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!
llp. 0 such that, for any choice of n 9 1~ and x l . . . . . Xn 9 X, 1. A Haar-like system fjk we define using sets of positive measure (A., jKjj=Ok= 0 such that
(51)
Martingales and singular integrals in Banach spaces
257
To see this, let K denote/72, the Hilbert space of sequences x - - ( x j ) j ~ / o with X j E I-I and IIx IlK -- (Z~--0 Ilxk 112) 1/2 finite. Let F and G be the K-valued martingales defined by
Fn=(fn,O,O .... )
and
G n - - ( d o . . . . . d,,O,O . . . . ).
Their respective difference sequences D and E satisfy lID, IlK - l I E , IlK. In addition,
IIF~ IlK - I I f ~ IIn and IIGn IlK -- (Y~'~=0 Ildk II2) i/2. So, by Theorem 14, IS(f ) I p-
IIGllp ~< (p* - 1)llFllp-
(p* - 1)llfllp.
This gives the left side of (51) and IIFIIp ~< (P* - 1)llGIIp gives the right side. Apart from the constants, special cases of (51) for real-valued martingales go back to the 1920s and 1930s with the work of Khintchine, Littlewood, Paley, Marcinkiewicz, and Zyground on Rademacher series, Haar series, and sums of independent random variables. Apart from the constants, the inequality was proved for all real-valued martingales by Burkholder [28]. Many other proofs are now known, including Gundy's proof [86] based on his martingale version of the Calder6n-Zygmund decomposition. Pittenger [125] proved that if p ~> 3, then the best constant on the right side of (51) is p - 1. The proof given above has the advantage of yielding the best constant for the right side if p ~> 2 and the best constant on the left if 1 < p ~< 2; in both of these cases the constant is p - 1. The best constants in the other cases are not known. Inequality (51) does not carry over with finite constants to any Banach space that is not isomorphic to a Hilbert space. This is clear from a result of Kwapiefi [100]; see the proof of the first part of Theorem 14 above. The following theorem provides a martingale square function inequality in the setting of rearrangement invariant function spaces; see Lindenstrauss and Tzafriri [ 107] for background on such spaces. If p ~ [1, oo), then L p (0, 1) and many other Orlicz spaces are examples. Here B is a rearrangement invariant function space on (0, 1) and the upper Boyd index qB of B plays an important role. It is the least q 6 [ 1, oo] such that the norm of the dilation operator Ds on B satisfies
IIDsII ~< s llq
for all s E (0, 1),
where ( D s x ) ( t ) = x ( t / s ) if t 6 (0, s], ( D ~ x ) ( t ) = 0 if t E (s, 1), and the function x : ( 0 , 1) ~ R belongs to B. The following theorem is due to Johnson and Schechtman [96]; also see Antipa [4] and Hitczenko [90]. THEOREM 15. Let B be a rearrangement invariant function space on (0, 1) with finite upper Boyd index qB. There exist positive real numbers c and C depending only on qB such that if f is a real-valued martingale on (0, 1) with respect to Lebesgue measure on the er-algebra of Borel subsets of (0, 1), then
clls(f)
B
IIf*llB
CIIs(z)IIB.
(52)
Conversely, if either side of this two-sided inequality holds for all such martingales (the filtrations must also vary), then the upper Boyd index qB is finite.
258
D.L. Burkholder
This contains the Burkholder-Davis-Gundy inequality [43], which is the inequality
c•
<~EcI)(f*) <~C• EcI)(S(f)),
where q~ is a nondecreasing continuous function on [0, oo], convex on [0, ~ ) , with q~(0) = 0, and satisfying r ~< yq~(,k) for all )~ > 0. The positive real numbers c• and C• can be chosen to depend only on the positive real number y. The condition of convexity can be omitted if the martingales f are assumed to be dyadic or otherwise satisfy the conditions assumed in the earlier paper of Burkholder and Gundy [42]; see Burkholder [29] for a short treatment of both cases. For the convex case, the associated rearrangement invariant function space B is an Orlicz space. Its upper Boyd index (see p. 139 of Lindenstrauss and Tzafriri [107] and the references given there) is qB = inf{q" sup
clgO~t)/c1900tq < ~}.
~.,t>~l
5.5.
Differential subordination of harmonic functions
The function U defined in (47) has other applications. Here is one of them. Let n be a positive integer, D a domain of R n, H a real or complex Hilbert space, K the Hilbert space I-In, u and v harmonic functions on D with values in H, and Vu = (OlU . . . . . OnU). Suppose that v is differentially subordinate to u: for all x E D,
IIv (x)ll
Ilvu(x)ll .
(53)
Fix a point ~ E D and suppose that Ilw(~)lln ~ Ilu(~)lln. Let Do be a bounded subdomain satisfying ~ Do C Do U ODo C D. Denote by/z the harmonic measure on 0 Do with respect to ~. If 1 ~< p < ec, let
IIu IIp = sup
Do
Do
IIu II~ d#
,
where the supremum is taken over all such Do. Then (54)
Ilollp ~ ( p * - 1)llullp.
To see this, recall that U majorizes the function V, where U and V are as in the proof of part (ii) of Theorem 14. It is not hard to show that U (u, v) is superharmonic on D. Therefore,
f0Do
V(u,
v)d#
<~fo
Do
U(u, v)d/z
~< U(u(~),
v(~)) <~O,
which gives (54). See Burkholder [39] for the details and an application to Riesz systems.
Martingales and singular integrals in Banach spaces
259
In the classical setting, n = 2, D is the open unit disk, ~ - - 0 , v(0) = 0, H = R, and the function v is conjugate to u so the Cauchy-Riemann equations give equality in (53). For (54), the conjugacy condition has been replaced by the weaker condition of differential subordination, a condition that makes sense for domains of R n . It is not yet known whether or not p* - 1 is the best constant in (54), as it is for the analogous martingale inequality proved above. For related work, see Burkholder [41], Wang [137], Bafiuelos and Wang [9,10], and Choi [49,50].
5.6. The Beurling-Ahlfors transform The function U of (47) is useful also in the study of the Beurling-Ahlfors transform, the singular integral operator B on L p (C), 1 < p < oc, that maps f to its convolution with - 1/re z 2. It is important in the study of quasiconformal mappings and some other parts of analysis; see, for example, the paper of Iwaniec and Martin [93] and the references given there. The norm liB lip of B is not yet known precisely, except for [IBII2 - 1, but such knowledge would be helpful in some applications. It is known that p*-l~
for a l l l < p < o o .
See Lehto [104] or Iwaniec [92] for the left side. The right side is due to Bafiuelos and Wang [9], where they use U in their proof. Iwaniec [92] conjectures that IIBIIp = P* - 1 and, indeed, this would follow if a conjecture of Bafiuelos and Wang could be shown to be true; see Bafiuelos and Lindeman [8]. The Bafiuelos-Wang conjecture is that
fc
U(Of, m Of) ~< 0
for all infinitely differentiable functions f :C ---> C with compact support, where
1 Of---~
Of i a--x- Oy
and
af--~
l(O
~xx+i-O-Ty .
Some numerical support for these and several closely related conjectures is contained in Baernstein and Montgomery-Smith [7], as well as a remarkable integral identity for the special function U.
6. Martingale characterization of spaces with the R N P
In the study of Banach spaces, the Radon-Nikod3)m property (RNP) has a long and rich history going back at least to the 1930s. Some of the early background is sketched in B.J. Pettis's foreword and the authors' introduction to Vector Measures by Diestel and Uhl [56]. In this influential book, the Radon-Nikod3)m property plays a major role, as it does in a number of articles in this Handbook. After recalling the definition of the RNR we
260
D.L. Burkholder
examine just one of the many necessary and sufficient conditions on a Banach space in order that it have the RNP. Let B be a Banach space and f" a a-algebra of subsets of S2. Here a function q9 : U ~ B is a vector measure if it is a-additive on U. Let I~01denote the measure (from U to [0, cx~]) of total variation. Then q) hasfinite variation if I~oI(s2) is finite. The Banach space B has the Radon-Nikodym property with respect to (I2, 3c, P) if for each vector measure q) of finite variation that is absolutely continuous with respect to the probability measure P there is an integrable function g: S2 --+ B such that ~0(A) = fAg dP for all A 6 U, and B has the RNP if it has the RNP with respect to every probability space. Reflexive spaces, hence UMD spaces, have the RNP, but some nonreflexive spaces such as s also have it; see pp. 217-219 of Diestel and Uhl [56] for three long and helpful lists: a list of some spaces that have the RNP, a list of some that do not, and a list of conditions on a Banach space, each equivalent to the RNP. After Doob's discovery [57] that L l - b o u n d e d martingales converge almost everywhere, it was natural to ask: Under what conditions on a Banach space B, do L l - b o u n d e d martingales converge almost everywhere? Partial results were obtained by Chatterji [47] and Scalora [133]. The complete answer is given in the following key theorem. THEOREM 16. A Banach space B has the RNP if and only if every L~-bounded martingale converges almost everywhere. The "only if" part of this theorem is due to A. Ionescu Tulcea and C. Ionescu Tulcea [91 ]; see also the note by Bellow [ 11 ]. For the "if" part, see Chatterji [48]. For much else that is related, see the books by Diestel and Uhl [56], Bourgin [23], Egghe [70], and Edgar and Sucheston [69]. PROOF (ONLYIF). In all of its essentials, the following proof is the same as Lamb's in the scalar-valued case; see Lamb [101]. Suppose that B has the RNP and f is a Bvalued martingale on (s U, P) with Ilflll finite (recall that Ilflll = SUPn~>0 Ilfn Ill)- Then f converges a.e. as can be seen as follows. Let )~ > 0 and define r:S-2 ~ [0, cx~] by r(co) = inf{n: Ilfn(co)ll > )~}. We shall see below that F = (frAn)n>~O converges a.e. This implies that f converges a.e. on the set {f* ~< )~} since f = F on this set. The complementary set satisfies P(f* > )~) ~< Ilflll/)~ by the Doob maximal inequality applied to the submartingale (llfnll)n>>O, so P(f* < ec) = 1 by the finiteness of Ilflll. Consequently, f converges a.e. The first step in the proof that F converges a.e. is to show that the expectation of F* is finite. Observe that F* ~< )~ on {r = ec} and F* = IIf~ II on {r < ec}. Therefore, E F* ~< )~ + limn--,oc fr<<,nIIf~ II where the integral is equal to n
~ f{r=k} k=0
Ilfkll ~ ~
k=0
f ~ =k/ IIfn II ~ IIfn II1 ~ IIf Ill
since (llfn II)n~0 is a submartingale and {r = k} 6 3ck. Here (Un)n~O is any filtration with respect to which f is a martingale. Thus, E F* ~< )~ + IIf Ill.
Martingales
and singular
i n t e g r a l s in B a n a c h
spaces
261
The next step is to show that there is an integrable function Foc such that
(55)
F, = E ( F ~ I 5t-n) for all n ~> 0.
Let . , 4 - ~n~=of'n and ~ be the least a-algebra containing the algebra A. Define r ".T'~ ~ B by qgn(A) -- fA Fn and the measure ~ ".T'~ ~ [0, cx~) by ~p(A) -- fA F*, and note that Ilqgn(A)ll ~< 7r(A). This leads to a vector measure qg~ : ~ ' ~ ~ B satisfying ~p~(A) = l i m n ~ ~pn(A) for all A E ~ as follows. By (1), F is a martingale: Fn -- Y]~=0 l{k~
A EA
where A A Aoc denotes the symmetric difference of A and Aoc. The limit q%c not only exists and but also satisfies IIq%c(A)ll ~< ~p(A) for all A in Uoc. So it is absolutely continuous with respect to ~k, hence with respect to P. By the RNP, there is an integrable 9coo-measurable function Foc such that
q)oc(A) -- f a F ~ dP
for all A 6 Foc.
In particular, this holds for all A E 9c, and gives
fA
in -- q?n(A) -- q)oc(A) -- fA F ~ .
This proves (55), which yields the a.e. convergence of F as follows. Let S be the family of all simple functions Goc : S2 --+ B that are measurable with respect to the algebra A. The martingale G defined by Gn -- E(Goc ] ,T'n) converges a.e. since Gn = G ~ for all large n. By the contraction property of conditional expectations, 11Fn - Gn I11 ~ ]]Foc - Goc ill. Therefore, lim sup IIFn m , n -+ oc
-
Fm I1~ ~< 2 inf IIFoc - Goc
I1, = 0.
G oc E S
Since (llFm+k - Fm II)k~>0 is a submartingale, Doob's inequality gives XP (
sup
IIFk -Fm It > k) ~< if Fn -Fm Ill
for all X > 0.
m <~k <~n
Accordingly, F converges a.e. so f does also. This completes the proof of the "only if" part of Theorem 16. PROOF (IF). Let (I2, U, P) be a probability space and ~0 : U ~ B a vector measure with finite variation that is absolutely continuous with respect to P. We can assume that Iq0l, the total variation measure, satisfies Iq0[(~2) = 1 so that (I-2, .T', I~0[) is also a probability space.
D.L. Burkholder
262
We can also assume that ~- is countably generated (see Johnson and Lindenstrauss [95] in this Handbook). Consequently, there is a sequence of finite partitions Hn of I2, each set in Hn being a union of sets in Hn+l, such that ~ is the smallest a-algebra containing all of the partition sets. By the classical Radon-Nikod3~m theorem, there is an integrable function g such that I~ol(C) - fc g dP for all C E ~ . We shall show below that there is a bounded B-valued ~-measurable function f ~ such that
qg(C) - fc f ~ dl~ol for all C 6 .Y.
(56)
Therefore, qg(C) - fc f ~ g dP for all C E 9t'. Consequently, B has the RNP. The existence of f ~ is proved as follows. Let H + = {A E Hn: Iqgl(A) > 0} and define fn : S-2 ---->B by qg(A) ~IA. AEH +
I~ol(a)
Note that IIin II ~ 1. Let 9t'n be the algebra generated by Hn : if B E f n , then B is the union of a family of partition sets in Hn. It is easy to see that the sequence f is a martingale relative to (~n)n)0 on the probability space (I2, 9t', I~ol), and that qg(B) = f , fn dl~ol for all B E 9t'n. So by the assumption of the "if" part of the theorem, there is a a B-valued ~ measurable function f ~ such that 2 ~> IIA - f ~ II ~ 0 almost everywhere as n increases to oo. Here define ~ ' ~ --+ B by O(C) = f ~ dl~01 and note that I~Pl ~< I~ol on 3t'. If B 6 3t'n and C E ~ , then II~0(n) - 7s(n) II <~f~ Ilfn - f ~ l l algol---> 0 and
fc
II o(c> -
II: [l o(c> -
- ~o(B)
+
II
21~ol(B A C).
But inf{l~0l(B A C): B E ~n for some n ~> 0} = 0 and (56) follows. This completes the proof of Theorem 16. [3 The martingale characterization of spaces with the RNP given in Theorem 16 should be compared with the following martingale-transform characterization of UMD spaces [30]. THEOREM 17. A Banach space B is UMD if and only if every -+-l-transform of every
L~-bounded martingale converges almost everywhere.
7. Martingale characterization of spaces with the A R N P
A complex Banach space B has the analytic Radon-Nikodym property (ARNP) if for every vector measure q9:/3[0, 2zr) --+ B of finite variation satisfying
fo ree -in~ dqg(0) -- 0 for all integers n < 0,
(57)
Martingales and singular integrals in Banach spaces
263
there is an integrable function g:[0, 2rr) ~ B such that q)(B) = fB g(O)dO for all B B[0, 27r), the o--algebra of Borel sets of [0, 27r). For example, see Bukhvalov [26], Dowling [64], and Hensgen [88]. An earlier and equivalent definition of the ARNP is contained in the work of Bukhvalov and Danilevich [27]. Let D be the open unit disk of C. The Hardy class H~ (D) is the family of analytic functions F : D ~ B satisfying
sup fo 27r IIF (re iO) ][P dO < oo, O
and H~C(D) is the family of the bounded analytic functions F. A complex Banach space B has the ARNP if and only if for each F ~ H~C(D), the limit of F(re i~ as r t 1 exists for almost all 0 6 [0, 27r). The same class of Banach spaces B is obtained if H ~ ( D ) is replaced by H~ (D) where p ~> 1. There is an important martingale characterization of complex Banach spaces with the ARNP due to Edgar [68]. Suppose that (X2, U, P) is a probability space on which is defined an independent sequence Z of Steinhaus random variables: Zn is uniformly distributed on the unit circle, the boundary 0 D of D. Let 5On be the a-algebra generated by Z0 . . . . , Zn. Then an analytic martingale (starting at x 6 B) is any sequence f of the form fn -- x + ~-~=1 vkZk where the function vk" X2 --> B is fk_l-measurable and integrable. By (1), f is a martingale. Typically, P is the Haar measure on the infinite polydisk s = (0 D) ~176 and the value of fn at a point (e i~176 , e i01 . . . . ) in s is written as fl
f , (0) = x + ~
~0k(00 . . . . . Ok-1 )e i~ 9
k=l
THEOREM 18 ([68]). A complex Banach space B has the ARNP if and only if every L 1bounded analytic martingale converges almost everywhere. This theorem and the corresponding theorem for the RNP (Theorem 16) have quite different proofs. We omit Edgar's proof. The two theorems yield an obvious corollary, which is also a consequence of (57) and the vector-valued version of the E and M. Riesz theorem due to Ryan [132], as was observed by Bukhvalov [26]. COROLLARY 19. A complex Banach space with the RNP has the ARNP. But these two properties are not equivalent. The Banach space L~ (0, 1) has the ARNP (see the next section) but it is well known that it does not have the RNP (see Diestel and Uhl [56]). Further work related to the ARNP includes Dowling and Edgar [65], Garling [77], Ghoussoub and Maurey [80], Ghoussoub, Maurey, and Schachermayer [82], Ghoussoub, Lindenstrauss, and Maurey [81 ], and Bu [24].
D.L. Burkholder
264
8. Characterization of AUMD spaces 0 (B) the least ot 6 [1 cx~] such Here let B be a complex Banach space, 0 < p < oc, and Otp that 0 (B) IIf IIp Ilgllp ~ Ofp for all f and g where f is a B-valued analytic martingale (see Section 7) and g is a -+-l-transform of f (see Section 2). If the numbers E~ E {1 , - 1} in the definition of g are replaced by complex-valued U(~_l)v0-measurable functions v~ that have their values in the closed unit disk, then the best constant may change a little but its finiteness, or lack of it, will not be affected. There will be no change if the vk have their values in [ - 1 , 1]. The space B is an A UMD space if Ctp 0 (B) is finite for some p 6 (0 c~); equivalently, for all p ~ (0, 0o). This equivalence can be seen by using the inequality IIf*llp ~< el/pllfl]p, which holds for all p E (0, c~) and all analytic martingales [77], using the trivial inequality IIf IIp ~< IIf* IIp, and observing that the argument in the dyadic case (see Section 2 above) shows that if Ilg*llp ~< cp(B)llf*llp holds for some p E (0, cx~) and all f and g as above, then it holds for all p 6 (0, ~x~). Notice that if p E (1, oc), then Otp 0 (B) ~< tip(B) So a UMD space is an AUMD space. But the converse is not true: the space L ~ (0, 1) is AUMD but is not UMD because it is not reflexive. That L ~ (0, 1) is AUMD is an immediate consequence of the fact that C is UMD, hence is AUMD, and Theorem 9 of Garling [77]: if B is AUMD, then so is L~ (0, 1) for all p E [1, cx~). Note that Ilf* Ill ~ ellfll l, a special case of Garling's inequality for analytic martingales, implies that an AUMD space is an ARNP space" if B is AUMD and f is an L~-bounded analytic martingale, then f * is integrable so there is an integrable function foc such that fn = E (fee [ .T'n) as in (55), which implies that f converges almost everywhere. Therefore, by Edgar's theorem above, the space B is ARNE Using multipliers, Blower [ 16] proved the following characterization of AUMD. THEOREM 20. A complex Banach space B is AUMD if an only if there is a real number y such that
fo re
• n~O
mnan einO dO <, ~' fo 2rc
Z ane inO
dO
n >~O
for all sequences (an)n>~O in B with an = 0 for all large n, and all complex-valued sequences (mn)n>~O such that Imnl ~ 1 and n2[mn+l - 2mn + mn-ll ~ 1. This implies that B is AUMD if and only if for all such sequences m, the linear transformation Tm defined by Tm (e in0) = mne inO extends to a bounded linear operator on the Hardy space H 1 (0 D). Piasecki [ 120,121 ] has two quite different characterizations of AUMD. The first one can be viewed as the subharmonic analogue of the ~'-convex characterization of UMD given in Theorem 1. The other characterizes AUMD in terms of tangent analytic martingales: Piasecki proves for AUMD an analogue of McConnell's inequality [ 112] for tangent martingales in UMD spaces.
Martingales and singular integrals in Banach spaces
265
References [1] D.J. Aldous, Unconditional bases and martingales in L p (F), Math. Proc. Cambridge Phil. Soc. 85 (1979), 117-123. [2] H. Amann, Linear and Quasilinear Parabolic Problems, I: Abstract Linear Theory, Birkh~iuser, Basel (1995). [3] T. Andr, Contractive projections in Lp spaces, Pacific J. Math. 17 (1966), 391-405. [4] A. Antipa, Doob's inequality for rearrangement-invariant function spaces, Rev. Roumaine Math. Pures Appl. 35 (1990), 101-108. [5] N. Asmar, E. Berkson and T.A. Gillespie, Representations of groups with ordered duals and generalized analyticity, J. Funct. Anal. 90 (1990), 206-235. [6] N.H. Asmar, B.P. Kelly and S. Montgomery-Smith, A note on UMD spaces and transference in vectorvalued function spaces, Proc. Edinburgh Math. Soc. 39 (1996), 485-490. [7] A. Baernstein and S. Montgomery-Smith, Some conjectures about integral means of Of and 0 f, Acta Universitatis Upsaliensis 64 (1999), 92-109. [8] R. Bafiuelos and A. Lindeman, A martingale study of the Beurling-Ahlfors transform in R n , J. Funct. Anal. 145 (1997), 224-265. [9] R. Bafiuelos and G. Wang, Sharp inequalities for martingales with applications to the Beurling-Ahlfors and Riesz transforms, Duke Math. J. 80 (1995), 575-600. [10] R. Bafiuelos and G. Wang, Orthogonal martingales under differential subordination and applications to Riesz transforms, Illinois J. Math. 40 (1996), 678-691. [ 11 ] A. Bellow, For the historical record, Measure Theory (Oberwolfach, 1983), Lecture Notes in Math. 1089, Springer, Berlin (1984), 271. [12] E. Berkson, T.A. Gillespie and P.S. Muhly, Abstract spectral decompositions guaranteed by the Hilbert transform, Proc. London Math. Soc. 53 (1986), 489-517. [13] E. Berkson, T.A. Gillespie and P.S. Muhly, Generalized analyticity in UMD spaces, Arkiv frr Math. 27 (1989), 1-14. [ 14] O. Blasco, Hardy spaces of vector-valued functions: Duality, Trans. Amer. Math. Soc. 308 (1988), 495507. [15] O. Blasco, Boundary values of functions in vector-valued Hardy spaces and geometry on Banach spaces, J. Funct. Anal. 78 (1988), 346-364. [16] G. Blower, A multiplier characterization of analytic UMD spaces, Studia Math. 96 (1990), 117-124. [ 17] S. Bochner and A. E. Taylor, Linearfunctionals on certain spaces of abstractly-valued functions, Ann. of Math. 39 (1938), 913-944. [ 18] J. Bourgain, Some remarks on Banach spaces in which martingale difference sequences are unconditional, Ark. Mat. 21 (1983), 163-168. [19] J. Bourgain, Extension of a result of Benedek, Calder6n and Panzone, Ark. Mat. 22 (1984), 91-95. [20] J. Bourgain, Vector-valued singular integrals and the H 1-BMO duality, Probability Theory and Harmonic Analysis, Marcel Dekker, New York (1986), 1-19. [21] J. Bourgain and W.J. Davis, Martingale transforms and complex uniform convexity, Trans. Amer. Math. Soc. 294 (1986), 501-515. [22] J. Bourgain and H. Rosenthal, Martingales valued in certain subspaces of L |, Israel J. Math. 37 (1980), 54-75. [23] R.D. Bourgin, Geometric Aspects of Convex Sets with the Radon-Nikod~m Property, Lecture Notes in Math. 993, Springer, Berlin (1983). [24] S. Bu, On the analytic Radon-NikodSm property for bounded subsets in Banach spaces, J. London Math. Soc. 47 (1993), 484--496. [25] A.V. Bukhvalov, Continuity of operators in spaces of vector functions with applications to the theory of bases, J. Sov. Math. 44 (1989), 749-762. [26] A.V. Bukhvalov, On the analytic Radon-NikodSm property, Function Spaces, B.G. Teubner, Stuttgart (1991), 211-228. [27] A.V. Bukhvalov and A.A. Danilevich, Boundary properties of analytic and harmonic functions with values in Banach space, Math. Notes 31 (1982), 104-110.
