313 PA, 220, 294 P(S),7 7r,';"n (f), 342 1Q,1Q+,7 Q(A), 166, 233
R(K),187 R(U),213 RaP. ), 167 RA(a), 177 JR. JR.+ JR.- JR.+. 7 JR. ,9:285' , PA(a),169
x
s, 37, 47 suppf,309 S(A),263 SE,35
ISBN 978-0-19-920654-4
1111111 III
9 780199 206544