266
D.L. Burkholder
[28] D.L. Burkholder, Martingale transforms, Ann. Math. Statist. 37 (1966), 1494-1504. [29] D.L. Burkholder, Distribution function inequalities for martingales, Ann. Probab. 1 (1973), 19-42. [30] D.L. Burkholder, A geometrical characterization of Banach spaces in which martingale difference sequences are unconditional, Ann. Probab. 9 (1981), 997-1011. [31] D.L. Burkholder, Martingale transforms and the geometry of Banach spaces, Probability in Banach Spaces, III (Medford, 1980), Lecture Notes in Math. 860, Springer, Berlin (1981), 35-50. [32] D.L. Burkholder, A nonlinear partial differential equation and the unconditional constant of the Haar system in L p, Bull. Amer. Math. Soc. 7 (1982), 591-595. [33] D.L. Burkholder, A geometric condition that implies the existence of certain singular integrals of Banachspace-valued functions, Conference on Harmonic Analysis in Honor of Antoni Zygmund, Chicago, 1981, Wadsworth, Belmont, CA (1983), 270-286. [34] D.L. Burkholder, Boundary value problems and sharp inequalities for martingale transforms, Ann. Probab. 12 (1984), 647-702. [35] D.L. Burkholder, An elementary proof of an inequality of R.E.A.C. Paley, Bull. London Math. Soc. 17 (1985), 474-478. [36] D.L. Burkholder, Martingales and Fourier analysis in Banach spaces, Probability and Analysis (Varenna, 1985), Lecture Notes in Math. 1206, Springer, Berlin (1986), 61-108. [37] D.L. Burkholder, A proof ofPetczyhski's conjecture for the Haar system, Studia Math. 91 (1988), 79-83. [38] D.L. Burkholder, Sharp inequalities for martingales and stochastic integrals, Colloque Paul L6vy (Palaiseau, 1987), Ast6risque 157-158 (1988), 75-94. [39] D.L. Burkholder, Differential subordination of harmonic functions and martingales, Harmonic Analysis and Partial Differential Equations (El Escorial, 1987), Lecture Notes in Math. 1384, Springer, Berlin (1989), 1-23. [40] D.L. Burkholder, Explorations in martingale theory and its applications, Ecole d'l~t6 de Probabilit6s de Saint-Flour X I X - 1989, Lecture Notes in Math. 1464, Springer, Berlin (1991), 1-66. [41] D.L. Burkholder, Strong differential subordination and stochastic integration, Ann. Probab. 22 (1994), 995-1025. [42] D.L. Burkholder and R. E Gundy, Extrapolation and interpolation of quasi-linear operators on martingales, Acta Math. 124 (1970), 249-304. [43] D.L. Burkholder, B.J. Davis and R.E Gundy, Integral inequalities for convex functions of operators on martingales, Proc. Sixth Berkeley Symp. Math. Statist. Prob., Vol. 2 (1972), 223-240. [44] D.L. Burkholder, R.E Gundy and M.L. Silverstein, A maximal function characterization of the class H P, Trans. Amer. Math. Soc. 157 (1971), 137-153. [45] A.E Calder6n and A. Zygmund, On the existence of certain singular integrals, Acta Math. 88 (1952), 85-139. [46] A.E Calder6n and A. Zygmund, On singular integrals, Amer. J. Math. 78 (1956), 289-309. [47] S.D. Chatterji, Martingales of Banach-valued random variables, Bull. Amer. Math. Soc. 66 (1960), 395398. [48] S.D. Chatterji, Martingale convergence and the Radon-Nikodym theorem in Banach spaces, Math. Scand. 22 (1968), 21-41. [49] C. Choi, A submartingale inequality, Proc. Amer. Math. Soc. 124 (1996), 2549-2553. [50] C. Choi, A weak-type inequality for differentially subordinate harmonic functions, Trans. Amer. Math. Soc. 350 (1998), 2687-2696. [51] P. C16ment and B. de Pagter, Some remarks on the Banach space valued Hilbert transform, Indag. Math. (N.S.) 2 (1991), 453-460. [52] E Cobos, Some spaces in which martingale difference sequences are unconditional, Bull. Polish Acad. Sci. Math. 34 (1986), 695-703. [53] T. Coulhon and D. Lamberton, Rdgularitg L p pour les ~quations d'~volution, S6minaire d'Analyse Fonctionnelle, 1984/1985, Publ. Math. University Paris VII 26 (1986), 155-165. [54] G. David and J.L. Journ6, A boundedness criterion for generalized Calder6n-Zygmund operators, Ann. of Math. 120 (1984), 371-397. [55] W.J. Davis, D.J.H. Garling and N. Tomczak-Jaegermann, The complex convexity of quasi-normed linear spaces, J. Funct. Anal. 55 (1984), 110-150. [56] J. Diestel and J.J. Uhl, Vector Measures, Amer. Math. Soc., Providence, RI (1977).
Martingales and singular integrals in Banach spaces
267
[57] J.L. Doob, Regularity properties of certain families of chance variables, Trans. Amer. Math. Soc. 47 (1940), 455-486. [58] L.E. Dor and E. Odell, Monotone bases in Lp, Pacific J. Math. 60 (1975), 51-61. [59] G. Dore, L P regularity for abstract differential equations, Functional Analysis and Related Topics (Kyoto, 1991), Lecture Notes in Math. 1540, Springer, Berlin (1993). [60] G. Dore and A. Venni, On the closedness of the sum of two closed operators, Math. Z. 196 (1987), 189201. [61 ] G. Dore and A. Venni, Some results about complex powers of closed operators, J. Math. Anal. Appl. 149 (1990), 124-136. [62] G. Dore and A. Venni, Maximal regularity for parabolic initial-boundary value problems in Sobolev spaces, Math. Z. 208 (1991), 297-308. [63] I. Doust, Well-bounded and scalar-type spectral operators on L P spaces, J. London Math. Soc. 39 (1989), 525-534. [64] EN. Dowling, Representable operators and the analytic Radon-NikodSm property in Banach spaces, Proc. Royal Irish Acad. 85A (1985), 143-150. [65] EN. Dowling and G.A. Edgar, Some characterizations of the analytic Radon-NikodSm property in Banach spaces, J. Funct. Anal. 80 (1988), 349-357. [66] A. Driouich and O. E1-Mennaoui, On the inverse Laplace transform for Co-semigroups in UMD-spaces, Arch. Math. (Basel) 72 (1999), 56-63. [67] G.A. Edgar, Complex martingale convergence, Banach Spaces (Columbia, MO, 1984), Lecture Notes in Math. 1166, Springer, Berlin (1985), 38-59. [68] G.A. Edgar, Analytic martingale convergence, J. Funct. Anal. 69 (1986), 268-280. [69] G.A. Edgar and L. Sucheston, Stopping Times and Directed Processes, Cambridge University Press, Cambridge (1992). [70] L. Egghe, Stopping Time Techniques for Analysts and Probabilists, London Math. Soc. Lecture Notes Series 100, Cambridge University Press, Cambridge (1984). [71] E Enflo, Banach spaces which can be given an equivalent uniformly convex norm, Israel J. Math. 13 (1972), 281-288. [72] C. Fefferman and E.M. Stein, H p spaces ofseveral variables, Acta Math. 129 (1972), 137-193. [73] D.L. Fernandez, On Fourier multipliers of Banach lattice-valued functions, Rev. Roumaine Math. Pures Appl. 34 (1989), 635-642. [74] T. Figiel, On equivalence of some bases to the Haar system in spaces of vector-valued functions, Bull. Polon. Acad. Sci. Math. 36 (1988), 119-131. [75] T. Figiel, Singular integral operators: a martingale approach, Geometry of Banach Spaces (Strobl, Austria, 1989), London Math. Soc. Lecture Notes Series 158, Cambridge University Press, Cambridge (1990), 95-110. [76] D.J.H. Garling, Brownian motion and UMD-spaces, Probability and Banach Spaces (Zaragoza, 1985), Lecture Notes in Math. 1221, Springer, Berlin (1986), 36-49. [77] D.J.H. Garling, On martingales with values in a complex Banach space, Math. Proc. Cambridge Phil. Soc. 104 (1988), 399-406. [78] D.J.H. Garling, Random martingale transform inequalities, Probability in Banach Spaces 6, Birkhfiuser, Boston (1990), 101-119. [79] S. Geiss, Martingale transforms and applications to the Banach space theory, Dissertation, Jena (1995). [80] N. Ghoussoub and B. Maurey, Plurisubharmonic martingales and barriers in complex quasi-convex Banach spaces, Ann. Inst. Fourier 39 (1989), 1007-1060. [81] N. Ghoussoub, J. Lindenstrauss and B. Maurey, Analytic martingales and plurisubharmonic barriers in complex Banach spaces, Banach Space Theory, Contemp. Math., Vol. 85, Amer. Math. Soc., Providence, RI (1989), 111-130. [82] N. Ghoussoub, B. Maurey and W. Schachermayer, Pluriharmonicallly dentable complex Banach spaces, J. Reine Angew. Math. 402 (1989), 76-127. [83] Y. Giga and H. Sohr, Abstract L p estimates for the Cauchy problem with applications to the Navier-Stokes equations in exterior domains, J. Funct. Anal. 102 (1991), 72-94. [84] M. Giga, Y. Giga and H. Sohr, LP estimate for abstract linear parabolic equations, Proc. Japan Acad., Ser. A 67 (1991), 197-202.
268
D.L. Burkholder
[851 G. Godefroy, Renormings of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 781-835. [86] R.E Gundy, A decomposition of L 1-bounded martingales, Ann. Math. Statist. 39 (1968), 134-138. [87] J.A. Guti6rrez, On the boundedness of the Banach space-valued Hilbert transform, Dissertation, University of Texas, Austin, TX (1982). [88] W. Hensgen, Hardy-Riiume vektorwertiger Funktionen, Dissertation, Munich (1986). [891 W. Hensgen, On the dual space of H P (X), 1 < p < cx~, J. Funct. Anal. 92 (1990), 348-371. [901 P. Hitczenko, A remark on the paper "Martingale inequalities in rearrangement invariant function spaces" by W.B. Johnson and G. Schechtman, Functional Analysis (Austin, 1987-89), Lecture Notes in Math. 1470, Springer, Berlin (1991), 177-182. [911 A. Ionescu Tulcea and C. Ionescu Tulcea, Abstract ergodic theorems, Trans. Amer. Math. Soc. 107 (1963), 107-124. [921 T. Iwaniec, Extremal inequalities in Sobolev spaces and quasiconformal mappings, Z. Anal. Anwendungen 1 (6) (1982), 1-16. [931 T. Iwaniec and G. Martin, Riesz transforms and related singular integrals, J. Reine Angew. Math. 473 (1996), 25-57. [941 R.C. James, Super-reflexive Banach spaces, Can. J. Math. 24 (1972), 896-904. [951 W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84. [96] W.B. Johnson and G. Schechtman, Martingale inequalities in rearrangement invariant function spaces, Israel J. Math. 64 (1988), 267-275. [97] H. Kbnig, Vector-valued Lagrange interpolation and mean convergence of Hermite series, Functional Analysis (Essen, 1991), Lecture Notes in Pure and Appl. Math. 150, Dekker, New York (1994), 227-247. [981 H. K6nig and N.J. Nielsen, Vector-valued L p-convergence of orthogonal series and Lagrange interpolation, Forum Math. 6 (1994), 183-207. [99] K. Kunen and H. Rosenthal, Martingale proofs of some geometrical results in Banach space theory, Pacific J. Math. 100 (1982), 153-175. [ 1001 S. Kwapiefi, Isomorphic characterizations of inner product spaces by orthogonal series with vector valued coefficients, Studia Math. 44 (1972), 583-595. [101] C.W. Lamb, A short proof of the martingale convergence theorem, Proc. Amer. Math. Soc. 38 (1973), 215-217. [102] J.M. Lee, On Burkholder's biconvex-function characterization of Hilbert spaces, Proc. Amer. Math. Soc. 118 (1993), 555-559. [103] J.M. Lee, Biconcave-function characterisations of UMD and Hilbert spaces, Bull. Austral. Math. Soc. 47 (1993), 297-306. [104] O. Lehto, Remarks on the integrability of the derivatives of quasiconformal mappings, Ann. Acad. Sci. Fenn. Ser. A1 Mat. 371 (1965), 3-8. [105] J. Lindenstrauss and A. Petczyfiski, Contributions to the theory of the classical Banach spaces, J. Funct. Anal. 8 (1971), 225-249. [106] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I: Sequence Spaces, Springer, New York (1977). [107] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces II: Function Spaces, Springer, New York (1979). [108] J. Marcinkiewicz, Quelques thdorkmes sur les sdries orthogonales, Ann. Soc. Polon. Math. 16 (1937), 84-96. [1091 B. Maurey, Systkme de Haar, S6minaire Maurey-Schwartz, 1974-1975, l~cole Polytechnique, Paris (1975). [110] T.R. McConnell, On Fourier multiplier transformations of Banach-valued functions, Trans. Amer. Math. Soc. 285 (1984), 739-757. [111] T.R. McConnell, A Skorohod-like representation in infinite dimensions, Probability in Banach Spaces, V (Medford, 1984), Lecture Notes in Math. 1153, Springer, Berlin (1985), 359-368. [112] T.R. McConnell, Decoupling and stochastic integration in UMD Banach spaces, Probab. Math. Stat. 10 (1989), 283-295. [113] S. Monniaux and J. Priiss, A theorem of the Dore-Venni type for noncommuting operators, Trans. Amer. Math. Soc. 349 (1997), 4787-4814.
Martingales and singular integrals in Banach spaces
269
[114] A.M. Olevskii, Fourier series and Lebesgue functions, Uspehi Mat. Nauk 22 (1967), 237-239 (Russian). [115] A.M. Olevskii, Fourier Series with Respect to General Orthogonal Systems, Springer, New York (1975). [ll6] R.E.A.C. Paley, A remarkable series of orthogonal functions I, Proc. London Math. Soc. 34 (1932), 241264. [117] A. Pdczyfiski, Structural theory of Banach spaces and its interplay with analysis and probability, Proceedings of the International Congress of Mathematicians (Warsaw, 1983), PWN, Warsaw (1984), 237-269. [118] A. Petczyfiski, Norms of classical operators in function spaces, Colloque Laurent Schwartz, Ast6risque 131 (1985), 137-162. [119] A. Petczyfiski and H. Rosenthal, Localization techniques in LP spaces, Studia Math. 52 (1975), 263-289. [120] M. Piasecki, A geometrical characterization of A UMD Banach spaces via subharmonic functions, Demonstratio Math. 30 (1997), 641-654. [121] M. Piasecki, A characterization of complex AUMD Banach spaces via tangent martingales, Demonstratio Math. 30 (1997), 715-728. [122] A. Pietsch and J. Wenzel, Orthonormal Systems and Banach Space Geometry, Cambridge University Press, Cambridge, UK (1998). [123] G. Pisier, Un exemple concernant la super-rdflexivitd, S6minaire Maurey-Schwartz, 1974-75, l~cole Polytechnique, Paris (1975). [124] G. Pisier, Martingales with values in uniformly convex spaces, Israel J. Math. 20 (1975), 326-350. [125] A.O. Pittenger, Note on a square function inequality, Ann. Probab. 7 (1979), 907-908. [126] J. Prtiss, Evolutionary Integral Equations and Applications, Birkh~iuser, Basel (1993). [127] J. Prtiss and H. Sohr, On operators with bounded imaginary powers in Banach spaces, Math. Z. 203 (1990), 429-452. [128] M. Riesz, Sur les fonctions conjugudes, Math. Z. 27 (1927), 218-244. [129] J.L. Rubio de Francia, Fourier series and Hilbert transforms with values in UMD Banach spaces, Studia Math. 81 (1985), 95-105. [1301 J.L. Rubio de Francia, Martingale and integral transforms of Banach space valued functions, Probability and Banach Spaces (Zaragoza, 1985), Lecture Notes in Math. 1221, Springer, Berlin (1986), 195-222. [131] J.L. Rubio de Francia and J.L. Torrea, Some Banach techniques in vector valued Fourier analysis, Colloq. Math. 54 (1987), 273-284. [132] R. Ryan, The F. and M. Riesz theorem for vector measures, Indag. Math. 25 (1963), 408-412. [133] ES. Scalora, Abstract martingale convergence theorems, Pacific J. Math. 11 (1961), 347-374. [134] E.M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, Princeton (1970). [135] E.M. Stein, Topics in Harmonic Analysis Related to the Littlewood-Paley Theory, Princeton University Press, Princeton (1970). [136] E.M. Stein, Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals, Princeton University Press, Princeton (1993). [137] G. Wang, Differential subordination and strong differential subordination for continuous-time martingales and related sharp inequalities, Ann. Probab. 23 (1995), 522-551. [1381 J. Wenzel, Mean convergence of vector-valued Walsh series, Math. Nachr. 162 (1993), 117-124. [1391 J. Wenzel, Ideal norms associated with the UMD-property, Arch. Math. 69 (1997), 327-332. [1401 E Zimmermann, On vector-valued Fourier multiplier theorems, Studia Math. 93 (1989), 201-222. [1411 A. Zygmund, Trigonometric Series I, II, Cambridge University Press, Cambridge (1959).
This Page Intentionally Left Blank
CHAPTER 7
Approximation Properties Peter G. Casazza Department of Mathematics, The University of Missouri, Columbia, MO 65211, USA E-mail: pete @casazza, math. missouri, edu
Contents 1. The basis property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. The approximation property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. The bounded approximation property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. The commuting bounded approximation property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. The 7r-property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. The finite dimensional decomposition property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7. The uniform approximation property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. The compact approximation property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9. Approximation properties in non-separable spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
H A N D B O O K OF THE G E O M E T R Y OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 271
273 280 286 291 295 298 304 308 310 313
This Page Intentionally Left Blank
Approximation Properties
273
The approximation property, which already appeared in Banach's book in 1932 [5], plays a fundamental role in the structure theory of Banach spaces. The first systematic study of the variants of the approximation property was initiated by Grothendieck in 1955 [35]. At that time the main properties were the approximation property, the bounded approximation property and the basis property. It was not until 1972 that Enflo [23] produced the first example of a Banach space which fails the approximation property. Today, there is a "chain" of approximation properties in between the standard ones. Our goal is to introduce the important variants of the approximation properties and the relationships between them. In the process we will consider the various counter-examples and how they fit into the theory without giving the details of their constructions. We will not emphasize the applications of the approximation properties, but many of them can be found throughout the handbook. As we will see, there is a rich theory surrounding the approximation property for Banach spaces and the counter-examples merely illustrate the delicate nature of the relationships between the various properties. What one wants from this theory is to be able to prove that a given space has the weakest possible form of the approximation property and then be able to conclude that it actually has the strongest possible form. For example, we will see that a Banach space has the bounded approximation property if and only if it is a complemented subspace of a Banach space with a basis. But, there are Banach spaces with the bounded approximation property with fail to have bases. However, a complemented subspace of L p for 1 < p < cx~ (which immediately has the bounded approximation property) actually has a basis. Also, Lusky [69] has shown that the Disk Algebra A has a basis without using any deep results from complex function theory, but instead relies on results from the approximation property. Our emphasis will be on building the theory necessary to pass the various approximation properties from a space to its complemented subspaces, or from the dual space to the space etc. and finding the minimal conditions for passing from one of these properties to another. We will give only some representative proofs illustrating the basic techniques in the area. The counter-examples are quoted mostly to illustrate that these properties really are distinct in the general case, to show that positive results are best possible, and to explain some of the technical difficulties encountered in the proofs. These examples also make contact with other important areas such as Banach algebras, C*-algebras et al.
1. The basis property Basis theory is a large subject, and we will not attempt to cover even a fraction of it here. We will concentrate on those results which relate basis theory to the approximation properties. We refer the reader to the Basic Concepts section of this Handbook [47] for the basic definitions and results concerning bases. Here we use the term basis to mean Schauder basis. This concept was introduced by Schauder in 1927 [90]. But Schauder assumed his bases were normalized and that the dual functionals were continuous. Later, Banach [5] removed these unnecessary assumptions. The basis problem appeared in Banach's book [5] in 1932, p. 111: "On ne sait pas tout espace du type (B) separable admet une base?" But it was not until 1972 that Enflo answered this question in the negative [23]. Later, Szarek [95] constructed a Banach space X
P.G. Casazza
274
with a finite dimensional decomposition (see Section 6) which fails to have a basis. Also, Read [86] constructed a Banach space satisfying the commuting bounded approximation property which fails the :r-property (see Section 5). Both of these spaces are Banach spaces which fail to have bases despite being isomorphic to a complemented subspace of a Banach space with a basis, namely C2 (see Definition 1.10 and Proposition 6.7). Whenever possible we will refer to the space of Szarek since Read has not submitted his manuscript for publication. Although every separable Banach space embeds into a Banach space with a basis, namely C[0, 1], there are separable Banach spaces which never embed complementably in a space with a basis. One class of such spaces are the Banach spaces which fail the approximation property discussed in Section 2. There is a simple criterion due to Petczyfiski and Wojtaszczyk [82] (later improved independently by Petczyfiski [80] and Johnson, Rosenthal and Zippin [50]) for checking when a Banach space is complemented in a space with a basis. To state this result, we need a definition.
DEFINITION 1.1. A sequence of non-zero finite rank operators (Ai) from a Banach space X to itself is called a (unconditional)finite dimensional expansion of the identity if for all xEX, x -- Z
Aix, i
(and the series converges unconditionally). It is immediate that a finite dimensional expansion of the identity is the same as the bounded approximation property (see Section 3). Requiring the operators to be non-zero in Definition 1.1 is a convenience but not a necessity. We now have
PROPOSITION 1.2. A Banach space X is isomorphic to a complemented subspace o f a Banach space with a basis if and only if X has a finite dimensional expansion o f the identity. PROOF. (=~) If Y is a Banach space with a basis (Xi,X*) and X is a complemented subspace of Y with projection P : Y -+ X, let I -- {i: Pxi 7/= 0}. Now, the operators A i x -- x i*( x ) P x i , for i 6 I, form a finite dimensional expansion of the identity for X (,=) Here we will just do the case where the operators Ai are rank 1 and in Section 6 we will do the general case. We let Y be the family of sequences (xi), xi ~ A i ( X ) for which Y~i xi converges in X. We define addition and multiplication coordinatewise in Y and let the norm be given by ]l(xi)ll- supn II ~in___lxillx. Then Y is a Banach space with a basis (Yn) consisting of any choice of Yn = (xi) with xi = 0 for all i ~ n, and Xn :7/=O. Finally, Ux = (Anx) is an isomorphic embedding of X into Y and P(xi) = (An(Y'~i xi)) is a bounded linear projection of Y onto X. [5] Similarly, one can show that a Banach space X is isomorphic to a complemented subspace of a Banach space with a unconditional basis if and only if X has a unconditional 1-dimensional expansion of the identity. It is known [50] that if X* has a basis, then so does X. In fact, in this case X has a shrinking basis and so X* has a boundedly complete basis. To see that the converse of
Approximation Properties
275
this fails, we need a basic tool for constructing examples which is a generalization due to Lindenstrauss [63] of a result of James [38]. PROPOSITION 1.3. If X is a separable Banach space, there is a separable Banach space Z with Z** having a boundedly complete basis and Z * * / Z is isomorphic to X. Moreover, Z*** ~- Z* 9 X*. PROOF. Let (xn) be a sequence of which is dense in the boundary of the unit ball of X. The space Z consists of all the sequences of scalars z - - (ai) for w h i c h Y~i aixi = 0 and Pj
j=l
Z +1aixi
2) 1/2 <~,
(1)
i=pj-i
where the supremum is taken over all choices of integers m and 0 = P0 < pl < " " < Pm. It follows that for every z = (ai) E Z, the s e r i e s ~-~i aixi converges in X. The main point of the proof is to show that Z** can be identified with the space of all sequences for which (1) holds. Then the operator T: Z** --+ X given by T (ai) = ~--~iaixi is bounded, and by the density of (xi) in the boundary of the unit ball of X, it is a quotient map. Moreover, the unit vectors of Z form a boundedly complete basis of Z. The moreover part of the theorem comes from the fact that for every Banach space Z there is a norm 1 projection from Z*** onto Z*. The projection is just the map which takes every bounded linear functional on Z** to its restriction to Z. D Now it is simple to construct a Banach space with a basis whose dual fails to have a basis (even the approximation property). PROPOSITION 1.4. There is a Banach space X with a basis so that X* is separable and fails the approximation property.
PROOF. We take a separable reflexive Banach space X failing the approximation property (see Section 2) and choose Z with Z** having a basis and Z*** ~ Z* @ X*. Then Z*** fails to have a basis - e v e n the approximation property. D Johnson and Schechtman (see [48]) have shown that there are subspaces of co with bases whose dual spaces fail to have bases (or even the approximation property). Also, there are subspaces of g l which have bases whose dual spaces fail the approximation property (see Corollary 6.12). Obviously, any subspace of a Hilbert space (being itself a Hilbert space) has a basis. This property does not characterize Hilbert spaces but spaces having this property must be very close to Hilbert spaces. Kwapiefi's theorem [59] states that any Banach space which is both type 2 and cotype 2 must be isomorphic to a Hilbert space. But, Szankowski [91] shows that if every subspace of a Banach space X has the approximation property then for every e > 0, X has type ( 2 - e) and cotype (2 + e). Johnson [44] showed there are nonHilbert spaces with the property that every subspace (even every subspace of every quotient
276
P. G. C a s a z z a
space) has a basis. The first such example was the convexified Tsirelson space T 2. In 1972, at the same time that Enflo gave the first counter-example to the approximation problem, Tsirelson [97] constructed the first Banach space which contains no subspaces isomorphic to co, or g p, for 1 ~< p < o~. Figiel and Johnson [28] gave an analytic description of the norm on the dual space of the space constructed by Tsirelson, and this soon became known as Tsirelson's space T. The p-convexification of T was also constructed by Figiel and Johnson [28]. To define these spaces we fix 1 ~< p < oo, and let (tn) be the unit vectors in the linear space of finitely non-zero sequences Coo. For two subsets E, F of the natural numbers, we will write E < F to mean max E < min F. Also, for E C N and x - - ~-]m an tn we write E x =: Y~n~E antn. For x = Y~n antn, we inductively define a sequence of norms (ll" IIm)mC~_-0 on Coo by
IIx Iio : max lan I, /7
and for all m ~> 0,
Ilxllm+l - max
IIEjxll
IlXllm, ~ max
,
(2)
j:l
where the inner max is taken over all choices of finite subsets k <~ E1 < E2 < ... E~ of N. Since (llx [Im )m= oo 1 is increasing we can let
IIxll -
lim IIxllm
m---> oo
for all x for which this limit is finite. Now, p-convexified Tsirelson space TP is the completion of Coo in this norm. A parallel theory developed for modified Tsirelson's spaces where we assume the (Ei) are just disjoint subsets of N with k <<,Ei. However, Casazza and Odell [17] showed that these two constructions produce equivalent norms. The unit vectors (tn) form a 1-unconditional basis for T p and none of these spaces contains a subspace isomorphic to co or gq, 1 ~< q < oo. Johnson [44] showed that every subspace of every quotient space of (a subspace of) T 2 has a basis. When stronger properties of these spaces were developed [18], it was discovered that the subspace constructed by Johnson was isomorphic to the whole space T 2. It follows easily from the definition that any n disjointly supported normalized blocks (yi)in=l o f (tk)Ln are 2-equivalent to the unit vector basis of ~2. Duality arguments show that this property holds also in T*, which also yields norm 2 projections onto the blocks spanning ~ in T 2. Casazza, Johnson and Tzafriri [12] showed that we can iterate this process (see also [18,85]) to exponentially increase the number of blocks one can take supported after n and still have a good copy of the unit vector basis in a Hilbert space. Now, applying Proposition 7.3, we can show that every n-dimensional subspace E of span(tk)~=n is 8-complemented and 8-isomorphic to ~ . Now, if E is any 2n-dimensional subspace of T 2, and F -- E N span(tk)~__n+l then dim F ~> n. It follows that every 2n-dimensional subspace of T 2 contains an n-dimensional subspace which is 16-isomorphic to a Hilbert space and 16-complemented in T 2. Pisier
277
Approximation Properties
[84] introduced the notion of a weak Hilbert space and gave several equivalent formulations for the definition. One of them is that X is a weak Hilbert space if there is a constant M > 0 so that every 2n-dimensional subspace of X contains an n-dimensional subspace which is M-complemented in X and M-isomorphic to s So by the discussion above, T 2 is a weak Hilbert space. Nielsen and Tomczak-Jaegermann [78] have shown that every separable weak Hilbert space which is a Banach lattice is very much like T 2. In particular, every subspace of every quotient space of these spaces has a basis. Pisier [84] showed that all weak Hilbert spaces have the approximation property (see Section 7 for a discussion of weak Hilbert spaces and the uniform approximation property). It is an open question whether every separable weak Hilbert space has a basis. But Maurey and Pisier (see [74]) have shown that every separable weak Hilbert space has a finite dimensional decomposition (see Section 6). The converse is also an open question. That is, if X is a separable Banach space and every subspace of X has a basis, must X be a weak Hilbert space? There is some strong evidence towards a positive answer to this question given by Mankiewicz and Tomczak-Jaegermann [72,73]. They showed that a Banach space X for which every subspace of X and every quotient space of X has a basis (with uniformly bounded basis constants) is a weak Hilbert space. The results in [73] are much stronger than we have stated and actually are finite dimensional in nature and give quantitative estimates of the weak type 2 and weak cotype 2 constants in terms of the uniform basis constants for all subspaces (respectively quotient spaces) of X. Although there are infinite dimensional spaces which are never complemented in spaces with bases, the finite dimensional situation is much better. This is a result of Petczyfiski [80] (proved independently without the exact constants in [50]). PROPOSITION 1.5. For every finite dimensional space E and every e > O, there is a finite dimensional space F D E, with a norm one projection P from F onto E, so that F has a basis with basis constant <~ (1 + e). n forE PROOF. Let dim E -- n and choose an Auerbach system ( X i , X i*)i=l rn + j, 0 <~ r < n, 1 <<.j <<.n, put Ui(x) -- -ff1X* j(x)xj such that yi E Ui (E). Define a norm on F by
IlYll =
max
l<~k<~n 2
For e a c h i - -
Let F be all sequences y -- (yi)/.n21
~Yi
i=1
I f w e choose ei E U i ( E ) for 1 ~
1
I IE, IIg-~ll ~< 1 and IIv II ~ 1 + n2.
E by U ( y i ) -- ~ yi, we see F]
The situation for unconditional bases is much more delicate. Although every Banach space contains a subspace with a basis, Gowers and Maurey [33] have shown that they need not contain unconditional basic sequences. There are classical Banach spaces, such as C[0, 1] and L 110, 1] which do not embed into any Banach space with a unconditional
278
P.G. Casazza
basis. Also, Lindenstrauss [60] observed that el has a subspace D with a basis given by the vectors (en - 1 (ezn -k- e2n+l)), while D fails to have a unconditional basis. It is immediate that if X has a unconditional basis and X* is just separable, then el does not embed into X and so the unconditional basis for X is shrinking and hence its dual functionals form an unconditional basis for X*. Also, any pre-dual X of el without an unconditional basis (such as C (co'~ provides an example of a Banach space which fails to have a unconditional basis but its dual space has a unconditional basis. Proposition 1.2 becomes more delicate in this setting also. PROPOSITION 1.6. I f X is complemented in a Banach space with a unconditional basis, then X has a unconditional finite dimensional expansion o f the identity. The converse fails. However, if X has a unconditional finite dimensional expansion of the identity by rank 1 operators, then X is complemented in a Banach space with a unconditional basis. PROOF. The first part of Proposition 1.5 is exactly the same as the first part of the proof of Proposition 1.2. The last statement in Proposition 1.5 is also done exactly as in the proof of Proposition 1.2 with the slight modification in the definition of the space Y where we require the series Y~i xi there to converge unconditionally. For the converse, Kalton and Peck [54] have constructed a Banach space Z which fails to have a unconditional basis but is an unconditional sum of two dimensional Banach spaces. That is, there is a sequence En of subspaces of Z with dim En = 2, for all n = 1, 2 . . . . and for every z E Z there is a unique sequence of vectors Zn E En with z -- Y~n Zn and this series converges unconditionally. It follows that the operators An z = Zn give an unconditional expansion of the identity by rank 2 operators while it is known [46] that Z is not complemented in any Banach space with a unconditional basis. [] There is a partial converse for Proposition 1.6 stating that a Banach space with a unconditional finite dimensional expansion of the identity is isomorphic to a subspace of a Banach space with a unconditional basis (see Section 6). The different results we are getting for complemented subspaces of spaces with bases and spaces with unconditional bases can be explained by some fundamental results of Gordon and Lewis [32]. We say that a Banach space X has GL-LUST (Gordon-Lewis local unconditional structure) if there is a constant K > 0 so that for every finite dimensional subspace E of X there is a Banach space F with a 1-unconditional basis and operators T : E --+ F and S : F --+ X with S T = liE and IISIIIITll ~< g . It is easily checked that a complemented subspace of a Banach space with a unconditional basis must have GLLUST. Gordon and Lewis [32] and a host of other authors explored some deep consequences of a space having GL-LUST. By Proposition 1.5, every finite dimensional space is 1-complemented in a finite dimensional space with a basis whose basis constant is at most 2. One consequence of the work on GL-LUST is that if we fix K, we cannot find for every finite dimensional space E a finite dimensional space F so that E O1 F has a K-unconditional basis. The picture for complemented subspaces of Banach spaces with unconditional bases is quite unsettled at this time. The one positive result here is due to Kalton and Wood [55].
Approximation Properties
279
THEOREM 1.7. If X is a complex Banach space with a 1-unconditional basis and Y is a 1-complemented subspace of X, then Y has a 1-unconditional basis. Rosenthal [87] and Flinn [29] have given separate proofs of this result. Benyamini, Flinn and Lewis [6] have constructed a (real) Banach space without a 1-unconditional basis which is 1-complemented in a space with a 1-unconditional basis. But, their complemented subspace does have a unconditional basis. So the general question is still open. PROBLEM 1.8. Does every complemented subspace of a Banach space with a unconditional basis have a unconditional basis? This is the atomic version of the general question: Is a complemented subspace of a Banach lattice isomorphic to a Banach lattice? More than likely this question has a negative answer. A weaker conclusion might have a chance of holding. PROBLEM 1.9. If Y is complemented in a Banach space with a unconditional basis, does Y have a complemented subspace with a unconditional basis? Special cases of these problems are also of interest. For example, if Y is a complemented subspace of Lp[O, 1], 1 < p < cx~, does Y have a unconditional basis? Complemented subspaces of Lp which are not isomorphic to a Hilbert space are called s and it is known [50] that separable s have bases. In fact [79], s have bases (xi) so that supnd(span(xi)in=j, g~) < oo. It is also known [15] that if gp r X has an unconditional basis, then X has an unconditional basis. One way to give a positive answer to the s would be to answer the corresponding question for Lp[O, 1]. If Lp 9 X has a unconditional basis, must X have an unconditional basis? This would answer the s question since it is known [4] that if X is complemented in Lp, then L p "~ L p 9 X. Although every subspace of T 2 has a basis, this space has subspaces without unconditional bases. This comes from the GL-LUST considerations of Komorowski [53]. We ask PROBLEM 1.10. If every subspace of a Banach space X has a unconditional basis (or just GL-LUST), must X be isomorphic to a Hilbert space? There is evidence to suggest that Problem 1.10 has a positive answer [53,57]. We end this section with a discussion of a class of universal spaces with bases constructed by Johnson [42]. We choose a sequence of finite dimensional spaces (En) which are dense in the Banach-Mazur distance in the family of all finite dimensional spaces. Also, let (Fn) be dense in the family of finite dimensional spaces with a 1-unconditional basis. DEFINITION 1.1 1. For 1 ~< p < cx~ let
n
~p
n
~p
and for p - 0 , the sums above are replaced by c0-sums.
P.G. Casazza
280
We note for later reference that C p has the property that for every finite dimensional space G and for every e > 0 and for every natural number m, there is an n > m and an operator T: G --+ En such that II T II II T -1 II ~< 1 + e. To see that we can assume this l property, let Cp = (Y~n @Gn)ep, where (Gn) is a listing of the (En) where each En appears ! ! infinitely many times in the sequence (Gn). Then clearly Cp is isometric to ( ~ E3Cp)ep. Note that C p is isometric to a subspace of C p' . Also, given any Ek and any m, the space (Y~in_ 1 ~)Ei)~pEDpEkis (1 -+- e)-isomorphic to one of the Ei's and we must have i > m just ! by dimension considerations. It follows that C p is isometric to a complemented subspace of Cp and hence by the Petczyfiski decomposition method Cp ~ C'p" An application of Proposition 1.5 shows that C p has a basis. The simplest way to do this is to observe that we can rearrange the En so that each Ezn G E2n+! has a basis (x~)~" l
~,~,~t'xn~kni)i:l)n~
with basis constant <~ 2. Then l__ is a basis for (a space isometric to) Cp with basis constant ~< 2. Johnson [42] showed the universality of C! and U1. PROPOSITION 1.12. If X is a separable Banach space, then X* is isometric to a norm 1
complemented subspace of C~. If X is separable and has GL-LUST then X* is isometric to a norm 1 complemented subspace of U~. PROOF. Pick a sequence G l C G2 C ... of finite dimensional subspaces of X whose union is dense in X. We can find natural numbers kl < k2 < ... and operators Tn : Gn --+ Ek, with II Zn II II T~- 1 II <~ 1 + 1. Define T" C1 --+ X by T (Xn) = Y~n Tn 1xk,. Then T is a quotient n map, and so T* is an isometry. Also, T* T-1 is a projection onto the range of T*. The other case is similar. D
2. The approximation property If a Banach space X has a basis (Xn)n~__l with basis projections Pn, then for any Banach space Y and any compact operator T E B(X, Y), l i T - PnTll ~ O. The question o f w h e t h e r the converse was true (known as the approximation problem) was an open problem for quite some time until until Enflo [23] gave a counter-example in 1972. At the end of this section we will discuss the various examples concerning the approximation property. NOTATION 2.1. For Banach spaces X, Y, we denote by U ( X , Y), the family of finite rank operators T : X --+ Y, and let U ( X ) = f ' ( X , X). DEFINITION 2.2. A Banach space X is said to have the approximation property (AP for short) if for every compact set K in X and every e > 0, there is a T 6 9t'(X) so that IIT x - x II ~< e, for every x E K. Every Banach space with a basis has the approximation property. To see this, we just observe that if K is a compact set in X, then l i m n ~ I1(I - en)lK II = 0, where (Pn) are the basis projections. In this proof, the finite rank operators approximating the identity on K are actually projections. It is an open question whether the approximation propertyis equivalent to this "projection form" of the approximation property. A simple calculation
Approximation Properties
281
shows that every complemented subspace of a space with the approximation property must have the approximation property. Grothendieck [35] initiated the study of the variants of the approximation property and the relations between them. One important tool he used was the topology of uniform convergence on compacts sets. NOTATION 2.3. If X and Y are Banach spaces, we let r denote the topology on B(X, Y) of uniform convergence on compact sets in X. The topology r is the locally convex topology on B(X, Y) generated by the seminorms of the form IITII~c - s u p { l l T x l l : x ~ K } , where K ranges over the compact subsets of X. We write (B(X, Y), r) for this space. Grothendieck [35] identified the dual space of (B(X, Y), r). PROPOSITION 2.4. The continuous linear functionals on (B(X, Y), r) consist of all func-
tionals 4) of the form oo
0(3
~ ( r ) -- Z y * ( r x i )
'
xi 9 X
'
y* 9 X* , Z
Ilxi IIIlyi* II < ec.
l
i=1
i=1
PROOF. If 4) has such a representation, assume 0 r Xi and choose Oti > 0 tending to ec so that ~-~i Oli Ilxi III Y* II - M < ec. Then K -- xi /Ilxi Iloti U {0}, is compact and
I (r i
liy*lillx ll llW(xi/llxiil e)ll
<<.
MilrgiK.
i
Conversely, if 4) is a linear functional on B(X, Y) satisfying ]4~(T)] ~< MIITIIx, with K compact, a well known construction (see, e.g., [66, Proposition 1.e.2]) shows that we may assume that K = c--o-~(xi) where ]]xi ]]--+ 0. Let S: B(X, Y) --+ (Y | Y G . . ")co = Z be given by S(T) = ( T x l , Tx2 . . . . ). Since ]4~(T)I ~< MIIS(T)I], there is a linear functional on S B ( X , Y) so that 4~(T) = ~ ( S T ) . By the Hahn-Banach theorem, we can extend this to a linear functional on Z, and hence to an element of (Y* G Y* G . . ")e~. That is, there must exist Yi 9 Y* so that Z i IlYi II < oc and ~ ( T ) -- Y~i Yi T ( x i ) . I--I The next result due to Grothendieck [35] relates the approximation property to the question of approximating compact operators. THEOREM 2.5. For a Banach space X, the following statements are equivalent:
(1) X has the approximation property. (2) There is a net (Tk))~A of finite rank operators on X converging to the identity op-
erator on X uniformly on compacta, i.e. T)~ --+ I in (B(X), r). (3) For every Banach space Y, .F(Y, X) is dense in (B(Y, X), r). (4) For every Banach space Y, U ( X , Y) is dense in (B(X, Y), r). (5) For every choice of (xi) in X and (x[) in X* such that ~ i []xi I] ]ix*l] < oc and
Y~i x*(x)xi - O, for all x 9 X, we have Y~.i x*(xi) - O. (6) For every Banach space Y, every compact operator T 9 B(X, Y) and every e > O, there is a T 9 .F(X, Y) with IIT - T1 II < 8.
282
P.G. Casazza
PROOF. The equivalence of (1) and (2) is immediate from the definitions. The equivalence of (1) and (5) is a consequence of Proposition 2.4. (1) means that the identity operator is in the r-closure of U(X) in (B(X, Y), r). This happens if and only if every r-continuous linear functional which vanishes on the operators of rank 1, also vanishes on the identity operator. But, in light of Proposition 2.4, this is exactly what (5) is saying. Letting X = Y in (3) and (4) shows that they both imply (1). Now assume (1). Let T 9 L(Y, X). For every compact set K in Y, T (K) is compact in X. Hence, given e > 0, by (1) there is a T1 9 9t-(X) so that IIT~Ty - Tyll <<,e, for all y 9 K. Since Tl T 9 9t'(Y, X), this proves (3). A similar construction proves that (1) implies (4). The equivalence of (1) and (6) can be found in [35,66]. [] Theorem 2.5 has several important consequences. First, if X fails the approximation property, letting Y = span(x/) in (5) of Theorem 2.5, we see that for every subspace Y C Z C X, Z fails the approximation property. That is, if X fails the approximation property, then there is a separable subspace Y C X so that for every subspace Y C Z C X, Z fails the approximation property. Another immediate consequence of the equivalence of (1) and (5) in Theorem 2.5 is PROPOSITION 2.6. If X* has the approximation property, then so does X. Hence, if X is reflexive, then X has the approximation property if and only if X* has the approximation property. Condition (6) of Theorem 2.5 raises an obvious question, which is still unsolved. Namely, is there an internal characterization of the approximation property in terms of approximating compact operators by finite rank operators on X ? Formally, we are asking. PROBLEM 2.7. Let X be a Banach space with the property that for every e > 0 and for every compact operator T 9 B(X), there is a T1 9 9t-(X) with liT - T1 II ~< E. Must X have the approximation property? Given the symmetry of (3) and (4) in Theorem 2.5, it is natural to ask what happens if we reverse the rolls of X and Y in Condition (6). The answer is another result of Grothendieck
[35]. PROPOSITION 2.8. l f X is a Banach space, then X* has the approximationproperty if and only if for every Banach space Y, every e > 0 and every compact operator T 9 B(X, Y), there is a T1 9 ~ ( X , Y) such that IIT - T111 ~< e. During his investigation of the approximation problem Grothendieck found many interesting equivalent formulations of the problem. Although we now have counter-examples to the general problem, Grothendieck's results still have interest since they show how to transfer these examples to some examples in classical analysis. PROPOSITION 2.9. The following three assertions are equivalent:
Approximation Properties (1) Every Banach space has the approximation property. (2) Every matrix A - - ( a i j ) i,~ j= l of scalars, f o r which limi
283
a i j - - O,
and }-~i maxj laij l < 0(3 and A2 __ O, also satisfies trace A - - Z i (3) Every continuous function K ( r , s ) on [0, 1)| [0, 1],forwhich
f o r all i - 1,2 . . . . . a i i - - O.
fo K ( r , s ) K ( s , t ) d s - - O ,
for every r and t, satisfies fd K (s, s) ds - 0. We now pass to a discussion of the various counter-examples to the approximation property and some of their consequences. After Enflo [23] constructed the first Banach space failing the approximation property, this was quickly followed by a number of important examples concerning the approximation property. Almost immediately, Figiel [25] and Davie [20,21 ] gave greatly simplified constructions for spaces failing the approximation property and showing that there are subspaces of g p, 2 < p, failing the approximation property. Szankowski [92] then constructed subspaces of g p, 1 ~< p < 2, which fail the approximation property. We will start, however, at the end of the story, since it puts these results into proper context. There is a general criterion for a space to fail the approximation property. This is a modified version of a criterion used by Enflo in his original solution to the approximation problem. This criterion actually identifies spaces which fail the weaker property the compact approximation property (compact AP for short) studied in Section 8. A Banach space is said to have the compact approximation property if it satisfies Definition 2.2 with the finite rank operators there replaced by compact operators. PROPOSITION 2.10. Let X be a Banach space and assume there are sequences ( x i ) in
X and (x*) in X*, a sequence (Fi) offinite subsets of X and an increasing sequence of integers (ki ), so that the following hold: (1) x i (xi ) -- 1, f o r every i -- 1, 2 . . . . . (2) sup/Ilxi II < ~ . :r (3) x i --+ 0 in the weak-star topology. (4) For every T E B ( X , Y), if we let rio(T) = 0 and f o r n ~ 1 1
kn
fin(T)-- kn ) ix;(Txi)' then f o r every n = 1,2 . . . . we have I/~n(T)-/~n-l (T)I ~< sup{llTxll: x E Fn}. (5) E n sup{ IIx I1: X E F~ } < ~ . Then X fails to have the compact approximation property (and hence the approximation property).
284
P. G. Casazza
Although this criterion looks formidable, in practice it allows the construction of spaces designed specifically to satisfy this criterion to produce counter-examples to the approximation property. In an important series of constructions, Szankowski carried out this program to produce a spectacular collection of counter-examples to the approximation property. The first was the construction of a Banach lattice which fails the approximation property [91 ]. PROPOSITION 2.11. For every 1 <<, r < p < e~, there is a sublattice o f (Lr[0, 1] 9 Lr [0, 1] 0 " " ")ep which fails to have the compact approximation property. The lattice Lr[0, 1] itself cannot have any sublattices which fail the approximation property since its sublattices are all order isometric to L p ( v ) , for a suitable measure v. Szankowski's next construction [92] (see [66, Theorem 1.g.2, p. 103]) was of subspaces of p, p r 2, which fail the approximation property. The real significance of this construction is the surprisingly exact representation of the set of vectors in e p which span a subspace failing the approximation property. PROPOSITION 2.12. For every 1 <~ p :/= 2 < cx~, there is a subspace o f g,p which fails the compact approximation property. PROOF. We will work with the case 1 ~< p < 2. We let X be the space of all sequences x -- ( a i ) ~ 4 so that
Ilxll =
Z =-
lail2
BEAn
where An is a carefully chosen partition of {2n, 2 n -+- 1 . . . . . 2 n+l - 1}. Since X is an gpsum of e~'s, it follows that X "~ ep. We let (ei)~xz=4be the unit vector basis of this space and let Z be the closed subspace of X spanned by the sequence Zi =
e2i
--
e2i+l -+- e 4 i
-+-
e4i+l + e4i+2 + e4i+3.
Then Z fails to have the compact approximation property. Let us recall a result of Krivine [58] and Maurey and Pisier [76]. For a Banach space X, if p(X) _ sup{p: X is of type p},
q (x) _ inf{q" X is of cotype q },
then X contains almost isometric copies of s p ( X ) and s q ( X ) , for n - 1 , 2,
....
Because of the
structure of the proof of Proposition 2.12, we can see that any Banach space which contains subspaces uniformly isomorphic to g np . for. n .- 1. 2 .. , will have such a construction on it. This observation combined with the preceding discussion yields a major theorem in this area.
Approximation Properties
285
THEOREM 2.13. If X is a Banach space and every subspace of X has the compact approximation property, then X is of type 2 - E and cotype 2 + e, for every e > O. For later reference, we mention that because of the structure of compact sets in sums of Banach spaces, the approximation property passes through gp-sums. PROPOSITION 2.14. If (Xn) is a sequence of Banach spaces with the approximation property, then for every 1 <~ p < cx~ the space (Y-~,I | (and (~-~n | has the approximation property. Now we consider other counter-examples concerning the approximation property. Since there are reflexive spaces (even subspaces of g p, p r 2) which fail the compact approximation property, using the construction of Section 1, following Proposition 1.3, we see that there is a Banach space Z so that Z** has a boundedly complete basis while Z*** fails to have the compact approximation property. It follows that the dual space of the universal space Cl (which has a basis - see Definition 1.10) fails to have the compact approximation property. We also have the construction of Johnson and Schechtman (see [48]) of a subspace of co with a basis whose dual space fails to have the approximation property. Johnson [44] has constructed spaces of the form X - ( ~ n ~_ ~k,, p,,)e2, for carefully chosen Pn $ 2 and kn t c~ so that every subspace of X has the approximation property, but X has subspaces without bases. Casazza, Garcfa and Johnson [10] combined Enflo's construction [23] with Johnson's construction just cited to produce Banach spaces X, Y so that every subspace of every quotient space of both X, Y have the approximation property (even a finite dimensional decomposition) but X @ Y has a subspace failing the compact approximation property. Since there are Banach spaces with bases whose duals fail the approximation property, it follows from Theorem 2.5 and Proposition 2.8 that there exists Banach spaces X, Y and a compact operator T e B ( X , Y) so that T is a r-limit of finite rank operators but T is not a norm limit of finite rank operators. Figiel (unpublished) observed that if X is reflexive and Y is either separable or reflexive, then any compact operator T E B ( X , Y) which is a r-limit of finite rank operators is also a norm limit of finite rank operators. Since there is a Banach lattice which fails the compact approximation property, it follows that U~ fails the approximation property by Proposition 1.12. In particular, there is a Banach space with a 1-unconditional basis whose dual space fails the approximation property. Szankowski [93], in yet another spectacular construction, resolved the much worked on question of the approximation property for B (H). THEOREM 2.15. For infinite dimensional Hilbert spaces H, B ( H ) fails the approximation property. This was the first naturally occurring example shown to fail the approximation property. Godefroy and Saphar [31] observed that B ( H ) / K ( H ) fails the approximation property. It is still unknown at this time if H ~ has the approximation property. In [83,84], Pisier gave a solution to a problem of Grothendieck related to the approximation property. Grothendieck raised the question: If X and Y are Banach spaces for which the injective and projective tensor products coincide (i.e., X + Y - X ~ Y), must either X
286
P G. Casazza
or Y be finite dimensional? Pisier [83,84] constructs an infinite dimensional Banach space X so that X + X -- X ~ X, both algebraically and topologically. Such a space must fail the approximation property. THEOREM 2.16. Any Banach space Y of cotype 2 can be embedded isometrically into a Banach space X of cotype 2 such that (1) X + X - X ~ X . (2) X and X* are both of cotype 2. (3) Every operator from X into X which is in the norm closure of the finite rank operators is nuclear. Moreover, if Y is separable then the same is true for X. It is known [84] that if X has the approximation property and both X and X* are of cotype 2, then X is isomorphic to a Hilbert space. Pisier then used this construction to answer a question of Lindenstrauss" Does every separable Banach space contain uniformly complemented subspaces uniformly isomorphic to g~, ~ or g ~ ?
COROLLARY 2.17. There is an infinite-dimensional Banach space X and a number 6 > 0 such that every finite rank projection P on X satisfies
II P [I >~ 6 (rank P) 1/2. We end this section with a brief discussion concerning the impact of counter-examples to the approximation problem on the study of Banach algebras. A left approximate identity in a Banach algebra A is a net (e~)~eA of elements of A such that limx Ilexx - x II - 0, for all x E A. One consequence of Theorem 2.5 (2) is that if a Banach space X has the approximation property, then f ( X ) (the closure of the finite rank operators in B ( X ) ) has a left approximate identity. The converse is an open question related to Problem 2.7. The other variants of the approximation properties we will examine in later sections also have formulations in terms of (bounded) approximate identities. We will not examine this topic in detail but refer the reader to [11,22,34,89] or books on Banach algebras for the current results on this subject. There has been little attention paid to the approximation property for Banach lattices and whether there are stronger results in this setting. We refer to Nielsen [77] for the most recent results on this subject. We ask PROBLEM 2.18. For a Banach lattice X, does the approximation property imply the positive approximation property? Does the bounded approximation property imply the positive bounded approximation property (see Section 3)?
3. The bounded approximation property The definition of the approximation property (Definition 2.2) puts no restriction on the norm of the finite rank operator which approximates the identity on a compact set. Yet, as
Approximation Properties
287
we saw in the case of X having a basis, the basis projections are providing the required approximation and this set of projections is uniformly bounded (independent of the compact set K). Now we introduce this stronger form of the approximation property. DEFINITION 3.1. Let X be a Banach space and 1 ~< )~ < cx~. We say that X has the h - b o u n d e d approximation property (X-BAP for short) if for every e > 0 and every compact set K in X, there is a finite rank operator T (i.e. T E f ( X ) ) so that IITll ~< )~ and IIT x - x II ~< e for every x E K. We say that X has the b o u n d e d approximation property (BAP for short) if X has the )~-BAP, for some ~. Finally, a Banach space is said to have the metric approximation property (MAP for short) if it has the 1-BAP. The compact sets in Definition 3.1 can be replaced by finite sets. To see this, let K be a compact set in a Banach space X and let e > 0 be given. We can find a finite set (Xi)i__ n 1 SO that K C U i ~ l B ( x i , e/3)~). Now, if T is a finite rank operator with IlTxi - xill <<.e / 3 , for 1 ~< i <~ n, then II Z x - x II ~< e, for every x E K. It is immediate that a space with a basis has the bounded approximation property and that a space with the bounded approximation property has the approximation property. The converse of both of these implications is false. For the first implication, we observe that if X has the X-BAP and if P is a bounded linear projection from X onto Y then Y has the X II P IIBAP. Now let X be a reflexive Banach space which is complemented in a space with a basis while X itself fails to have a basis (see Proposition 6.7). Then X has the bounded approximation property but fails to have a basis. A counter-example to the converse of the second implication will be done later in this section. As we have seen in Section 2, a space with a basis has the bounded approximation property and the same argument shows that a space with a monotone basis has the metric approximation property. The converse of these results fails as we will see later in this section. All/2p-spaces have the bounded approximation property. If ( X , ) is a sequence of Banach spaces, let X = (y~ | (or a c0-sum). Then it is immediate that X has the bounded approximation property if and only if the Xn all have the bounded approximation property with a uniform bound on their BAP constants. We now show that finite rank operators approximating the identity close enough on a finite dimensional subspace can be perturbed to equal the identity on that subspace. LEMMA 3.2. Let X be a Banach space, let E be an n-dimensional subspace o f X and let T : X --+ E be onto. Let k ~ n and let F be a k-dimensional subspace o f X such that ]]T[F -- IF[] < e < 1, where (1 - e) -1 ek < 1. Then there is a rank n operator S on X such that SIF = I F , I I S - Tll < (1 - e ) - l e k ] l T l l , and S * ( X * ) = T * ( X * ) . Moreover, if T is a projection then S can be chosen to be a projection. PROOF. By our assumption, U = T IF is an invertible operator on F with 11U [I < 1 + e and II U- 1 II < (1 - e ) - 1. Hence, [IU - 1 _ I IF II < e(1 - e ) - 1. By a result of Kadec and Snobar (see Proposition 9.12 in [96]) we can choose a projection P from E = T X onto T F with IIP II ~< k = dim T F. Let V = U-1 p + l i e - P, and S = V T. Now IIV - l i E II = [[U - 1 P - PII < ke(1 - e) -I < 1. Hence V : E -+ X is one-to-one and so T* and S* have the same r a n g e - since the null space of T is a finite co-dimensional subspace of X. Clearly SIF = IIF and S is a projection if T is. The rest of the lemma is easily derived from here. D
288
P.G. Casazza
As an immediate consequence of L e m m a 3.2 we have the following THEOREM 3.3. The following properties f o r a Banach space X are equivalent: (1) X has the bounded approximation property. (2) There is a uniformly bounded net (T~) of finite rank operators on X which tends strongly to the identity on X. (3) There is a ~ ~ 1 so that for everyfinite dimensional subspace E C X there is a finite rank operator T on X such that IIT ]1 <<.)~ and T x = x, f o r all x ~ E. In the case of separable spaces, we can use induction on this procedure to prove
COROLLARY 3.4. Let X be a separable Banach space. Then X has the h-bounded approximation property if and only if there is a sequence of finite rank operators (Tn) on X converging strongly to the identity so that Tm Tn -- Tn, f o r all n < m, and lim SUPn IITn II <~ ~. We will call such a sequence (Tn) an approximating sequence (or a h-approximating sequence if the ~ is important). What actually occurs in Corollary 3.4 is slightly stronger than we state. In particular, if (Tn) is a sequence of finite rank operators on a separable Banach space X which converges strongly to the identity on X, then there is an approximating sequence (Sn) on X so that for some k l < k2 < ..- lim, ]lTk, -- S, II = 0. Since there are spaces with bases whose duals fail the approximation property, we see that the bounded approximation property does not pass to dual spaces. However, it does pass from dual spaces down to the space. This was observed by Grothendieck [35] for the metric approximation property and by Johnson [41 ] for the general case. We first introduce some notation. A Banach space X is said to have the h-duality bounded approximation property provided there is a net (T~) of finite rank operators on X uniformly bounded by )~ which converges strongly to the identity on X and so that (T*) converges strongly to the identity on X*. Equivalently, X has the )~-duality bounded approximation property provided that, for each e > 0 and each pair of finite dimensional subspaces E of X and F of X*, there is a finite rank operator T on X with IIT II ~< ~, IITx - x ll ~< e]lx II, for all x 6 E and liT*x* - x*ll ~< ellx*l] for all x* 6 F. PROPOSITION 3.5. I f X* has the h-bounded approximation property than X has the )~duality bounded approximation property. PROOF. An application of Helly's Theorem (or local reflexivity) shows that we may assume that X* has the )~-bounded approximation property given by weak-star continuous operators. So choose a net of finite rank operators (T~) on X so that lim sup~ IIT~ II ~< )~ and lima T ' x * = x* for all x* 6 X*. Then (T~) converges weakly to the identity on X and so there are disjoint finite convex combinations of the T~ (which must be chosen carefully since this is only a net) which form a net and converge strongly to the identity on X. [] We will see later in this section that there are Banach spaces with the approximation property which fail the bounded approximation property. First, we recall another important
Approximation Properties
289
result of Grothendieck [35] that for separable dual spaces, the approximation property implies the metric approximation property. We refer the reader to Lindenstrauss and Tzafriri [66] for the best proof of this result. THEOREM 3.6. A separable dual space with the approximation property has the metric approximation property.
There are several important consequences of Theorem 3.6 which we will now explore. THEOREM 3.7. Every reflexive Banach space with the approximation property also has the metric approximation property. PROOF. This is immediate from Proposition 3.5 and Theorem 3.6 in the separable case. For the non-separable case we also need Proposition 9.6. [] In light of the counter-examples to the various forms of the approximation property, we see that Theorem 3.6 is quite a powerful result in that it uses the existence of a large number of finite rank operators on a separable dual Banach space to draw the conclusion that there is an equally large number of uniformly bounded finite rank operators on the space. The proof of Theorem 3.6 has always been a little mysterious. It relies on the fact that for a separable dual Banach space X, the sets Bx and Bx, are compact metric in their respective weak-star topologies. Because of this, Grothendieck's result has never been generalized to non-separable duals. PROBLEM 3.8. If X is a (non-separable)dual space with the approximation property, must X have the metric approximation property? Now we will consider the relationships between the approximation property, the bounded approximation property and the metric approximation property. The main example here comes from a construction of Figiel and Johnson [27]. PROPOSITION 3.9. Fix 1 <~ )~ < cx~. If X is a Banach space which has the )~-bounded approximation property in every equivalent norm, then X* has 2)~(1 + 4)~)-bounded approximation property. In the case of the metric approximation property, Proposition 3.11 was strengthened by Johnson [42] to: If X has the metric approximation property in every equivalent norm, then X* has the metric approximation property. If X is a Banach space with a separable dual X* having the approximation property, then for every equivalent norm I" I on X, (X*, 1. I) still has the approximation property and hence the metric approximation property. In light of Proposition 3.5 we now have THEOREM 3.10. Let X be a Banach space with a separable dual space. The following are equivalent: (1) X* has the approximation property.
290
P.G. Casazza
(2) X* has the metric approximation property. (3) X has the metric approximation property in all equivalent norms. In particular, co has the metric approximation property in every equivalent norm. Proposition 3.9 will allow us to construct several important examples. COROLLARY 3.1 1. (1) There is a Banach space which has the bounded approximation property but fails the metric approximation property. (2) There is a separable Banach space (which even has a separable dual) with the approximation property which fails to have the bounded approximation property. PROOF. Let X be a separable Banach space with a basis whose dual is separable and fails to have the approximation property (see Proposition 1.4). By Proposition 3.11, for every natural number n, there is an equivalent norm I. I, on X so that (X, I" I,) fails the n-bounded approximation property. Now (X, l" I,) is a Banach space with the bounded approximation property which fails the metric approximation property. Also, (y~, q)(X, I" I,))e2 fails the bounded approximation property (see the discussion following Definition 3.1) but has the approximation property by Proposition 2.14. D Left open here is another important question. PROBLEM 3.12. If a Banach space X has the bounded approximation property, must X have the metric approximation property in an equivalent norm? Does g l have the metric approximation property in every equivalent norm? Does g~ have the metric approximation property in every equivalent dual norm? There is a very nice classification of spaces which have the bounded approximation property proved independently by Petczyfiski [80] and Johnson, Rosenthal and Zippin [50]. Although we cannot prove this result until Section 6 (see Proposition 6.9), we will state it now.
THEOREM 3.13. A separable Banach space has the bounded approximation property if and only if it is isomorphic to a complemented subspace of a space with a basis. Moreover, if X is reflexive and has the bounded approximation property then X is isomorphic to a complemented subspace of a reflexive space with a basis. The classifications of the approximation property due to Grothendieck and appearing in Theorem 2.5 have corresponding formulations for the bounded approximation property. The proof is the corresponding bounded versions of the proof of Theorem 2.5. THEOREM 3.14. For a Banach space X, the following statements are equivalent: (1) X has the )~-bounded approximation property. (2) For every Banach space Y, the finite rank operators from Y to X of norm <~)~ are r-dense in the unit ball of (B(X, Y), r).
Approximation Properties
291
<
(3) For every K > O, every Xn E X, x* E X* with Y~n IlXn IIIIx~ II ~ , and satisfying I ~ x ~ ( T x ~ ) l <~ KIITII for every finite rank operator T on X, must also satisfy )~g.
I~ x~,(xn)l <.
There are unconditional forms of the metric approximation property [13]. The motivation for this concept comes from the observation that if X is separable and has an approximating sequence (Tn), then letting An = Tn - Tn-1 (where To = 0 ) we have for all x 6 X, x = Y~n An (x). We say that a separable Banach space X has the unconditional metric approximation property (unconditional MAP for short) if X has an approximating sequence (Tn) for which l i m n ~ I I I - 2Tn II = 1. The justification for this terminology comes from the following result of Casazza and Kalton [ 13]. THEOREM 3.15. A separable Banach space X has the unconditional metric approximation property if and only if for every e > 0 there is an approximating sequence (Tn) so that if An = Tn - Tn-1 (and To = O) then
sup
~OiAi[I
~ 1 +s.
N,On=-+- I i=1
4. The commuting bounded approximation property As we will see, having an approximating sequence (Tn) whose elements commute will yield much stronger results. This property was first isolated by Johnson and Rosenthal
[51]. DEFINITION 4.1. A Banach space X has the )~-commuting bounded approximation property ()~-commuting BAP for short) if there is a net (T~)~eI of finite rank operators on X converging strongly to the identity such that lim sup~ IIT~ II ~< )~, and for all or, fl we have T~ T/~ = Tr T~. We say that X has the commuting bounded approximation property (commuting BAP for short) if it has the )~-commuting bounded approximation property for some X~l.
Clearly the commuting bounded approximation property implies the bounded approximation property. The converse of this fails in the non-separable case (see Corollary 9.4) and is an open question in the separable case. PROBLEM 4.2. Does every separable Banach space with the bounded approximation property have the commuting bounded approximation property? Later in this section we will see that, in the separable case, the metric approximation property implies the commuting metric approximation property. In the case of separable spaces, we can greatly simplify the commuting bounded approximation property. This comes from a result of Casazza and Kalton [ 13]. Recall the definition of a )~-approximating sequence on X given in Corollary 3.4.
P.G. Casazza
292
PROPOSITION 4.3. Let X be a separable Banach space with a ~-approximating sequence (Tn). If ~--~ m [[TnTn+l - Tn+l Tn 1[ < oo, then X has a )~-approximating sequence ( S n ) satisfying Sn Sm = Smin(n,m), for all n ~ m. As an immediate consequence we have THEOREM 4.4. For a separable Banach space X, the following are equivalent: (1) X has )~ the commuting bounded approximation property. (2) There is a s sequence (Tn) on X with Tn Tm = Tmin(n,m), for n =/=m. (3) There is a s sequence (Tn) on X with
lim llZn Zm - Zm Tn ll = 0,
for all m.
n
(4) There is a s
sequence (Tn) on X with
lim IITn Tm - Tm Tn [I = O. f/,m
Problem 3.13 asks if the bounded approximation property implies the metric approximation property in an equivalent norm. Johnson [42] showed that the answer is yes for the commuting bounded approximation property. PROPOSITION 4.5. Every separable Banach space with the commuting bounded approximation property can be renormed to have the commuting metric approximation property. PROOF. Let (Tn) be an approximating sequence satisfying Tn Tm = Tmin(n,m), for n ~ m. We define a new norm on (X, [[. [[) by: [x[ = SUPn [[Tnx[[. For each natural number n let In(n-k-l)
S~=-
1 n
~
Ti,
i=l (n-1)n+l
Then (Sn) is an approximating sequence, with SnSm = Smin(n,m) ISnl---> 1.
for n ~ m and D
Casazza and Kalton [13] gave a strong converse to Johnson's result. THEOREM 4.6. Let X be a separable Banach space with the metric approximation property. Then X has the commuting metric approximation property. PROOF. Let X be a Banach space and (Tn) be an approximating sequence (recall that this implies that Tm Tn = Tn for m > n) satisfying I]Tn II <~ 1 + an, where Y~n an = / 3 < oo. For t > 0 define the operators
Vn(t ) -e
- n t exp
t
Tk k=l
--e-n' Z j--0
~(T1 +".+
Tn) j.
Approximation Properties
293
Then
IIW~(t)II <~e-ntexp( t~[lTkll)k-O <~ e/~t"
One shows that S(t)x = limn~e~ Vn(t)x, exists for all x e X. A calculation shows that S(t) is compact for all t > 0. Also, if x E Tn (X) then S(t)x = Vn (t)x --+ x, and as t ~ 0, IIS(t)ll --+ 1. Finally, since X has the metric approximation property, there are finite rank operators Rn so that IIRn - S(1/n)ll--+ O. Since S(t) has the semigroup property (i.e., S(t + s) = S(t)S(s)), the S(1/n) satisfy the requirements of Theorem 4.4(3). D The previous two results show that the metric approximation property has an isomorphic equivalent form: the commuting bounded approximation property. There is also a reverse metric approximation property. We say that X has reverse metric approximation property if there is an approximating sequence (T,) with l i m , ~ III - Tnl[ -- 1. It can be shown [13] that the reverse metric approximation property implies the commuting bounded approximation property. We also have the following result of Casazza and Kalton [ 13] concerning renormings of a space with the commuting bounded approximation property which we will need later. THEOREM 4.7. Let X have the commuting bounded approximation property and 1 <<,ot <
2. Then X can be renormed to have a commuting approximating sequence (Tn) with lim n----~ o o
IIT~ II :
lim n----~ o G
III
-
ot
Tn II =
l,
and
lim sup] Tn- T~J ~ .
1
n---~ o o
In particular, X can be renormed to have simultaneously the metric approximation property and the reverse metric approximation property. As we have seen, it is an open question whether the bounded approximation property implies the commuting bounded approximation property. Now we turn our attention to some cases where this conclusion is valid. The following result is due to Johnson, Rosenthal and Zippin [50], although it was tacitly contained in earlier work of Johnson [41 ]. PROPOSITION 4.8. If X is separable and X* has the )~-bounded approximation property
(or if X has the )~-bounded approximation property and is isomorphic to a dual space) then X has the X commuting bounded approximation property. In particular, every separable reflexive Banach space with the approximation property has the metric commuting bounded approximation property.
294
P G. Casazza
PROOF. Assume X* has the bounded approximation property. By Proposition 3.5, there is a net (Tc~) of finite rank operators on X which converges strongly to the identity on X while its net of adjoints (T*) converges strongly to the identity on X*. By induction, we extract a sequence (Tn) from the net with the property that limm--,~ IITm Tn - Tn II -- 0 and limm~ec IIT* T* - T* I I - 0 for all n = 1, 2 . . . . . Taking adjoints through the second limit we see that limm--,oc IITnT m - Tn II - 0 , for all n = 1, 2 . . . . . and so the hypotheses of Theorem 4.5(3) are satisfied. The other case is handled similarly. For the last part we just apply Proposition 2.6 and Theorem 3.6. M Zippin [99] has shown that a Banach space has a separable dual if and only if it embeds into a space with a shrinking basis. Combining the above results, we obtain the complemented version of Zippin's theorem. THEOREM 4.9. Let X be a Banach space with a separable dual. The following statements are equivalent: (1) X* has the bounded approximation property. (2) X embeds complementably into a space with a shrinking basis. (3) X has the shrinking commuting bounded approximation property. That is, there is a commuting approximating sequence (Tn) on X with (T*) converging strongly to the identity on X*. PROOF. It is immediate that (2) =~ (3) =~ (1). The implication (1) =~ (2) follows from Proposition 6.9 (or from the results of [50]). D Our next result [ 13] is useful for our constructions in Section 6. PROPOSITION 4.10. Let X be a separable Banach space and suppose there is a separable reflexive Banach space Y so that X 9 Y has the commuting bounded approximation property. Then X has the commuting bounded approximation property.
PROOF. Let (Sn) be an approximating sequence for X ~3 Y with SnSm = Smin(n,m), for all n 7~ m, and let P be the projection of this space onto X. Since (P Sn) converges strongly to zero on the reflexive space Y we have that limn (I - P)* (P Sn)*X* = 0, weakly for every x* 6 X*. It follows that PSn (I - P) converges weakly to zero in B(X, Y). Thus, we may pass to convex combinations Rn of (Sn) (which is still an approximating sequence for X • Y) but so that ( P R n ( I - P)) converges to zero in norm. Letting Tn = PRn, a calculation shows that limn,m-+~ IITn Tm - Tm Tn II - 0. Applying Theorem 4.4(4) finishes the proof. D We end this section with the unconditional version of the commuting metric approximation property similar to the unconditional metric approximation property encountered in Section 3. We say that X has the commuting unconditional metric approximation prop-
Approximation Properties
295
erty if for every s > 0 there is a commuting approximating sequence (Tn) for which if An = Tn - Tn-1 (where To = 0 ) w e have
sup
N,Oi=-+-I
~OiAi
~
i=1
There is a good classification of the commuting unconditional metric approximation property due to Godefroy and Kalton [30]. THEOREM 4.1 1. Let X be a separable Banach space with the approximation property.
Then the following conditions are equivalent: (1) X has the unconditional metric approximation property. (2) X has the commuting unconditional metric approximation property. (3) For every s > O, X is isometric to a 1-complemented subspace of a Banach space with a (1 + s)-unconditional finite dimensional decomposition.
5. The re-property In Section 3 we observed that if X has a basis, then it has the bounded approximation property with an approximating sequence given by projection operators. Lindenstrauss [61 ] isolated this property which today is called the re-property. DEFINITION 5.1. A Banach space X is said to have the rez-property if there is a net of finite rank projections (S~) on X converging strongly to the identity on X with lim supc~ IlSc~II ~< )~. The case )~ = 1 is called the metric re-property. A space with the rezproperty for some )~ is said to have the re-property. It is clear that every space with a basis has the the re-property given by the basis projections. The converse is false due to the example of Szarek [95] (see Proposition 6.7). Clearly, the re-property implies the bounded approximation property. The converse of this fails due to an important example of Read [86] of a reflexive Banach space with the bounded approximation property (and hence the commuting bounded approximation property) which fails the re-property. Actually, Read's space gives more. PROPOSITION 5.2. There is a complemented subspace of fails the re -property.
C2 (see Definition
1.10) which
We ask PROBLEM 5.3. For a separable Banach space, does the re-property imply the commuting bounded approximation property? Separable is needed here since in Section 9 we will show that this implication fails for non-separable spaces. In the case of the metric re-property, we do have an answer to Problem 5.3 due to Johnson [39].
296
P.G. Casazza
PROPOSITION 5.4. If X is separable and has the metric re-property, then X has the commuting metric re-property (and hence the finite dimensional decomposition property). PROOF. Let (P/7) be an approximating sequence of finite rank projections on X with 17 IIPnll = 1 + en and ]-In~__l(1 + e/7) < cx~. For n > j let Sj = PjPj+l "" Pn. For each j , Sj = l i m n ~ Sj exists pointwise and gives a commuting sequence of finite rank projections on X converging strongly to the identity. D Applying L e m m a 3.2 we obtain a stronger form of the zr-property for separable spaces. PROPOSITION 5.5. For a separable Banach space X, the following are equivalent: (1) X has the rex-property. (2) There is a X-approximating sequence of finite rank projections (Sn) on X for which SmSn -- Sn, for all m ~ n. The re-property does not pass to the dual space since there are Banach spaces with bases whose dual spaces fail the approximation property (see Section 1). The next result shows that the only reason the re-property does not pass to a dual space is that the dual space may fail the bounded approximation property. This is a result of Johnson, Rosenthal and Zippin [501. THEOREM 5.6. If X* has the rex-property, then so does X. If X has the rex-property and X* has the bounded approximation property, then X has the shrinking reu-property for some #. That is, there is a net of finite rank projections (S~) converging strongly to the identity on X so that (S*) converges strongly to the identity on X*. PROOF. The severe technicalities of the proof in [50] come from trying to give the best calculation for the respective )~. If we ignore this, we can get quickly to the basic idea of how one produces projections on X from projections on X*. If X* has rex-property then by Proposition 3.5, X has the bounded approximation property. So if E is a finite dimensional subspace of X and F is a finite dimensional subspace of X*, we can choose a finite rank operator S on X with IISII ~< x + 1 and so that SIE = IE and then choose a finite rank projection Q on X* with IIQII ~< )~ + 1 and with QIF = IIF and S*(X*) C Q(X*). A simple application of local reflexivity to L e m m a 3.2 shows that we may as well assume that there is a projection P on X with P* = Q. Now, P ' S * = S* implies that S = S P. This fact yields that the operator S + P - S P is a projection on X which is the identity on E and its adjoint is the identity on F. If we consider the set of pairs (E, F) so that E is a finite dimensional subspace of X and F is a finite dimensional subspace of X*, this becomes a net when it is ordered by inclusion of the subspaces. By the above, we can construct projections PE,F on X with PE,F -- I on E and P{,F -- I on F. It follows that (PE,F) gives the shrinking re-property. Assume now that X has the rex-property and X* has the bounded approximation property. If we choose finite dimensional subspaces E C X and F C X*, we can find a finite rank operator S on X and a finite rank projection P on X so that PIE -- I[E and S*[F -- IF
Approximation Properties
297
and and the norms of the operators are functions of)v and the bounded approximation property constant alone. Now, if Q = S + P - SP, then Q is our finite rank projection on X which equals the identity on E and Q* = S* + P* - P* S* is our finite rank projection on X* which equals the identity on F. [5 The central theme in the 7r-property is whether we can construct large classes of uniformly bounded finite rank projections on a Banach space. Read's example [86] says that even if we have a large number of uniformly bounded finite rank operators on a space (even a commuting family) this is not enough to induce a large number of projections on the space. So we now turn to the question of constructing projections on a Banach space. It is not surprising that a bounded operator which is sufficiently close (in norm) to its square will induce a projection on a Banach space. The exact measure of closeness required was found by Casazza and Kalton [ 13] and is surprisingly large. PROPOSITION 5.7. Let X be a Banach space, T a bounded linear operator on X, and assume that I ] T - T2[] = 0 < 1/4. Then there is a projection P on X such that {x: T x = x } C P ( X ) C T ( X ) and 1( I+2]ITII ) IIPII <~ ~ 1 + ( 1 - 4 0 ) 1 / 2 " PROOF. Define
m,
S
mzO
and P--
21(I - (I - 2T) S)
Power series manipulation and some calculating show that this works. Now we have the result of Casazza and Kalton showing that 1/4 is the best constant. THEOREM 5.8. Let X be a separable Banach space with an approximating sequence (Tn) f o r which limsuPn__+oo ]]Tn - T2]] < 1/4. Then X has the 7c-property. Moreover, there is a Banach space X which fails the 7c-property, but f o r which there is an approximating sequence (Tn) on X satisfying 1 lim sup] Tn - T2[] -- ~. n---+ oo
PROOF. By Corollary 3.4 and the statements following it, we may assume that our approximating sequence (7",) has the additional property that Tm Tn = Tn, for all m > n. Now the result follows easily from Proposition 5.7. For the second part of the theorem, we observe that Read's example [86] of a reflexive Banach space with the commuting bounded approximation property but failing the 7r-property, has an approximating sequence (T,) with lim s u p , ~ I]Tn - T2 ]l -- 1/4 by Theorem 4.7. U
298
P G. Casazza
An important tool for constructing finite rank projections on a Banach space is due to Johnson [40]. This construction, and its variations, have had widespread application in the area. So we will now consider this technique in detail. Recall the definition of the universal spaces with a basis Cp given in Definition 1.10. THEOREM 5.9. I f X has the bounded approximation property, then X 9 Cp has the yrproperty, f o r all 1 <<,p < ~ and p = O. PROOF. Recall that Cp = (Y~i ~[~Ei)gp where the Ei are dense in all finite dimensional spaces. Fix a finite dimensional subspace F of X and let H -'- Y ~ in: I OEi for a fixed n. Let T be a finite rank operator on X so that T I F - - I and IIT[I ~< )~ + 1. Let E = (I - T ) T ( X ) and choose an m > n so that d (E, Em) < 2, and let L be an isomorphism from E onto Em with IlL II = 1 and ILL-1 II < 2. Let P be the natural norm 1 projection of C p onto H and let Pm be the natural projection of Cp onto Em. Define S : X 9 Cp --+ X @ Cp by:
S(x, y) - ( T x + L -1 PraY, L ( I - T ) T x + L ( I - T ) L -1 PmY + PY). Then S is a finite rank projection on X G C p which is the identity on both F, H and with norm a function of the norm of T alone. [] In the next section we will see some strengthenings of Theorem 5.9. In particular, if X has the bounded approximation property then X @ C ~ has a basis. If X has the commuting bounded approximation property then X @ C p has a basis for all 1 ~< p ~< oo. A consequence of Theorem 5.9 and Proposition 4.10 is that Problems 4.2 and 5.3 are equivalent in the sense that either both have positive answers or both have negative answers. There are many variations of the above construction which have been adapted for special uses. Casazza [7] used a variation of this technique to show that James' space J is primary. Casazza [8] gave another variation of this technique showing that whenever Y is a complemented subspace of a reflexive Banach space X with a (sub) symmetric basis then X r Y has a basis. Yet another variation of this method was used by Casazza [9] to show that for separable Banach spaces X with the commuting bounded approximation property, there is a subspace Y of X so that both Y and X r Y have finite dimensional decompositions.
6. The finite dimensional decomposition property A Schauder basis decomposes a Banach space into a direct sum of one-dimensional subspaces. A cruder form of this is to decompose the space into finite dimensional subspaces. DEFINITION 6.1. A sequence (En) of finite dimensional subspaces of a Banach space X is called a (unconditional)finite dimensional decomposition for X (FDD (UFDD) for short) if for every x ~ X there is a unique sequence Xn ~ En so that x -- Y~n Xn (and this series converges unconditionally). In this case we will write X = Z n ~ En and say X has a FDD (finite dimensional decomposition).
Approximation Properties
299
As observed in the Basic Concepts Chapter of this Handbook [47], all the basic properties for bases hold in this setting. The decomposition constant is sup~ [1Pn 1] < oc, where the Pn are the natural decomposition projections given by Pn (y~oc n Xi. We i=1 X i ) = ~i=1 have the notions of shrinking and boundedly complete finite dimensional decompositions and the standard results that a unconditional finite dimensional decomposition for X is shrinking if and only if el does not embed into X and that it is boundedly complete if and only if co does not embed into X. It is clear that every space with a basis has a finite-dimensional decomposition (even into 1-dimensional subspaces). The converse is false as Szarek [95] has constructed a space with a finite dimensional decomposition which fails to have a basis. Also, a complemented subspace of a space with a finite dimensional decomposition (or even a space with a basis) need not have a finite dimensional decomposition. This is an example of Read [86] (see Proposition 6.7). It is also clear that a space with a finite dimensional decomposition has the Jr-property given by the finite dimensional decomposition projections. The converse is an open question. PROBLEM 6.2. Does every separable Banach space X with the Jr-property have a finite dimensional decomposition? We see from the definitions that X has a finite dimensional decomposition if and only if there is an approximating sequence (P~) of commuting projections on X, and in this case the finite dimensional decomposition is just X = ~-,n G(Pn - P n - l ) ( X ) where P0 = 0. Casazza [9] isolated these two properties as classifying spaces with finite dimensional decompositions. THEOREM 6.3. A separable Banach space has a finite dimensional decomposition if and only if it has both the commuting bounded approximation property and the Jr-property. PROOF. Let (jrn) be an approximating sequence of finite rank projections on X and let (Tn) be a )~-commuting approximating sequence of finite rank operators on X (so Tn Tm -rmin(n,m), for all n 7~ m). Let Y be
Y -- U T~X* -- span U T~x*. The second equality follows from the fact that Tin*T~* = T~*, for all m > n. This property also yields that (T~*) is a commuting approximating sequence for Y. By Lemma 3.2, we can choose n l < n2 < ... and finite rank operators (Si) on X with (1) IIsi II ~< 2)~. (2) Si Jri = Jri .
(3) S*l (X*)= T~ (X*) c Y. Now let Qi -- Jri Si. A direct calculation shows that the Qi are projections which by (2) above converge strongly to the identity on X. Now, even though X* may fail the approximation property, we can use Lemma 3.2 to mimic the construction in the proof of Theorem 5.6 using Y in place of X*. Now, by induction we can find uniformly bounded finite
300
P.G. Casazza
rank projections Pn on X so that Pm Pn = Pn and Pm*P* - Pn for all m >~ n, and Pn*(X*) C Y. Finally, (Pn) is the required commuting sequence of finite rank projections on X converging strongly to the identity on X and (Pff) converges strongly to the identity on Y. [3 This result makes Problem 6.2 equivalent to Problems 4.2 and 5.3. In this setting, Proposition 5.4 and Proposition 3.8 can be formulated as THEOREM 6.4. A separable Banach space X with any of the following properties has a finite dimensional decomposition: (1) X has the metric Jr-property. (2) X* has the re-property. (3) X has the re-property and is isomorphic to a dual space. In particular, every separable reflexive space with the Jr-property has a finite dimensional decomposition. (4) X has the 7r-property and X* has the bounded approximation property. The importance of finite dimensional decompositions does not stem from the fact that they have properties common with spaces with bases. Their importance comes from the fact that there are results for spaces with finite dimensional decompositions whose analogues are either false or unknown for spaces with bases. For example, Johnson [44] has constructed a Banach space X with the property that every subspace of X has a finite dimensional decomposition, but not every subspace has a basis. Also, it is unknown if every separable weak Hilbert space has a basis. But Maurey and Pisier (see [74]) have shown that every separable weak Hilbert space has the 7r-property, and hence a finite dimensional decomposition. It is unknown if every separable Banach space has a subspace with a basis so that the corresponding quotient space has a basis. But it is known [51 ] that every separable Banach space X has a subspace Y so that both Y and X~ Y have a finite dimensional decomposition. Moreover, if X* is separable, both these finite dimensional decompositions can be chosen to be shrinking. The above constructions can be found in Lindenstrauss and Tzafriri [66]. Also, the above constructions use an important technique in this area called "blocking techniques" which can also be found in [66]. One use for finite dimensional decompositions is in the construction of Banach spaces with bases. PROPOSITION 6.5. Let X = Y~n GEn be a finite dimensional decomposition (with FDD constant K ) f o r a Banach space X. If each (En) has a basis (xn)~" 1 with basis constant bounded by M, then ((xn)~n 1)n~__l is a basis for X with basis constant ~ K M . This yields our earlier stated result of Petczyfiski [80] and Johnson, Rosenthal and Zippin [50]. THEOREM 6.6. For a separable Banach space X the following are equivalent: (1) X has the bounded approximation property. (2) X has a finite dimensional expansion of the identity. (3) X is isomorphic to a complemented subspace of a space with a basis.
Approximation Properties
301
PROOF. (2) =~ (1) is obvious and (3) =~ (2) follows from the proof of Proposition 1.2. So let X have (1) and we prove X is complemented in a space with a basis. By the argument of the proof of Proposition 1.2 (where we now allow the operators Ai to be just finite rank) we conclude that X is complemented in a space Y with a FDD, say Y ~- Y]~n @En. By Propositions 1.5 and 6.5, we conclude that Y • C p has a basis. M Now we will take a closer look at the examples of Szarek [95] and Read [86]. PROPOSITION 6.7. (1) There is a complemented subspace of C2 with a finite dimensional decomposition which fails to have a basis. (2) There is a complemented subspace of C2 which fails the Jr-property. PROOF. These constructions are quite difficult and cannot be covered here. But we will discuss at an intuitive level how they are carried out. We choose a sequence (En) of finite dimensional Banach spaces with the property that the basis constant of En G ~2 is >/n. To get a space with a finite dimensional decomposition which fails to have a basis, we take X -- (y-~'GEn)e2. It follows that X is a complemented subspace of C2. We now have a space with a unconditional finite dimensional decomposition and it cannot have a basis taken from the En's. The main problem is that this space might have a basis coming from Proposition 1.4. To prevent this from happening, we choose the En's carefully as subspaces of s with Pn ,~ 2. The main point then is to show that we can pick the En's so that En @ s fails to have a good basis, and choose p~ so that any subspace of s of dimension n-1
~< Y]~i=l dim Ei is a good Hilbert space. Now, in order to build a good basis for E~, the span of {E1 . . . . . En- 1} is too small to be of help, while elements chosen from the span of En+l . . . . have the property that any set of dim En of these elements are spanning a good Hilbert space and therefore cannot help either. For (2) we have the same subspaces En, only now we need to glue them together more carefully. We now need to form the En's into a Banach space in such a way that we do not introduce new projections. We do this by overlapping the En's carefully and gluing them together in such a way that there are uniformly bounded finite rank operators passing through the overlaps but no good projections. This program is carried out by carefully embedding the E~ 's (with overlap) into C2. This is a very delicate operation and the complexity of Read's paper reflects this. In both cases above, the space X we construct is isomorphic to a complemented subspace of C2 and X has the bounded approximation property. Since C2 is reflexive, X has the commuting bounded approximation property and so by Proposition 6.8, X | C2 has a basis. That is, in both cases above, our space X is complemented in a space with a basis. E3 Proposition 6.5 does not hold for unconditional finite dimensional decompositions since the Kalton-Peck space [54] has a unconditional finite dimensional decomposition into two dimensional subspaces but fails to have a unconditional basis (or even be complemented in a space with a unconditional basis). However, a space X has a unconditional finite dimensional decomposition if and only if it has a unconditional finite dimensional expansion of the identity. Also, every Banach space with a unconditional finite dimensional decomposition embeds into a Banach space with a unconditional basis (see [66, Theorem 1.g.5, p. 51 ]). We ask
302
P.G. Casaba
PROBLEM 6.8. If X has a unconditional finite dimensional decomposition and X has GLLUST, must X have a unconditional basis? Casazza and Kalton [14] have given a partial positive answer to Problem 6.8: If X has GL-LUST and a unconditional finite dimensional decomposition X = ~ O En with SUPn dim En < ~ , then X has an unconditional basis (whose basis elements may be chosen from the En's). In particular, what the Kalton-Peck space [54] is missing for it to have a unconditional basis is GL-LUST. Another general method for constructing bases uses the universal spaces C p of Definition 1.10. This is an immediate consequence of the construction of Johnson (see Theorem 5.9). PROPOSITION 6.9. If X is separable and has commuting bounded approximation property, then X G Cp has a basis, for all 1 ~ p <<,cx~. Moreover, if X G Cp has a basis for some 1 < p < cx~, then X has the commuting bounded approximation property. PROOF. By Theorem 5.9, X 9 Cp has the Jr-property. Since it clearly has the commuting bounded approximation property, we could apply Theorem 6.3 to show it has a basis. However, if we go back to the proof of Theorem 5.9, we can see directly that the projections constructed there are commuting as long as our original approximating sequence (Tn) was commuting. Either way, X @ C p has a finite dimensional decomposition ~n G En. But Cp ~- Cp @ Cp and so we just need to observe that ( ~ n GEn) G Cp has a basis. But this is a simple application of Propositions 1.5 and 6.5. The moreover part of the proposition follows from Proposition 4.10. [] An important question is whether Proposition 6.8 holds (for 1 < p < cx~) with only the assumption that X has the bounded approximation property. By Proposition 4.10, this question is equivalent to Problems 4.2 and 5.3. Lusky [68] gives a positive answer to this question for p = 0 (see also Lusky [70]). PROPOSITION 6.10. If X is separable and has the bounded approximation property, then X G Co has a basis. It was well known for some time (see [95] for a proof) that every separable space X containing co has the local basis property. That is, there is a K >~ 1 and finite dimensional subspaces E1 C E2 Q ... of X with U En dense in X and each Ei has a basis with basis constant ~< K. If X is a Banach space for which every subspace has the local basis property, then there is a constant K > 0 so that for every finite dimensional subspace E of X there is another finite dimensional subspace F of X so that span (E U F) ~2 E G F and E • F has a basis with basis constant ~< K. That is, if every subspace of X has the local basis property that every finite dimensional subspace of X is well complemented in a finite dimensional subspace of X with a good basis. One consequence of Lusky's proof of Proposition 6.10 is that every/21-space has this property, i.e., every subspace of a s has the local basis property.
Approximation Properties
303
PROPOSITION 6.1 1. There is a constant C > 0 so that f o r every 1 <<, p <, 2 and every finite dimensional subspace E o f L p[O, 1], there is a finite dimensional subspace F o f ~1 so that the basis constant o f E q) l F is <, C. PROOF. Let (E,)n~=l be a sequence of finite dimensional quotient spaces of ~ for m = 1, 2, 3 . . . . which are dense (in the Banach-Mazur distance) in the family of all finite dimensional quotients of s Let
Y =
GEn n-- 1
9 co
Then Lusky [68, Corollary 2.2] shows that the space Y has a shrinking basis (xn, xS) with basis constant, say M. In particular, Y* has a basis (x*). But
Y* '~
@E* n--1
61
is isomorphic to a subspace of el and (E*) is dense in the family of all finite dimensional subspaces of g l. Also, g p is uniformly finitely representable in ~1 with constant, say K. Hence, for any finite dimensional subspace E of L p[O, 1], E is K-isomorphic to some space E* which in turn is 1-complemented in the space Y* with a basis. This is all we need for the proposition. [] COROLLARY 6.12. There is a subspace X o f g~l with a basis so that X* fails the approximation property. PROOF. Let X = Y* from the proof of Proposition 6.11. Now, Y has a basis and we claim that Y* fails the approximation property. To see this, choose a subspace Z of el which fails the approximation property and choose a sequence of finite dimensional subspaces Fl C F2 C ... of Z whose union is dense in Z. Then there are natural numbers (kn)~ec=1 so that Fn ~ 2 Ek*' 9Hence if
w-
eF, n=l
61
then by the proof of Proposition 1.12 we have that W* fails the approximation property and clearly W is isomorphic to a complemented subspace of Y. So Y fails the approximation property. [-1 We do not know if every subspace of L p[O, 1] has the local basis property for any 1 < p 7~ 2 < ec. These results concern constructing (local) bases from inside a Banach space. In general, this is a very delicate operation. There are a few bright spots here however. Let X be reflexive and have a symmetric basis and let Y be a complemented subspace
304
P.G. Casazza
of X. Then X G Y has a basis [8]. Also, if X is separable and has the commuting bounded approximation property, then there is a subspace Y of X so that both Y and X @ Y have finite dimensional decompositions [9]. However, Mankiewicz and Nielsen [71 ] have shown that there is a superreflexive space with a finite dimensional decomposition such that for every subspace Y of X, the space X ~3 Y fails to have a basis. Finally, if Y is any reflexive subspace of a space with a unconditional finite dimensional decomposition, then Y is isomorphic to a complemented subspace of a space with a unconditional finite dimensional decomposition if and only if Y has the approximation property [24]. In general, it is very difficult to discover if a space has a unconditional finite dimensional decomposition even if we assume it has a basis. PROBLEM 6.13. If X is complemented in a space with an unconditional basis, does X have a unconditional finite dimensional decomposition? What if we assume that X has a finite dimensional decomposition? Problem 6.11 is unsolved even in concrete cases such as the Orlicz spaces or the Lorentz spaces. There is one positive but technical result due to Alspach and Carothers [3] which solves this problem in very special circumstances. We end this section with a more general concept of an infinite decomposition for a Banach space. A sequence (X/7) of closed subspaces of a Banach space X is called a Schauder decomposition for X if for every x E X there is a unique sequence x/7 ~ Xn so that x - ~ n Xn. If the Xn are finite dimensional, this becomes a finite-dimensional decomposition. Many basis properties generalize trivially to this setting. For example, we have a uniformly bounded sequence of projections Pn Y~i=l xi) Y~i=l xi Conversely, every bounded sequence of projections (Pn) on a Banach space X for which Pn Pm = Pmin(n,m) and l i m n ~ Pnx = x for all x 6 X determines a Schauder decomposition Xn = (Pn - P n - 1 ) ( X ) for X (where P0 - 0). We also have the notion of shrinking and boundedly complete and unconditional decompositions. It was an open question until just recently whether every separable Banach space must have a Schauder decomposition. A counterexample was observed by Allexandrov, Kutzarova and Plichko [2]. OO
m
/7
.
PROPOSITION 6.14. There is a separable Banach space with no Schauder decomposition. PROOF. Since the hereditarily indecomposable space of Gowers and Maurey [33] contains g~ uniformly, it must have a subspace X failing the approximation property. If (X/7) is a Schauder decomposition for X, then there is an no so that X/7o is infinite dimensional (otherwise, this would be a finite dimensional decomposition for X and so X would have the approximation property). Now, the natural projection onto X/70 decomposes X which is a contradiction. 71
7. The uniform approximation property The property which corresponds to the bounded approximation property in the local theory of Banach spaces is the uniform approximation property (UAP). The uniform approxima-
Approximation Properties
305
tion property was introduced by Petczyfiski and Rosenthal [81] where they showed that L p(l z) has it for all 1 ~< p ~< cx~. DEFINITION 7.1. A Banach space X has the uniform approximation property (UAP for short) if there is a constant K and a function f (n) on the natural numbers so that for all n and all n-dimensional subspaces E C X, there is an operator T : X --+ X with rk T ~< f (n) such that 11T1[ ~< K and T [E = I[E. If for every 0 < e < 1, we can replace the K above by 1 + e, we say that X has the metric UAR And if the operators T can be chosen to be projections, we say that X has the uniform projection approximation property (UPAP for short). Given K > 1, let
kx(K,n)
-
sup
-
inf{rk T" T e L(X), IITII <~ K, TIE -- liE].
E C X, dim E = n
So X has the uniform approximation property if and only if there is a constant K such that k x ( K , n) is finite for all n, and in this case we say that X has the K-UAP. The uniform approximation property is much harder to check than the other approximation properties. The list of spaces known to have this property is quite small and includes only a few large classes of spaces: the L p-spaces and the reflexive Orlicz spaces [65] (both classes actually have the UPAP), and spaces obtained from these. Samuel [89] shows that ( ~ OLp)eq has the uniform approximation property. This was generalized by Heinrich [37] (see Theorem 7.5 below). A deep result of Jones [52] is that H 1 has the uniform approximation property. Szankowski [94] shows that the matrix spaces Sp fail the uniform approximation property (for p > 80). It is unknown if the disk algebra has the uniform approximation property. If true, this would be even more useful and interesting in light of the fact that it is unknown if H ~ has the approximation property. Clearly, if H is a Hilbert space then k/4 (1, n) = n. Hence if X is isomorphic to a Hilbert space, then there is a constant K such that
k x ( K , n) -- n
for all n.
The converse of this is also true by the complemented subspaces theorem of Lindenstrauss and Tzafriri [64]. From our discussion concerning the weak Hilbert space T 2 in Section 1, we have that T 2 has the uniform projection approximation property with KT2 ( 8 , n ) - - 2n (or even K T z ( K , n ) -- (1 + e)n for any fixed e and K -- K(e)). That is, for any 2ndimensional subspace E C T 2, E C span{(ti)in_=l , F}, where F is an n-dimensional subspace of span(ti)i~=n+l. But now F is 8-complemented in T 2. Johnson and Pisier [49] showed that such linear behavior of kx actually characterizes weak Hilbert spaces. THEOREM 7.2. A Banach space X is a weak Hilbert space if and only if there are constants K, C such that
k x ( K , n) <~Cn
for all n.
306
P.G. Casazza
The following result of Johnson (which appeared in [64]) reduces the checking of the uniform approximation property (at least in Banach lattices) to working with disjoint elements. PROPOSITION 7.3. There is a function N(k, e) <<,(2k2/e) k so that for any order complete Banach lattice L, and every k-dimensional subspace F of L, there are, for every e > O, N -- N(k, e) disjoint elements (xi)N1 in L and a linear operator T" F --+ X =: span(xi)Ul such that
II Tx - x II ~ ex for all x E F. Johnson [43] used this to show there are reflexive Banach spaces which are not sufficiently Euclidean (i.e., reflexive Banach spaces which do not contain uniformly complemented copies of g~). The disadvantage of this result is that it immediately jumps the value of the uniformity function to f ( K , n ) ~ n 2n. The advantage is that it trivially implies that spaces like gp and L p have the uniform projection approximation property since disjointly supported sequences in these spaces span 1-complemented subspaces. Local reflexivity implies that X has the uniform approximation property if and only if X** has the uniform approximation property. It was open for a long time whether the uniform approximation property passes from a space to its dual. This was done by Heinrich [37] using ultra-power techniques. Recently Mascioni [75] gave a direct proof with the best estimates of the UAP constant for X* given the UAP constant for X. THEOREM 7.4. A Banach space X has the uniform approximation property if and only if X* has the uniform approximation property. Heinrich [37] demonstrated that ultra-product techniques represent a powerful tool for working with the approximation property. THEOREM 7.5. Let X be a Banach space.
(1) If X has the uniform approximation property and 1 ~ p ~ ~ then L p (lZ, X) has the uniform approximation property. (2) X has the (K + e)-uniform approximation property (respectively uniform projection approximation property), for all e > 0 if and only if each ultrapower (X)u has the )~-uniform approximation property (respectively uniform projection approximation property) if and only if each ultrapower (X)u has the ~-bounded approximation property (respectively Jrz-property).
We end this section by considering the relationship between the uniform approximation property and the other approximation properties. Certainly the uniform approximation property implies the bounded approximation property. There are spaces (even spaces with unconditional bases) which fail the uniform approximation property. For one, since the uniform approximation property passes to all dual spaces of the space, any space with a basis whose dual space fails the approximation property cannot have uniform approximation property. In particular the spaces C1, U1 of Definition 1.10 fail to have the uniform
Approximation Properties
307
approximation property. It follows from Proposition 4.8 that a separable space with the uniform approximation property also has the commuting bounded approximation property. This result fails in the non-separable case since goc has the metric uniform approximation property but fails the commuting bounded approximation property by Corollary 9.4. And if X is any Banach space with the uniform approximation property and the re)~-property, then by Theorem 5.6, all the duals of X have the rex-property. There are many natural open questions here, some of which would resolve earlier problems concerning the approximation properties. PROBLEM 7.6. Does the uniform approximation property imply the metric uniform approximation property? If the answer to Problem 7.6 is yes, then we can drop altogether the notion of K-UAE PROBLEM 7.7. Does the uniform approximation property imply the metric approximation property? Note that restated, Theorem 7.5 implies that Problems 7.6 and 7.7 are equivalent. Also, if these problems have a negative answer, then we can give a negative answer to Problem 3.9. That is, if X is a Banach space with the uniform approximation property and failing the metric approximation property, then there is an ultraproduct Y of X so that Y has the uniform approximation property and fails the metric approximation property. Now, Y* has even the uniform approximation property while failing the metric approximation property. Lindenstrauss and Tzafriri [65] have given a positive answer to Problem 7.6 for uniformly convex spaces. THEOREM 7.8. A uniformly convex Banach space with the uniform approximation property also has the metric uniform approximation property. We can also ask: PROBLEM 7.9. Does the uniform approximation property and the re-property imply the uniform projection approximation property? Or does the uniform approximation property itself imply the uniform projection approximation property? Because of the ultraproduct characterization of the uniform approximation property there should be a relationship between the uniform approximation property and g~-sums. PROBLEM 7.10. If X is a Banach space and ( ~ @X)e~ has the bounded approximation property does X have the uniform approximation property? Heinrich [37] has also introduced a uniform version of the approximation property. We say that a Banach space X has the uniform Grothendieck approximation property if there are uniformity functions f(., .) and g(., .) satisfying for every e > 0 and for every/3 = (/3n) with/3n --+ 0 and for every x, ~ X with IIx~II ~< r there is a finite rank operator T on X
308 with rk T ~< f (fl, e), proves [37]
P.G. Casazza
II T II ~ g(fl, e), and IITxn - Xn II ~ e for every n = 1, 2 . . . . . Heinrich
PROPOSITION 7.1 1. Let X be a Banach space. Then the following hold: (1) X has the uniform Grothendieck approximation property if and only if X* has it. (2) X has the uniform Grothendieck approximation property implies that Lp(#, X) has it for all 1 <<,p <<,oc. (3) X has the uniform Grothendieck approximation property if and only if each ultrapower (X)u has the approximation property. It is immediate that the uniform approximation property implies the uniform Grothendieck approximation property. We ask PROBLEM 7.12. Does the uniform Grothendieck approximation property imply the uniform approximation property? The answer to Problem 7.12 is "yes" in the super-reflexive setting [37]. Again, if Problem 7.12 has a negative answer, then by Proposition 7.11 we can construct a counterexample to Problem 3.8. We just take a space X with the uniform Grothendieck approximation property and failing the uniform approximation property. Then there is an ultrapower Y of X which has the uniform Grothendieck approximation property and fails the bounded approximation property. Now, Y* has the uniform Grothendieck approximation property (and hence the approximation property) while failing the bounded approximation property.
8. The compact approximation property The compact approximation property was somewhat mysterious until recently when the first major results were done. This was in part due to the fact that it was not clear that this property was actually distinct from the approximation property. DEFINITION 8.1. A Banach space X has the compact approximation property (compact AP for short) if for every e > 0 and every compact set K in X there is a compact operator T ~ L(X) so that I[Tx - x[I ~< e for all x 6 K. We say X has the )~-bounded compact AP if these approximating compact operators can be chosen with IIT II ~< )~ (independent of the compact set K). If )~ = 1 works, we say X has the metric compact AP.
If X has the )~-bounded compact approximation property for some ~, we say X has the bounded compact approximation property. The spaces failing the approximation property we encountered in Section 2 all fail the compact approximation property. It is immediate that the approximation property implies the compact approximation property. In fact, Theorem 2.5 yields PROPOSITION 8.2. A Banach space X has the )~-bounded approximation property if and only if X has both the approximation property and the )~-bounded compact approximation property.
Approximation Properties
309
Willis [98] has shown that the approximation property and the compact approximation property are not equivalent. PROPOSITION 8.3. There is a separable reflexive Banach space having the compact approximation property but failing the approximation property. Willis' construction is really another factorization theorem. Johnson [40] showed that if an operator T E B(X, Y) is a norm limit of finite rank operators then T factors compactly through C p (see Definition 1.10). Figiel [26] shows that every compact operator T E B(X, Y) factors compactly through a subspace of Cp. A close examination of Willis' proof yields THEOREM 8.4. Every compact operator T E B(X, Y) factors through a reflexive Banach space which has the metric compact approximation property. The basic properties of the compact approximation property are not resolved yet. PROBLEM 8.5. If X* has the compact approximation property, must X have the compact approximation property? The Lindenstrauss and Tzafriri proof [66] of Grothendieck's Theorem 3.6 [35] works for a special case of the compact approximation property. THEOREM 8.6. If X* is separable and has the compact approximation property given by weak-star continuous operators, then X* has the metric compact approximation property. But the general case here is open. PROBLEM 8.7. If X* has the compact approximation property (even just for X* separable) must X* have the metric compact approximation property? Casazza and Jarchow [ 11 ] have shown that the bounded compact approximation property does not pass from a dual space to the space. PROPOSITION 8.8. There is a Banach space which has the approximation property, does not have the bounded compact approximation property, but all of its duals are separable and have the metric compact approximation property. PROOF. Let Y be the reflexive space of Willis [98] which has compact approximation property and fails the approximation property. By Proposition 1.3, there is a Banach space Z with Z** having a basis and Z*** ~ Z* @ Y*. Now, Z*** fails the approximation property but has shrinking compact approximation property (since Z* has a shrinking basis and Y is reflexive and has the compact approximation property). By Proposition 3.11, we can
310
P G. Casazza
renorm X with I" In so that it fails the n-bounded approximation property (and hence it fails the n-compact approximation property by Proposition 8.2). Now,
x-
e ( z * * , I. In) E2
does the trick.
E]
Godefroy and Kalton [30] have extended Theorem 4.6 to show that a separable space with the metric compact approximation property has the commuting metric compact approximation property. The compact approximation property shows up naturally in many settings. For example, Dixon [22] shows that a Banach space has the )~-bounded compact approximation property if and only if K (X) has a )~-bounded left approximate identity. The compact approximation property shows up in the study of M-ideals. A subspace M of a Banach space X is called an M-ideal if there is a projection P on X* with M • as its kernel and for all x* E X* we have IIx*ll- IIPx*ll + I1(I - P)x*ll. M-ideals were first studied by Alfsen and Effros [ 1]. Harmand and Lima [36] proved that if X is a Banach space such that K (X) is an M-ideal in L(X), then X has the metric compact approximation property. Cho and Johnson [ 19] proved a strong converse to this by showing that if X is a subspace of ~p for 1 < p < cx~, with the compact approximation property, then K (X) is an M-ideal in L(X). After work by several authors, Kalton [53] has classified the separable Banach spaces X such the K (X) is an M-ideal in L(X). Gowers and Maurey [33] have exhibited classes of Banach spaces X for which every bounded operator on X is a strictly singular perturbation of the identity. It is still an open question whether there are Banach spaces X for which every bounded operator on X is a compact perturbation of the identity. Godefroy and Saphar [31 ] have asked whether there are any infinite dimensional Banach spaces X for which K(X) is reflexive? They then observed that the answer to this question is positive if there is a Banach space X which fails the compact approximation property and so that L(X) = K (X) G I.
9. Approximation properties in non-separable spaces Many of of the proofs of our earlier results do not depend upon the separability of the Banach space. We will list here which results (proofs) carry over to this setting and the few for which problems occur and why. THEOREM 9.1. Let X be a non-separable Banach space. If X* has the approximation property (respectively)~-bounded approximation property, zcz-property, the uniform approximation property, the uniform projection approximation property) then so does X. Missing from Theorem 9.1 is the compact approximation property (for which this is an open question even in the separable case) and the bounded compact approximation
Approximation Properties
311
property since this implication fails even in the separable case (see Proposition 8.8). Also missing from Theorem 9.1 is the commuting bounded approximation property. This also fails as shown by Casazza, Kalton and Wojtaszczyk [ 16]. To put this result into its proper framework, we first need a definition. DEFINITION 9.2. A non-separable Banach space X is said to have the separable complementation property if for every separable subspace Y in X, there is a separable subspace Z with Y C Z C X and Z is complemented in X. We now have [ 16] THEOREM 9.3. Let X be a non-separable Banach space with the commuting bounded approximation property. Then X has the separable complementation property. PROOF. Let X have the commuting bounded approximation property and let Y be any separable subspace of X. Choose an increasing sequence (En) of finite dimensional subspaces of Y whose union is dense in Y. By induction, we can find a sequence (Tn) of finite rank operators on X so that Tn Tm - - T m i n ( n , m ) for all n r m and Tn IEn -- IEn. Now, let Z be the separable subspace of X which is the closed linear span of the Tn (X). By switching to a pointwise convergent subnet of (Tn), we obtain a bounded linear operator P mapping X onto Z and a calculation shows that P - p2. IS] Since co is complemented in every separable Banach space containing it, we have COROLLARY 9.4. If X has the commuting bounded approximation property, then every subspace of X which is isomorphic to co is complemented in X. Hence, ~ does not embed into any Banach space with the commuting bounded approximation property
Of course, both the dual and the pre-dual of g ~ have the commuting bounded approximation property. A very useful tool for working in non-separable spaces is the "Lindenstrauss subspace" defined in [61 ]. PROPOSITION 9.5. Let X be a Banach space and let E be a finite dimensional subspace of X. Let k be an integer and let e > O. Then there is a finite dimensional subspace F of X containing E such that for every subspace Y of X containing E with dim Y / E -- k there is a linear operator T : Y --+ F with IITII ~< 1 + ~ and T x = x for every x ~ E. By refining Proposition 9.5 in the reflexive case, Lindenstrauss [61 ] shows THEOREM 9.6. Every reflexive Banach space has the separable complementation property with norm one projections. Actually, Lindenstrauss' method shows more. For a reflexive Banach space X and any separable subspaces Y C X and Z C X*, there is a projection P on X with separable range such that Y C P ( X ) , Z C P*(X*) and IIPII -- 1. We observed in Section 2 that if every
312
P.G. Casazza
separable subspace of X is contained in a separable subspace with the approximation property, then X has the approximation property. Johnson [42] proves the corresponding result for the bounded approximation property. This proof is based on ideas of Lindenstrauss [61 ] which are interesting in that it uses discontinuous, non-linear operators to construct a family of approximating bounded linear operators. THEOREM 9.7. Let X be a non-separable Banach space. Then X has the X-bounded approximation property (respectively the fez-property, the uniform approximation property, the uniform projection approximation property) if and only if every separable subspace of X is contained in a separable subspace with the ),-bounded approximation property (respectively the rc)~-property, the uniform approximation property, the uniform projection approximation p rope rty). PROOF. We first check the bounded approximation property. If X has the )~-bounded approximation property and Y is a separable subspace of X, let (En) be an increasing sequence of finite dimensional subspaces of Y whose union is dense in Y. Now we can construct by induction a sequence of finite rank operators (Tn) on X with the property that Tn+l IEnUTn(X) = I. Now letting Z be the closed linear subspace of X (which contains Y) generated by (Tn (X)) will do it. Conversely, assume that every separable subspace of X is contained in a separable subspace with the ~.-bounded approximation property. Let E be any finite dimensional subspace of X. By Proposition 9.5, there exists a separable subspace Y of X so that, given any finite dimensional subspace F of X with E C F, there exists an operator TF : F --+ Y with IITFll ~< 1 + 1/dim F and TFIE -- IE. By our hypotheses there is a separable subspace Z of X with Y C Z and Z has )~-bounded approximation property. Choose a finite rank operator L: Z --+ Z so that IILI[ ~< )~ and LIE = IE. Now consider the net SF : X --+ L Z of (non-linear and discontinuous) functions defined by
LTFx SFX --
0
x 6F, otherwise,
where the direction on the net is by inclusion on F. Since dim L Z < oc, a compactness argument yields a pointwise convergent subnet to say Lx for each x 6 X. Then, L is linear, llLll ~< )~, L has finite rank, and LIE = IE. The zrz-property is done similarly. For the uniform approximation property, the if part can be done as in the bounded approximation property case. For the only if part, choose a separable subspace Y of X and let (Yn) be dense in Sy with each Yn repeated countably many times in the sequence. Let (En) be the spans of all finite subsets of (yn). For every n, choose an operator Tn with IlTn II ~< K, rank Tn <~f ( d i m E n ) (where f ( n ) is the uniformity function for X) and so that 11(I - Tn)lEn II ~< 1/dimEn. Let Y1 = span Un Tn(X). Repeat this procedure on Y1 to get Y2 and so on. Let Z = Ui Yi. Then Y C Z, Z is separable and Z has the uniform approximation property. D Since all the approximation properties except the basis property and the finite dimensional decomposition property pass to complemented subspaces, Theorem 9.7 holds trivially for all these properties in the case that X is reflexive by Theorem 9.6. Another simple
Approximation Properties
313
consequence of Lindenstrauss' result is [18] that every weak Hilbert space X can be decomposed as X = H @ Y, where Y is a separable weak Hilbert space and H is isomorphic to a Hilbert space.
References [ 1] E.M. Alfsen and E.G. Effros, Structure in real Banach spaces L Ann. of Math. 96 (1972), 98-128. [2] G. Allexandrov, D. Kutzarova and A. Plichko, A separable space with no Schauder decomposition, Preprint. [3] D.E. Alspach and N.L. Carothers, Constructing unconditional finite-dimensional decompositions, Israel J. Math. 70 (1990), 236-256. [4] D.E. Alspach, E Enflo and E. Odell, On the structure of separable s (1 < p < cx~), Studia Math. 60 (1977), 79-90. [5] S. Banach, Theorie des Operations Lineaires, Warszawa (1932). [6] Y. Benyamini, P. Flinn and D.R. Lewis, A space without 1-unconditional basis which is 1-complemented in a space with a 1-unconditional basis, Longhorn Lecture Notes, Univ. of Texas (1983-1984), 145-149. [7] P.G. Casazza, James' quasi-reflexive space is primary, Israel J. Math. 26 (3-4) (1977), 294-305. [8] P.G. Casazza, Finite dimensional decompositions in Banach spaces, Cont. Math. 52 (1986), 1-31. [9] P.G. Casazza, The commuting bounded approximation property for Banach spaces, London Math. Soc. Lecture Notes 138 (1989), 108-128. [ 10] P.G. Casazza, C.L. Garcfa and W.B. Johnson, An example of an asymptotically hilbertian space which fails the approximation property, Preprint. [11] P.G. Casazza and H. Jarchow, Self-induced compactness in Banach spaces, Proc. Royal Soc. Edinburgh 26A (1996), 355-362. [12] P.G. Casazza, W.B. Johnson and L. Tzafriri, On Tsirelson's space, Israel J. Math. 47 (1984), 81-98. [13] P.G. Casazza and N.J. Kalton, Notes on approximation properties in separable Banach spaces, Lecture Notes of the London Math. Soc. 158 (1991), 49-65. [14] EG. Casazza and N.J. Kalton, Unconditional bases and unconditional finite-dimensional decompositions in Banach spaces, Israel J. Math. 95 (1996), 349-373. [15] EG. Casazza, N.J. Kalton and L. Tzafriri, Decompositions of Banach lattices into direct sums, Trans. Amer. Math. Soc. 304 (1987), 771-800. [16] P.G. Casazza, N.J. Kalton and E Wojtaszczyk, The commuting bounded approximation property for nonseparable Banach spaces, Unpublished notes. [17] EG. Casazza and E. Odell, Tsirelson's space and minimal subspaces, Longhorn Lecture Notes, University of Texas at Austin (1982-1983), 61-73. [18] P.G. Casazza and T. Shura, Tsirelson's Space, Springer Lecture Notes 1363 (1989). [ 19] C. Cho and W.B. Johnson, A characterization of subspaces X of g.p for which K (X), is an M-ideal in L (X), Proc. Amer. Math. Soc. 93 (1985), 466-470. [20] A.M. Davie, The approximation problem for Banach spaces, Bull. London Math. Soc. 5 (1973), 261-266. [21] A.M. Davie, The Banach approximation problem, J. Approx. Theory 13 (1975), 392-394. [22] P.G. Dixon, Left approximate identities in the Banach algebras of compact operators, Proc. Royal Soc. Edinburgh 104 (1986), 169-175. [23] P. Enflo, A counterexample to the approximation property for Banach spaces, Acta Math. 130 (1973), 309317. [24] M. Feder, On subspaces of spaces with a unconditional basis and spaces of operators, Illinois J. Math. 24 (1980), 196-205. [25] T. Figiel, Further counterexamples to the approximation problem, Unpublished notes. [26] T. Figiel, Factorization of compact operators and applications to the approximation problem, Studia Math. 45 (1973), 191-210. [27] T. Figiel and W.B. Johnson, The approximation property does not imply the bounded approximation property, Proc. Amer. Math. Soc. 41 (1973), 197-200.
314
P. G. Casazza
[28] T. Figiel and W.B. Johnson, A uniformly convex Banach space which contains no ~p, Compositio Math. 29 (1974), 179-190. [29] E Flinn, On a theorem of N.J. Kalton and G.V. Wood concerning 1-complemented subspaces of spaces having an orthonormal basis, Longhorn Notes, Univ. of Texas (1983-1984), 135-144. [30] G. Godefroy and N.J. Kalton, Approximating sequences and bidual projections, Quart. J. Math. 48 (2) (1997), 170-202. [31] G. Godefroy and P.D. Saphar, Three-space problems for the approximation properties, Proc. Amer. Math. Soc. 105 (1) (1989), 70-75. [32] Y. Gordon and D.R. Lewis, Absolutely summing operators and local unconditional structure, Acta Math. 133 (1974), 27-48. [33] W.T. Gowers and B. Maurey, The unconditional basic sequence problem, J. Amer. Math. Soc. 6 (1993), 851-874. [34] N. Gronbaek and G.A. Willis, Approximate identities in the Banach algebras of compact operators, Canadian Math. Bull. 36 (1993), 45-53. [35] A. Grothendieck, Produits tensoriels topologiques et espaces nucleires, Mem. Amer. Math. Soc. 16 (1953). [36] P. Harmand and A. Lima, Banach spaces which are M-ideals in their biduals, Trans. Amer. Math. Soc. 283 (1983), 253-264. [37] S. Heinrich, Ultraproducts in Banach space theory, J. Reine Angew. Math. 313 (1980), 72-104. [38] R.C. James, Separable conjugate spaces, Pacific J. Math. 10 (1960), 563-571. [39] W.B. Johnson, Finite-dimensional Schauder decompositions in 7v~ and dual 7r~ spaces, Illinois J. Math. 14 (1970), 642-647. [40] W.B. Johnson, Factoring compact operators, Israel J. Math. 9 (1971), 337-345. [41 ] W.B. Johnson, On the existence of strongly series summable Markuschevich bases in Banach spaces, Trans. Amer. Math. Soc. 157 (1971), 481-486. [42] W.B. Johnson, A complementably universal conjugate Banach space and its relation to the approximation problem, Israel J. Math. 13 (1972), 301-310. [43] W.B. Johnson, A reflexive Banach space which is not sufficiently Euclidean, Studia Math. 55 (1976), 201205. [44] W.B. Johnson, Banach spaces all of whose subspaces have the approximation property, Seminar d'Analyse fonct. Expose 16, Ecole Polytechnique (1979-1980); Special Topics of Applied Mathematics, NorthHolland, Amsterdam (1980), 15-26. [45] W.B. Johnson, Complementably universal separable Banach spaces: An application of counterexamples to the approximation property, Studies in Functional Analysis, R.G. Battle, ed., MAA Studies in Mathematics, Vol. 21 (1980). [46] W.B. Johnson, J. Lindenstrauss and G. Schechtman, On the relation between several notions of unconditional structure, Israel J. Math. 37 (1980), 120-129. [47] W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84.
[48] W.B. Johnson and T. Oikhberg, Separable lifting property and extensions of local reflexivity, Preprint. [49] W.B. Johnson and G. Pisier, The proportional u.a.p, characterizes weak Hilbert spaces, J. London Math. Soc. 44 (1991), 525-536. [50] W.B. Johnson, H.E Rosenthal and M. Zippin, On bases, finite dimensional decompositions, and weaker structures in Banach spaces, Israel J. Math. 9 (1971), 488-504. [51] W.B. Johnson and H.E Rosenthal, On co*-basic sequences and their applications in the study of Banach spaces, Studia Math. 43 (1972), 77-92. [52] E Jones, BMO and the Banach space approximation problem, Amer. J. Math. 107 (4) (1985), 853-893. [53] N.J. Kalton, M-ideals of compact operators, Illinois J. Math. 37 (1993), 147-169. [54] N.J. Kalton and N.T. Peck, Twisted sums of sequences spaces and the three space problem, Trans. Amer. Math. Soc. 255 (1979), 1-30. [55] N.J. Kalton and G.V. Wood, Orthonormal systems in Banach spaces and their applications, Math. Proc. Cambridge Philos. Soc. 79 (1976), 493-510. [56] R. Komorowski, On constructing Banach spaces with no unconditional basis, Proc. Amer. Math. Soc. 120 (1994), 101-109.
Approximation Properties
315
[57] R. Komorowski and N. Tomczak-Jaegermann, Banach spaces without local unconditional structure, Israel J. Math. 89 (1995), 205-226. [58] J.L. Krivine, Sous espaces de dimension fnite des espaces de Banach reticules, Ann. of Math. 104 (1976), 1-29.
[59] S. Kwapiefi, Isomorphic characterizations of inner product spaces by orthogonal series with vector valued coefficients, Studia Math. 44 (1972), 583-595. [60] J. Lindenstrauss, On certain subspaces ofs Bull. Acad. Polon. Sci. 12 (1964), 539-542. [61 ] J. Lindenstrauss, Extensions of compact operators, Mem. Amer. Math. Soc. 48 (1964). [62] J. Lindenstrauss, On non-separable reflexive Banach spaces, Bull. Amer. Math. Soc. 72 (1966), 967-970. [63] J. Lindenstrauss, On James' paper "separable conjugate spaces", Israel J. Math. 9 (1971), 279-284. [64] J. Lindenstrauss and L. Tzafriri, On the complemented subspaces problem, Israel J. Math. 9 (1971), 263269. [65] J. Lindenstrauss and L. Tzafriri, The uniform approximation property in Orlicz spaces, Israel J. Math. 23 (1976), 142-155. [66] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I, Sequence Spaces, Springer, Berlin (1977). [67] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces II, Function Spaces, Springer, Berlin (1979). [68] W. Lusky, A note on Banach spaces containing cO or C ~ , J. Funct. Anal. 62 (1985), 1-7. [69] W. Lusky, On Banach spaces with the commuting bounded approximation property, Arch. Math. (Basel) 58 (6) (1992), 568-574. [70] W. Lusky, On Banach spaces with bases, J. Funct. Anal. 138 (2) (1996), 410-425. [71 ] E Mankiewicz and N.J. Nielsen, A superreflexive Banach space with a finite dimensional decomposition so that no large subspace has a basis, Israel J. Math. 70 (1990), 188-204. [72] P. Mankiewicz and N. Tomczak-Jaegermann, Schauder bases in quotients of subspaces of s American J. Math. 116 (1994), 1341-1363. [73] P. Mankiewicz and N. Tomczak-Jaegermann, Structural properties of weak cotype 2 spaces, Canadian J. Math. 48 (1996), 607-624. [74] V. Mascioni, Some remarks on the uniform approximation property for Banach spaces, Studia Math. 96 (1990), 243-253. [75] V. Mascioni, On the duality of the uniform approximation property in Banach spaces, Illinois J. Math. 35 (1991), 191-197. [76] B. Maurey and G. Pisier, Series de variables aleatoires vectorielles independantes et proprietes geometriques des espaces de Banach, Studia Math. 58 (1976), 45-90. [77] N.J. Nielsen, The positive approximation property of Banach lattices, Israel J. Math. 62 (1988), 99-112. [78] N.J. Nielsen and N. Tomczak-Jaegermann, Banach lattices with property (H) and weak Hilbert spaces, Illinois J. Math. 36 (1992), 345-371. [79] N.J. Nielsen and E Wojtaszczyk, A remark on Bases in 12p-spaces with an application to complementably universal 12~-spaces, Bull. de l'Acad. Polon. Sci. 23 (1973), 249-254. [80] A. Pe~czyfiski, Any separable Banach space with the bounded approximation property is a complemented subspace of a Banach space with a basis, Studia Math. 40 (1971), 239-242. [81] A. Petczyfiski and H.E Rosenthal, Localization techniques in Lp-spaces, Studia Math. 52 (1975), 263-289. [82] A. Petczyfiski and E Wojtaszczyk, Banach spaces with finite dimensional expansions of identity and universal bases of finite dimensional subspaces, Studia Math. 40 (1971), 91-108. [83] G. Pisier, Counterexample to a conjecture ofGrothendieck, Acta Math. 151 (1983), 180-208. [84] G. Pisier, Factorization of Linear Operators and Geometry of Banach Spaces, C.B.M.S. Amer. Math. Soc., Vol. 60 (1985). [85] G. Pisier, The Volume of Convex Bodies and Banach Space Geometry, Cambridge Tracts in Math. 94, Cambridge University Press (1989). [86] C.J. Read, Different forms of the approximation property, Unpublished notes. [87] H.P. Rosenthal, On one complemented subspaces of Banach spaces with a one unconditional basis, according to Kalton and Wood, Israel Seminar in the Geometrical Aspects of Functional Analysis, Univ. of Tel Aviv (IX) (1983-1984), 1.1-1.16. [88] C. Samuel, Exemples d'espaces de Banach ayant la propriete de projection uniforme, Colloq. Math. 43 (1980), 117-126.
316
P. G. Casazza
[89] C. Samuel, Bounded approximate identities in the algebra of compact operators on a Banach space, Proc. Amer. Math. Soc. 177 (1993), 1093-1096. [90] J. Schauder, Zur Theorie stetiger Abbildungen in Funkionalraumen, Math. Zeit. 26 (1927), 47-65, 417-431. [91] A. Szankowski, A Banach lattice without the approximation property, Israel J. Math. 24 (1976), 329-337. [92] A. Szankowski, Subspaces without the approximation property, Israel J. Math. 30 (1978), 123-129. [93] A. Szankowski, B(H), does not have the approximation property, Acta Math. 146 (1981), 89-108. [94] A. Szankowski, On the uniform approximation property in Banach spaces, Israel J. Math. 49 (1984), 343359. [95] S.J. Szarek, A Banach space without a basis which has the bounded approximation property, Acta Math. 159 (1987), 81-98. [96] N. Tomczak-Jaegermann, Banach-Mazur Distances and Finite-Dimensional Operator Ideals, Pitman Monographs, Longman Scientific and Technical, New York (1989). [97] B.S. Tsirelson, Not every Banach space contains g.p or co, Funct. Anal. Appl. 8 (1974), 138-141. [98] G.A. Willis, The compact approximation property does not imply the approximation property, Studia Math. 103 (1992), 99-108. [99] M. Zippin, The embedding of Banach spaces into spaces with structure, Illinois J. Math. 34 (3) (1990), 586-606.
CHAPTER
8
Local Operator Theory, Random Matrices and Banach Spaces
Kenneth R. Davidson Department of Pure Mathematics, University of Waterloo, Waterloo, Ontario, Canada N2L 3G1 E-mail: krdavids @math. uwaterloo, ca
Stanislaw J. Szarek Equipe d'Analyse, University Paris VI, Case 186, 4, place Jussieu, 75252 Paris, France Department of Mathematics, Case Western Reserve University, Cleveland, 0H44106-7058, USA E-mail: sjsl 3 @po.cwru.edu
Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
319
1. Local operator theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
321
1.1. A l m o s t c o m m u t i n g H e r m i t i a n matrices 1.2. Unitary orbits of normal matrices 1.3. Quasidiagonality
.................................
....................................
.............................................
1.4. Extensions of pure states and matrix paving . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5. Hyper-reflexivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. R a n d o m matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
321 326 330 333 338 341
2.1. T h e overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Concentration of m e a s u r e and its c o n s e q u e n c e s ............................ 2.3. N o r m of a r a n d o m matrix and the S l e p i a n - G o r d o n l e m m a . . . . . . . . . . . . . . . . . . . . . . .
342 346 350
2.4. R a n d o m matrices and free probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5. R a n d o m vs. explicit constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
354 358
Acknowledgments References
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
360
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
360
H A N D B O O K O F T H E G E O M E T R Y O F B A N A C H SPACES, VOL. 1 Edited by William B. J o h n s o n and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 317
This Page Intentionally Left Blank
Local operator theory, random matrices and Banach spaces
319
Introduction Connections between Banach space theory and classical operator theory on Hilbert space are multifold. First, one can generalize notions and results involving linear operators on a Hilbert space to the Banach space context. Second, more often than not, the study of the latter area also involves linear operators, and so one can use methods developed in one of the fields to attack questions in the other. Thirdly, spaces of operators are typically Banach spaces, and so we may study them as such using the methods of geometry of Banach spaces. Of course, this is not a complete list of connections. In this survey we shall focus on the second and the third of the links hinted above, and on those of their aspects which are related to the local theory of Banach spaces and to the asymptotic properties of finite dimensional normed spaces, as the dimension approaches infinity. More specifically, we shall address the following two groups of issues; the corresponding parts of the article (Sections 1 and 2) can be read essentially independently. 1. Problems in operator theory which are local in the sense that they reduce to questions about finite matrices, and moreover the resulting questions are similar in flavor to ones studied in the Banach space context. For example, they might involve estimates independent of the dimension of the space on which the matrices act. Or, they may require the determination of approximate isomorphic or almost isometric asymptotics for some quantity as a function of the dimension. At the same time, the estimates needed are usually not optimal, a feature that would be characteristic to a representative Banach algebra or C*-algebra problem. 2. The asymptotic behavior of large dimensional random matrices, with particular attention to their spectral properties. This is a topic which has been of major interest both in the geometry of Banach spaces and in the theory of operator algebras. It is also related to numerous other fields, and has lately become a very hot subject in view of apparent analogies between the distribution of eigenvalues of large random matrices and of zeros of the Riemann ff function. This article will not focus on these connections, but will mention them where appropriate. More generally, we shall bring up examples where the methods of one of the fields appear to have relevance to the other. This relevance may be described by a theorem, or just hinted by a vague analogy. As these points of contact between Banach space theory and operator theory are known to relatively few people on both sides, we hope that popularizing them will give a boost to both areas. Before proceeding, we shall clarify what we mean by isomorphic, almost isometric and isometric estimates through this simple example. Let Tn denote the triangular truncation map on the n x n matrices which sends a matrix A = [aij ].n~,j--1 to "-l-nA -- [bij ] where
bij
_ J aij I0
i f / ~< j , ifi > j .
Suppose that we are interested in the behavior of IlTn I1 as n tends to infinity. Several answers of varying degrees of precision (respectively isomorphic, almost isometric and isometric) are possible:
K.R. Davidson and S.J. Szarek
320
(a) lien II is of order log n, meaning that there are universal constants c, C > 0 such that for all n, c log n ~< IIT~ II ~< C log n. (b) IIT~ II is equivalent to ~1 log n [8], meaning that IIT~ II
1
lim = --, n--+~ log n Jr
(c) liT. II = f(n), where f is an explicit function. We shall also say that (a) and (b) describe, respectively, the rough and precise asymptotic order of IIT~ II. As above, and unless indicated otherwise, c, cl, C, C t etc. will always stand for universal positive constants whose numerical values may vary between occurrences. We now describe those topics which we will address in this article. In Section 1, we will treat questions of local operator theory focusing on the following areas: 9 Almost commuting versus nearly commuting problems: given a pair of (e.g., selfadjoint) matrices whose commutant is small in norm, is it close to a pair of (selfadjoint) matrices which exactly commute? 9 Spectral distance problems: given two matrices (normal, unitary, etc.) which are close in norm, are their spectra close in some appropriate sense? 9 Approximation of large matrices by direct sums of smaller ones: for example, the Herrero-Szarek result on non-reducibility of matrices and the examples of Szarek and Voiculescu of exotic quasidiagonal operators. 9 The Kadison-Singer problem (reformulated as the paving problem): does there exist a positive integer k such that, given a square matrix A with zero diagonal, it is always possible to find diagonal projections P1, P2 . . . . . Pk such that k
k
and i=1
~
Pi APi
1
~IIAII?
i=1
9 Questions about hyper-reflexivity: in particular, do Tn | Tn, the tensor products of two copies of the algebra of n • n upper triangular matrices, have a common distance estimate analogous to Arveson's distance formula for nests? We shall not elaborate on the Sz. Nagy-Halmos similarity problem recently solved by Pisier, see [138] or [64]. Even though it conforms quite well to the spirit of this article, it is, in its heart, a complete boundedness/operator space question. We do not embark on those areas as both them and the similarity problem itself are well covered in the literature, cf. [134,140,139] and (in this collection) [141] for the former and [61] for the latter. On the other hand, we shall mention other less known questions about finite matrices that fit into our framework. In each case, we shall try to sketch the connections to infinite operator theory. In particular, the above problems relate respectively to the Brown-Douglas-Fillmore theory, similarity and perturbations of operator algebras, approximation and distance estimates from spectral data, quasidiagonality, and the uniqueness of extensions of pure states. Depending on whether a problem has been solved or not, we shall either hint at the ingredients of the proof or describe the current state of the art and the consequences for the original problem.
Local operator theory, random matrices and Banach spaces
321
We point out that, in addition to their significance for the infinite dimensional theory, many of the problems stated above are also clearly important from the numerical linear algebra point of view. In Section 2, we shall signal various ways in which random matrices interact with operator theory and the geometry of Banach spaces. We start with an overview of the subject by presenting a sampler of classical (and not so classical) results on asymptotic spectral properties of large random matrices. We shall show how some of these classical results can be rather routinely obtained using standard methods of Banach space theory, e.g., by exploiting the measure concentration phenomenon and various tools of probability in Banach spaces. Although some of these arguments are folklore in the Banach space circles, they are generally not known to the wider mathematical public. Our purpose is to just illustrate this approach to the topic of random matrices, and so we shall not aim at the strongest or most general results that could be so obtained. We shall also sketch the links between the subject of random matrices and Voiculescu's free probability and state some other subtle and interesting questions (mostly of the almost isometric nature) about norms and similar parameters of random matrices. These questions are not necessarily directly relevant to Banach space theory, but conceivably they can be approached with the aid of standard tools of the area. We feel that it would be a worthwhile project to review and unify the closely related random matrix results and rather dissimilar methods pertaining to those results coming from, among others, mathematical physics, probability and Banach space theory. In addition to clarifying the picture, the benefits could include stating results in a form immediately utilizable in other fields. This is often a problem. For example, large deviation and mathematical physics formulations frequently involve rescaling the quantities in question and then investigating their asymptotics as the dimension tends to c~. On the other hand, applications to the geometry of Banach spaces, convexity and computational complexity typically require estimates that are valid for a wide range of parameters in any given dimension. Finally, in the last section, we mention some of the technologies used in operator theory to exhibit phenomena analogous to the ones obtained in the Banach space theory via random methods. These technologies typically involve representation theory, but there are also links to graph theory and arithmetic geometry, to name a few.
1. Local operator theory We will discuss this area issues through a few important problems which are in various states of solution, as mentioned in the introduction.
1.1. A l m o s t commuting Hermitian matrices There has been dramatic progress made on the following matrix problem: PROBLEM 1.1. Given e > 0, is there a 6 > 0 so that: whenever A and B are n x n Hermitian matrices of norm one, for any n >~ 1, such that ]lAB - BA]l < 6, then there are Hermitian matrices A1 and B1 which exactly commute such that IIA - A lll < e and lib - B1 II < ~?
322
K.R. Davidson and S.J. Szarek
The condition IIAB - BAll < 6 colloquially says that the matrices A and B almost commute; while the conditions IIA - All] < e and IIB - B~ II < e for a commuting Hermitian pair A1 and B1 say that A and B nearly commute. So the question becomes:
Are almost commuting Hermitian matrices nearly commuting ? An interesting reformulation of the problem tums this into a question about normal matrices. The matrix T -- A + iB satisfies [T, T*] := TT* - T*T = - 2 i ( A B - BA). Thus the Hermitian pair {A, B } is almost commuting when T is almost normal, and exactly commuting when T is normal. Hence the pair is near to a commuting pair precisely when T is close to a normal matrix. Conversely, if T is an almost normal or nearly normal matrix, then the real and imaginary parts A = (T + T * ) / 2 and B = (T - T * ) / 2 i are almost commuting or nearly commuting, respectively. Thus another equivalent question is: Are almost normal matrices nearly normal? As with many of the problems we discuss in this article, the crucial point is the dimension-free character of the estimates. Indeed, the problem could be reformulated for compact Hermitian matrices. However that was not the appropriate way to go for the solution, as we shall see. Dimension dependent results were obtained early on by Pearcy and Shields [135] who establish that 6 = eZ/n will suffice. This was sharpened by the second author [158] to 6 = cel3/Z/x/-ff. We shall see however that these estimates are not nearly of the correct order. It is interesting to note that there is a negative answer to the analogue of Problem 1.1 for arbitrary operators on Hilbert space. Consider a weighted shift on ~2 with basis {ek: k ~> 1 } given by
Snek=
min{k, n}
ek+l
fork~>l.
It is easy to see that 1 [s.,s*~]--~P.
where Pn is the orthogonal projection onto span{el . . . . . en}. Thus for large n, this operator has small self-commutator; and thus its real and imaginary parts are almost commuting. However, every Sn is a compact perturbation of the unilateral shift $1. This operator is a proper isometry which is Fredholm of index - 1 and has essential spectrum tre(S1) -- 7s Therefore each Sn is also Fredholm of index - 1 with essential spectrum ere (Sn) = ql", as these properties are invariant under compact perturbations. It is also straightforward to show that any operator T with II$1 - T II < 1 has Fredholm index - 1. So again, this is also true for Sn. A normal operator N has index 0 because ker N = ker N*. Thus each Sn is at least distance 1 from a normal operator, and in fact this is exact. There are many variants of our problem, and most of them have turned out to have negative answers. Here is one example which has received a lot of attention. Fix a basis
Local operator theory, random matrices and Banach spaces
323
{ej" 0 <~ j <~ n} for C n+l, and consider the matrices 2j A e j -- ( 1 - --ff-)ej
for0 ~< j ~
2
Bej -- - f f - ~ / ( j 4- 1)(n -- j ) e j + |
for0 <~ j ~ < n - 1,
Ben = 0 . Easy computations show that A is Hermitian, II[B, B*]II < 4 / n and II[A, B]II < 2 / n . Voiculescu [ 173] showed that the triple {A, Re B, Im B } are almost commuting but not near to a commuting triple of Hermitian matrices. A refinement of this by the first author [55] showed that {A, B} was not close to a commuting pair {A', B'} where A' is self-adjoint but B' is not required to be normal. Finally Choi [51 ] showed that this pair is far (about 0.5) from any commuting pair with no conditions on A' or B'. Choi's obstruction involved the determinant of a polynomial in A and B. Curiously, the matrix B above is close to a normal matrix; and the pairs {A, Re B } and {A, ImB} are also near commuting. This was shown in [55] using a method introduced by Berg [28]. Berg developed a technique for perturbing direct sums of weighted shifts that are almost normal to obtain normal operators when no index obstruction prevents it, as in the example cited above. This was a special case of a deep problem in operator theory. This leads to a connection with the Brown-Douglas-Fillmore theory [47,48] of essentially normal operators. An operator T is essentially normal if [T, T*] is compact. The original operator theory question they analyzed was to find when there is a compact perturbation of T which is normal. It is evident that if T - )~I is Fredholm, then the Fredholm index must be zero. The operator T has image t = 7r(T) in the Calkin algebra s The Calkin algebra is a C*-algebra, and t is a normal element of it. Thus C* (t) is isomorphic to C(X) where X = o-(t) = ae(T). For each bounded component Ui of C \ o-(t), the index r t i = ind(t - )~i) is independent of the choice of a point )~i E Ui. A normal operator will have ni = 0 for every 'hole'. And it is possible to construct an essentially normal operator T with ae(T) = X and any prescribed indices r / i - - ind(T - )~i I ) . The BrownDouglas-Fillmore theorem establishes that these are the only obstructions. THEOREM 1.2 (Brown-Douglas-Fillmore). Suppose that T is an essentially normal operator Then there is a compact operator K such that T - K is normal if and only if ind(T - )~I) -- Of o r every )~ in C \ ae(T). Surprisingly, the proof of BDF did not follow usual operator theoretic lines. Instead they observed that for any operator T above, there is a short exact sequence 0
>/C
i
> C*(T)+/C
7/"
> C(X)
> O.
Thus T determines an extension of the compact operators by C(X). Conversely any such extension determines, up to a compact perturbation, an essentially normal operator such that 7r(T) = z, where z is the identity function on X. A third viewpoint identifies this extension with the monomorphism that identifies C(X) with the subalgebra C* (t) of the Calkin algebra. They put a natural equivalence on extensions which corresponds to identifying operators which are unitarily equivalent up to a compact perturbation. The set of
K.R. Davidson and S.J. Szarek
324
equivalence classes Ext(X) becomes a semigroup under the operation of direct sum. Moreover this construction makes sense for any compact metric space X, not just subsets of the plane. They established that, in fact, Ext(X) is a group. Moreover it is a generalized homology theory paired in a natural way with topological K-theory. This opened the door to a new era in C*-algebras in which K-theory and topological methods came to play a central role. Even the existence of a zero element is non-trivial. In the case of a single operator or planar X, this reduces to the Weyl-von Neumann-Berg theorem [27] which shows that every normal operator is a small compact perturbation of a diagonalizable operator. This is easily generalized to arbitrary metric spaces, and it shows that any representation of C(X) on Hilbert space is close in an appropriate sense to a representation by diagonal operators. BDF show by elementary methods that any essentially normal operator T is unitarily equivalent to a small compact perturbation of T 9 D, where D is a diagonal normal operator with a ( D ) = ae(D) = ae(T). This leads us somewhat afield from the original problem. However the connection is made via quasidiagonality. A set of operators T is quasidiagonal if there is an increasing sequence Pk of finite rank projections converging strongly to I such that l i m n ~ II[T, P]II = 0 for every T E T. In particular, if T is quasidiagonal and e > 0, then there is a sequence Pk such that
T-- Z(Pk-- Pk-1)T(Pk-- Pk-1)= K k>/1
is compact and IlK II < e. Indeed, one just drops to an appropriate subsequence of the original one. Thus T - K has the form ~ke> 1 Tk where T~ act on finite dimensional spaces. Such operators are called block diagonal. Salinas [149] showed that the closure of the zero element in Ext(X) in the topology of pointwise-norm convergence, where we think of an extension as a monomorphism from C(X) into/2(7-{)//(7, is the set of all quasidiagonal extensions. In particular, unlike the planar case, the trivial element need not be closed. He used this to establish homotopy invariance of the Ext functor. In the case of a single operator, this shows that if T is essentially normal with zero index data, then it has a small compact perturbation which is block diagonal, say
T-K-T'-
ZeTk. k>~l
Moreover, it follows that [T', T'*] -- y~ke> 1[Tk, Tk*]. Therefore lim
k---+ (x)
II t:Tk, :T;J II
-
0
A positive solution to our original question would yield normal matrices Nk such that limk~ IIT~ - N~ II - 0. Hence N -- Y~ke>1 Nk is a normal operator such that T - N is compact. This was the motivation for the work of the first author [55]. If T -- y~,e k~>l Tk is essentially normal with ae(T) = X, then after a small compact perturbation, T may be replaced
Local operator theory, random matrices and Banach spaces
325
with T G D where D is any normal operator with eigenvalues in X. Thus each summand Tk can be replaced by Tk 9 D~, where Dk is diagonal. An absorption theorem is proven: THEOREM 1.3 (Davidson). I f T is any n • n matrix, then there are an n x n n o r m a l matrix N a n d an 2n • 2n n o r m a l matrix M such that
IIM - Z 9 Nil ~< 75[1[T, Z*]l[ 1/2. One difficulty in applying this to the BDF problem is that any holes in the spectrum of T are obliterated by summing with a normal with eigenvalues dense in the unit disk. In [29], Berg and the first author showed that similar operator-theoretic techniques can be used to show first that every essentially normal operator with zero index data is quasidiagonal. And then refinements of the absorption principle which control the spectrum were used to prove the planar version of BDF. This had the advantage of providing quantitative estimates for 'nice' spectra. But the correct positive answer to our question came again from a more abstract approach. Huaxin Lin [113] considered T = ~ke>~l Tk as an element of the von Neumann algebra 93l -- Ilk~>l 9)tnk. A positive solution is equivalent to constructing a normal operator N = Y'~ke>1Nk asymptotic to T, that is, verifying l i m k ~
[ITk -- Nk ] l - 0. This is a per-
turbation by the ideal 3 -- ~ k ~>1 93ln~ of elements J = ~ ek/> 1 Jk where limk~ oc I[Jk 11= 0. Thus T determines an element t = q ( T ) in the quotient algebra 9Jt/~. The problem is reduced to lifting each normal element of 93I/3 to a normal element of 93t. This was accomplished by a long tortuous argument. THEOREM 1.4 (Lin). Given e > O, is there a ~ > 0 so that w h e n e v e r A a n d B are n x n H e r m i t i a n matrices o f n o r m one, f o r any n >~ 1, a n d ]lAB - BAll < 6, then there are H e r m i t i a n matrices A1 a n d B1 w h i c h exactly c o m m u t e such that ]IA - All] < e a n d
IIB - B1 II < ~. A remarkable new proof of Lin's theorem is now available due to Friis and ROrdam [78]. They take the same approach, but base their argument only on two elementary facts about 93l/3. It has stable rank one, meaning that the invertible elements are dense. And every Hermitian element can be approximated by one with finite spectra. Both of these results can be obtained by using well-known facts about matrices on finite dimensional spaces. From this, a short argument cleverly using no more than the basic functional calculus yields the desired lifting of normal operators. Thus the original question has a positive solution. (See Loring's book [ 115] for a C*-algebraic view of this problem and its solution.) In a second paper [79], they show how the planar case of the Brown-Douglas-Fillmore Theorem follows. An important matrix question remains: PROBLEM 1.5. What is the dependence of 3 on e for almost commuting Hermitian matrices?
K.R. Davidson and S.J. Szarek
326
Optimal estimates for 3 should be 0(82) as in the absorption results. However the Lin and Friis-Rcrdam results yield only qualitative information, and do not provide information on the nature of the optimal 6 function. A related question deals with pairs of almost commuting pairs of unitary matrices. Voiculescu [173] constructed the pair Uej --ej+lmodn and Vej --e2zrij/nej for 1 ~< j ~< n. It is evident that [I[U, V]ll < 2rc/n. However he showed that this pair is not close to a commuting pair of unitaries. He speculated that both this example and the commuting triples example occur because of topological obstructions. Asymptotically, the spectrum approximates a torus (or a two-sphere for Hermitian triples), and thus there is a heuristic justification for this viewpoint. Loring [ 114] showed that this was indeed the case for this pair of unitaries by establishing a K-theoretic obstruction coming from the non-trivial homology of the two-sphere. Exel and Loring [74] obtained an elementary determinant obstruction that shows that this pair of unitaries is far from any commuting pair, unitary or not. 1.2. Unitary orbits of normal matrices Consider a normal matrix N in 9Xk. By the spectral theorem, there is an orthonormal basis which diagonalizes N. So N is unitarily equivalent to diag(~,l . . . . . )~k) where the diagonal entries form an enumeration of the spectrum of N including multiplicity. Two normal matrices are unitarily equivalent precisely when they have the same spectrum and the same multiplicity of each eigenvalue. A natural question from numerical analysis and from the approximation theory of operators asks how the spectrum can change under small perturbations. The answer one gets depends very much on what kind of perturbations are allowed. In particular, if the perturbed matrix is also normal, then one obtains a more satisfying answer than if the perturbation is arbitrary. Consider first the case of a Hermitian matrix A. Then the eigenvalues may be enumerated so that ,kl ~> ... ~> ~.k. If B is a small perturbation of A, say IIA - B 1[ < e, then the Hermitian matrix R = (B + B*)/2 also satisfies IIA - R[[ < e, and so we may assume that the perturbation B is Hermitian to begin with. In that case, write the eigenvalues as # l ~> " " ~> #k. Then a 1912 argument due to Weyl [183] gives: THEOREM 1.6 (Weyl). Let A and B be k x k Hermitian matrices with eigenvalues ordered as )~l ~ "" >1 )~k and # l ~ "" >~ #k, respectively. Then sup [#j - ,kj [ ~< [] A - B II. l~j~k
Moreover, f o r each A, this value is attained by some such B. Indeed, suppose for example that )~j ~ # j . Consider the spectral subspace for A corresponding to {)~1. . . . . ~j } and the spectral subspace of B for {#j . . . . . #k }. Since the dimensions of these spaces add up to more than k, they contain a common unit vector x. Then (Ax, x) lies in the convex hull of {~1 . . . . . ~,j }, whence is ~> )~j. Similarly (Bx, x) <~ lZj. Hence
IIA -
BI[
~ ((A -
B ) x , x ) - ~(Ax, x) - (Bx, x) ~ ~,j - # j .
Local operator theory, random matrices and Banach spaces
327
Weyl's theorem shows that a perturbation of Hermitian matrices of norm e cannot move any eigenvalue more than e. One can reformulate this to say that if A and B are Hermitian with eigenvalues as above, then IIA - B II is bounded below by the spectral distance suPl <~j<~kI#j -- ~'j I. This spectral distance is achieved when A and B are simultaneously diagonalizable with eigenvalues matched in increasing order. The unitary orbit bt(A) of a matrix A is the set of all matrices which are unitarily equivalent to it. In finite dimensions, this is a closed set because the unitary group is compact. However, for operators this is no longer the case. The unitary orbit of a normal matrix consists of all normal matrices with the same spectral data. Thus one must be able to express the distance between these two unitary orbits solely in terms of the two spectra. Weyl's result shows that in the Hermitian case, this distance is precisely the spectral distance. More recently, significant attention has been paid to the more difficult normal case. Here is it no longer trivial to define a spectral distance. However, if M and N are normal matrices which are simultaneously diagonalizable, then this diagonalization amounts to a pairing of the two sets of eigenvalues. Among all possible permutations, there is one of least distance, and this is called the spectral distance: sp.dist(~.,/z) := inf
m a x [~,j yr CSk 1~ j ~ k
~Tr(j)[,
where ~ a n d / t represent the eigenvalues repeated according to their multiplicities. (This is really the quotient metric on ~ divided by the natural action of the symmetric group on cn.) Evidently, dist(L/(M), L/(N)) ~< sp.dist(a(M), a ( N ) ) , where a (M) denotes the spectrum of M including multiplicity. Many results in the literature (cf. [36,155]) show that under special hypotheses, the spectral distance is again the exact answer. See [33 ] for a full treatment of these ideas. However a recent result of Mueller (private communication) shows that even in three dimensions, it is possible that dist(L/(M), b/(N)) < sp.dist(a(M), a ( N ) ) . His example is explicit, but it is fair to say that we do not really understand why this strict inequality occurs. So it is of great interest that a result of Bhatia, Davis and McIntosh [35] obtains bounds independent of dimension: THEOREM 1.7 (Bhatia-Davis-McIntosh). that for all normal matrices M and N,
There is a universal constant c (> 1/3) such
dist(lg(M), lg(N)) ~ c sp.dist(a(M), a ( N ) ) . This result is obtained from a study of the Rosenblum operator on 9Ytk given by rM,N(X) = N X - X M . It is well known that this map is injective if a ( M ) and o'(N) are disjoint. In the case in which M and N are normal, there is a useful integral formula for the inverse which allows norm estimates. This integral formula requires finding a function f in LI(]R 2) with Fourier transform f satisfying f ( x , y) -- 1/(x + iy) for x 2 + y2 ) 1. The constant depends on 11f Ill, which may be chosen less than 3 [34].
328
K.R. Davidson and S.J. Szarek
Estimates of this type allow one to estimate the angle between spectral subspaces of nearby normal operators. For example, if E = EM(K) and F = EN(L) are the spectral projections of M and N respectively for sets K and L with dist(K, L) = 3, then
1 IIEFII ~ ~ I I A - nil. c~ When this is small, the ranges of E and F are almost orthogonal. These results can be extended to normal operators on infinite dimensional space. Here the unitary orbit of a normal operator is not closed. The unitary invariants are given by the spectral theorem in terms of the spectral measure and a multiplicity function. However, the Weyl-von Neumann-Berg theorem [27] shows that the closure of the unitary orbit of every normal operator contains a diagonal operator. The multiplicity and spectral measure do not play a role in this closure except for isolated eigenvalues. Indeed, two normal operators have the same unitary orbit if and only if they have the same essential spectrum, and the same multiplicity at each isolated eigenvalue. (See [60, Theorem 11.4.4].) It turns out that one can define an analogue of the spectral distance for spectra of normal operators. Once this is accomplished, the finite dimensional methods extend to show that the distance between unitary orbits is equal to the spectral distance for Hermitian operators [15] and greater than a universal constant times the spectral distance for normal operators [56]. Some of these results are valid for other norms on the set of matrices. Although we are usually interested in the operator norm because of its central role, other norms occur naturally. For example, the Schatten p-norms are defined as the gP norm of the (sequence of the) singular values. In particular, p = 2 yields the Hilbert-Schmidt norm I1" 112, sometimes also called the Frobenius norm, which is particularly tractable. More generally, there is a large class of norms which are unitarily invariant in the strong sense that II A II ~ = II U A V II for all unitaries U and V; for 9Ytn, there is a one-to-one correspondence between such norms and symmetric norms on C n (or ~n) applied to the sequence of singular values. For example, the Ky Fan norms are given by the sum of the k largest singular values. The set of (properly normalized) unitarily invariant norms is a closed convex family and the Ky Fan norms are the extreme points. One can define the spectral distance between the unitary orbits of two normal matrices with respect to any unitarily invariant norm by sp.dist~ (k, p,) " - inf Ildiag()~j - #~(j))II 7rESk
z"
(Again, this is really a quotient metric induced by the symmetric norm o n C n related to I1" II~.) A classical result for the Hilbert-Schmidt norm is due to Hoffman and Wielandt [96]. Note that there is no constant needed. THEOREM 1.8 (Hoffman-Wielandt). If M and N are normal matrices, then sp.dist2 (a (M), a ( N ) ) ~< I I M - NIl2.
Local operator theory, random matrices and Banach spaces
329
For general unitarily invariant norms, a result of Mirsky [ 125] extends Weyl's Theorem for Hermitian matrices. A more recent result of Bhatia [32] extends this to certain normal pairs. THEOREM 1.9 (Mirsky-Bhatia). Suppose that M and N are normal matrices such that M - N is also normal (e.g., M and N Hermitian). Then f o r every unitarily invariant norm, sp.distr (or(M), or(N)) <, i I M - NIIr. For unitary matrices, Bhatia, Davis and McIntosh [35] show that THEOREM 1.10. Suppose that U and V are unitary matrices. Then f o r every unitarily invariant norm, 2
--sp.distr
(or(U), o'(V)) ~< [ I U - Vlir.
Jr
More results along these lines may be found in Bhatia's books [31,33]. When matrices are not normal, there are generally no good estimates of distance in terms of their spectra. A simple example is to take an orthonormal basis el . . . . . en for C n, and set Jn to be the Jordan nilpotent Jnei = ei+l for 1 ~< i < n and Jnen 0. Then consider T -- Jn + ~(', en)el. We see that liT - J n [ I - E while cr(Jn) - {0} and or(T) consists of the nth roots of e. So the spectral distance is e 1/n. This is much more dramatic than the normal case. We refer the reader to the books [31,33 ] for a lot more material on these types of questions. Likewise when calculating the eigenvalues of a normal matrix numerically, we are dealing with its perturbation or approximation which is not necessarily normal. A small perturbation of a normal matrix may have spectrum which does not approximate the spectrum of the normal at all. Indeed, Herrero [91] shows that there is a normal matrix N in 9)tk with 1 E or(N) and a nilpotent matrix Q such that lIN - Qll < 5 k-1/2. This allowed him to establish an important infinite dimensional result: =
THEOREM 1.1 1 (Herrero). Suppose that N is a normal operator on Hilbert space with connected spectrum containing O. Then N is the norm limit o f a sequence o f nilpotent operators. One way to construct such examples is to take Q to be a weighted shift Qei --aiei+l for 1 ~< i < k and Q ek = 0. This is evidently nilpotent. However, if the weights slowly increase from 0 to 1 and back down again, it will follow that IIQ II = 1 and II Q* Q - Q Q* II is small on the order of O(k-1). We saw in the previous section that the matrix Q must be close to a normal matrix. For weighted shifts, a method known as Berg's technique [28] (cf. [91 ]) allows explicit perturbations to a normal matrix with estimates on the order of C II Q* Q - Q Q* IIl/R, which yields the desired example. Self-adjoint examples can be constructed, but the estimates are strikingly different. Hadwin [90] solved this problem by relating the problem of approximating a Hermitian matrix by a nilpotent one to the norm of triangular truncation, mentioned in the introduction.
330
K.R. Davidson and S.J. Szarek
THEOREM 1.12 (Hadwin). For k >>.1, let 6 ~ - i n f { l l A - QII" A, Q ~ 9Ytk, A - A*, IIAI] = 1, Q~ = 0 } . Then l i m k ~
6n logn = zr/2.
However, for the numerical analysis problem of calculating eigenvalues of a matrix which is a priori known to be normal, this deterioration of estimates is inessential. Indeed, if A0 is a given small but not necessarily normal perturbation of an unknown normal matrix A, and A1 is any normal approximant of A0, then by Theorem 1.7, the spectra of A and A 1 are close. However this argument is not constructive, as it does not tell us how to find A1. This leads to a refinement of Problems 1.1 and 1.5: given a matrix Ao which is known to be nearly normal, find a specific normal approximant with an explicit error estimate. In infinite dimensions, the description of those operators which are limits of nilpotents is a central result in the approximation theory of operators due to Apostol, Foia~ and Voiculescu [10], cf. [91, Chapter 5]. THEOREM 1.13. An operator T is the limit of a sequence of nilpotent operators if and
only if (i) T has connected spectrum and essential spectrum containing O, and (ii) whenever T - )~I is semi-Fredholm, the Fredholm index is O. These conditions are easily seen to be necessary by elementary means. The special case of normal limits mentioned above follows from the finite dimensional approximations and an application of the continuous functional calculus. However, the general results require a good model theory for a dense set of Hilbert space operators. This requires interplay between infinite dimensional methods and analytic function theory. We return to some considerations of this result in the next section.
1.3. Quasidiagonality The class of quasidiagonal operators and block diagonal operators were introduced in Section 1.1. A priori, one might expect that quasidiagonal operators behave more like matrices than arbitrary operators. However, in many ways, it has proven to be a difficult class of operators to deal with. Still, they do occur in important ways in both operator theory and C*-algebras. Here we will limit our attention to questions related to questions about matrices. The class of block diagonal operators of the form T = ~j(t3>/1 Tj which most closely mimics matrices are those in which the summands Tj are of uniformly bounded dimension, say m. Such operators, after a rearrangement of basis, can be written as a direct sum of k x k matrices with diagonal operator entries for 1 ~< k ~< m. The Weyl-von Neumann-Berg Theorem shows that every normal operator is the norm limit of diagonalizable operators. More generally, the same methods show that any finite set of commuting normal operators
Local operator theory, random matrices and Banach spaces
331
can be simultaneously approximated by diagonal operators. Thus the closure of the set of operators which are k x k matrices with diagonal entries is the set of k x k matrices with commuting normal entries. These operators are called k-normal. From a more algebraic point of view, the C*-algebra generated by an m-normal operator has the property that there is a family of irreducible representations of dimension at most m which separate points in the algebra. Conversely this condition implies that the operator is the direct sum of k-normal operators for k ~< m. In this section, we will concern ourselves with a variety of questions dealing with the approximation of operators in the quasidiagonal class. There are also many connections with C*-algebras which will have to be neglected. Consider first the question of whether every quasidiagonal operator can be approximated by m-normal operators with m unlimited. The first attempt would be to try to approximate large finite matrices by a direct sum of smaller matrices. This works very well for weighted shifts using Berg's technique [28]. This method has been exploited by Herrero [93] in obtaining good estimates for the distance between unitary orbits of weighted shifts. With Berg, the first author [29] expanded this method to work for general block tri-diagonal forms in their proof of a quantitative Brown-Douglas-Fillmore Theorem [47,48]. Further evidence comes from the infinite dimensional setting where Voiculescu's celebrated generalized Weyl-von Neumann Theorem [172] implies, among other things, that every operator is the norm limit of operators with many reducing subspaces. Indeed, every operator T is an arbitrarily small compact perturbation of an operator unitarily equivalent to Tt = T G A G A | where A is the image of T under a ,-representation of C* (T) which annihilates K~(7-/) n C* (T). This is especially good since T and T t have the same closed unitary orbit (whence T ~ is also the limit of operators unitarily equivalent to T). In spite of this positive evidence, the desired decomposition of large matrices into smaller ones is not possible [95]. THEOREM 1.14 (Herrero-Szarek). There is a computable constant 6 > 0 so that f o r every n >~ 2, there is an n x n matrix An o f norm one which cannot be approximated within 6 by any matrix which decomposes as an orthogonal direct sum o f smaller matrices. The proof is probabilistic. It measures the set of reducible operators, and shows that small balls centered there cannot cover the whole ball of 9Jtn. It is based, in particular, on precise estimates for metric entropy of the unitary groups U (n) and Grassmann manifolds [ 156,157] obtained by the second author while studying problems in geometry of Banach spaces (see [163] for similar more recent results motivated by questions in operator algebras and free probability). Herrero conjectured that the optimal value of 6 is probably 1/2, or at least close to 1/2. However, the 6 found in [95] was quite small, 1.712 x 10 -7. The approach from [24] (cf. Section 2.5) will certainly yield a much better, but still not quite optimal, constant. This negative evidence is not sufficient in itself to answer the question of approximation by m-normals. Nevertheless, the union of the set of m-normals over all finite m is not norm dense in the set of quasidiagonal operators. This was established by two different methods. The second author [159] pushed the probabilistic argument harder and was able to show the existence of block diagonal operators which cannot be uniformly approximated by m-normals.
332
K.R. Davidson and S.J. Szarek
Voiculescu [ 174,175] used more algebraic methods to find an obstruction to approximation. He showed that if T can be approximated by m-normals, then the injection of C* (T) into/3(7-/) is a nuclear map. This is equivalent to saying that C* (T) can be imbedded as a subalgebra of a nuclear C*-algebra. Due to the work of Kirchberg, this class is known to be the set of exact C*-algebras. Voiculescu then shows by non-constructive methods that many block-diagonal operators do not have this property. In the second paper, property T groups are exploited to yield a more constructive approach. These papers describe a path leading to explicit examples by using explicit representations of explicit groups with property T such as S L3 (Z). However, they stop just short of actually producing such an example. See Section 2.5 for further discussion of "constructivity" in this and other contexts. On the other hand, Dadarlat [53] has positive results which show that this exactness condition could be precisely the obstruction. He shows that if T is quasidiagonal and C* (T) contains no non-zero compact operators, then T is the limit of m-normal operators if and only if C*(T) is exact. Voiculescu's Weyl-von Neumann Theorem is an important tool in the approximation theory of Hilbert space operators. So when working with quasidiagonal operators, it is natural to want to use this result and stay within the set of quasidiagonal operators. Generally one wants to approximate T by operators unitarily equivalent to T G A or T G A ~ ) where A = p(rc(T)) and p is a faithful representation of C*(T)/(/C(7~) n C*(T)). We mention in passing that this somewhat resembles the "Petczyfiski decomposition method" used in Banach space theory. So the question arises whether p can be chosen to be quasidiagonal when T is quasidiagonal. Indeed a C*-algebra is called quasidiagonal if it has a faithful quasidiagonal representation. Clearly this property is preserved by subalgebras, but our question asks about specific quotients. It turns out to have a negative answer. Wassermann [ 180,181 ] constructs several counterexamples. He uses the fact that the reduced C*algebra of a non-amenable group cannot be quasidiagonal. His first example uses a connection of the free group to Anderson's example [7] of a C*-algebra 92 for which Ext(92) is not a group. His second paper exploits property T. On the other hand, in [62] it is shown that the quotient of a nuclear quasidiagonal C*-algebra by the compact operators remains quasidiagonal. Theorem 1.13 of Apostol, Foia~ and Voiculescu [10] characterized the closure of the nilpotent operators. An interesting corollary is that the closure of the algebraic operators, those satisfying a polynomial identity, is the set of operators satisfying only the second condition (ii) whenever T - XI is semi-Fredholm, the Fredholm index is 0. Another deep result of Apostol, Foia~ and Voiculescu [9] characterizes this class as the set of biquasitriangular operators. A triangular operator T is one which has an orthonormal basis {en: n ~> 1} in which the basis of T is upper triangular. Say that T is quasitriangular if it is the limit of triangular operators, normally with respect to different bases. An operator is biquasitriangular if both T and T* are quasitriangular. Since every matrix can be put into triangular form, it is easy to see that every block diagonal operator is triangular. Consequently, every quasidiagonal operator if biquasitriangular. The converse is far from correct. Within the class of quasidiagonal operators, one may ask for a description of the closure of all nilpotent or all algebraic quasidiagonal operators. Since quasidiagonal operators trivially satisfy condition (ii) of Theorem 1.13, the natural conjecture might well be that
Local operator theory, random matrices and Banach spaces
333
property (i) should describe the closure of quasidiagonal nilpotents while the closure of quasidiagonal algebraic operators might be all quasidiagonal operators. Unfortunately, the answer to the nilpotent question is no; and the algebraic question remains open. Herrero [94] considered the operator [ ~ / ] where D is diagonal with eigenvalues dense in [0, 1]. This operator is the direct sum of 2 • 2 matrices, and so is 2-normal. The spectrum and essential spectrum is [0, 1] which is connected and contains the origin. So this operator is the norm limit of nilpotent operators. However, there is a trace obstruction to being the limit of nilpotent quasidiagonal operators. A natural finite dimensional question could play a central role in solving this problem. PROBLEM 1.15. If T is an n x n matrix, can the distance from T to the set Nilk of nilpotent matrices of order k be estimated in terms of the quantity IITk II ~/ ~ ? On the other hand, it is easy to exhibit Herrero's operator as the limit of quasidiagonal algebraic operators- just approximate D by diagonal operators with finite spectrum. So he asked whether every quasidiagonal operator is the limit of quasidiagonal algebraic operators. In [62], a special case of this conjecture is verified. THEOREM 1.16 (Davidson-Herrero-Salinas). Suppose that T is quasidiagonal, its essential spectrum does not disconnect the plane, and C*(T)/(IC(~) A C*(T)) has a faithful quasidiagonal representation. Then T is the limit of quasidiagonal algebraic operators. We mention a problem from [62] in the spirit of this survey article. A positive solution would provide a positive answer to the algebraic approximation question. PROBLEM 1.17. Given e > 0, is there a constant C (e) independent of n such that for every matrix T with [[T [[ ~< l, there exists a diagonal operator D and an invertible operator W such that
IT -
WDW -1
and
]]W[[ [[W-' [1 ~ C(e)?
A compactness argument shows that for each fixed dimension n, there is a constant C(e, n) which works.
1.4. Extensions of pure states and matrix paving In this section, we will discuss a well-known problem of Kadison and Singer [99], which asks whether every pure state on the algebra D of diagonal operators on ~2 with respect to the standard orthonormal basis extends uniquely to a (necessarily pure) state on s163 We note that 79 is a generic discrete maximal abelian subalgebra or masa. It is proved in [99] (see also [4]) that the answer to the uniqueness question is negative for non-discrete masas. The interest in the Kadison-Singer problem lies, in particular, in the fact that a positive solution would shed new light on the structure of pure states on s163 In fact it would go a long way towards a simple characterization of such states. Indeed, pure states on 79 ~ s _~
K.R. Davidson and S.J. Szarek
334
C (fiN) are just point evaluations at elements of fiN, the Stone-Cech compactification of N. The canonical extension of a state from 79 to s is obtained by composing the state with the conditional expectation E which takes each T in s to its diagonal part in 79. In other words, if L/is the ultrafilter associated to a pure state q) on 79, then the extension 7t to/~(/~2) is given by ~ ( T ) = ( p ( E ( T ) ) = lim(Ten,en). n~/g
This suggests the open question raised in [99, w whether every pure state on/~(~2) is an extension of a pure state on some discrete masa, or indeed a given masa. However, it has been shown in [6], cf. [4, w that the two questions are in fact equivalent. See also [5] for a related positive result for the Calkin algebra. On the other hand, every pure state on/~(~2) is of the form qg(T) = limncU (Txn, Xn) for some sequence (Xn) of unit vector in ~2 and some ultrafilter in N [ 186]. Moreover, if q9 is not a vector state, the weak limit w-limneU Xn is necessarily 0. Since weak-null sequences in ~2 can be refined to be asymptotically orthogonal, it follows that the difficulty in settling both this and the uniqueness question lies in the difference between limits over ultrafilters and regular limits. Thus, at the first sight, the Kadison-Singer problem seems to be a strictly infinitedimensional question. However, it was shown by Anderson [4] that it is equivalent to a finite dimensional question known as the paving problem. PROBLEM 1.18. Does there exist a positive integer k such that, for any n >~ 1 and any matrix A 6 9Jtn with zero diagonal, one can find diagonal projections P1, P2, 9.., Pk 6 9Jtn such that
(i)
EPi=I
and
PiAPi
(ii)
i=1
1
~ IIAII?
i=1
This is clearly the kind of question that fits into our framework. To clarify the connection we will sketch the argument. Observe first that Problem 1.1 8 is formally equivalent to the following: Given s > O, does there exist a positive integer k such that, for any matrix T in s one can find diagonal projections P1, P2, ..., Pk such that ~ 1 Pi - I and k
k
E( ))Pi
(ii') i=1
E
PiTPi - E(T)
~<slITII?
i=1
Indeed, one gets s in place of 1/2 in the condition (ii) via iteration, with k depending on s. Passing from finite matrices with a uniform estimate on k to infinite matrices considered as operators on ~2, one obtains the partition of N corresponding to the decomposition of the identity on ~2 into a sum of projections from finite partitions via a diagonal argument. Let q9 be an extension to s of a pure state on 79. We claim that qg(D1TD2) = q)(D1)qg(T)q)(D2)
for D1, D2 E 79, T E s163
(l)
Local operator theory, random matrices and Banach spaces
335
We note that a priori ~p is multiplicative only on 79 as a point evaluation on C (fiN) _~ 79. We postpone the proof of (1) and show first how it allows to complete the argument showing that an affirmative answer to Problem 1.1 8 implies that q9(T) -- q9(E (T)) for T E/2 (~ 2 ). Let e > 0. Fix T ~ E(g2) with IIT II <~ 1, and let P1, P2 . . . . . Pk be the projections given by the affirmative answer to infinite variant of Problem 1.1 8. We now identify 79 with ~oo. P1, P2 . . . . . Pk correspond then to a partition of unity in ~ into a sum of k indicator functions of subsets. In this identification, qglz~ is obtained as a limit with respect to certain ultrafilter b/. It follows that among the numbers q9(Pi), i = 1, 2 . . . . . k, exactly one is equal to 1 and the others are all 0, depending on whether the corresponding subset belongs to b/ or not. Accordingly, by (1),
(P
Pi TPi i--1
-- Z
(P(Pi TPi) -- Z
i=1
(p(Pi)(p(T)(p(Pi) = (p(r).
i--1
Therefore
IIk
e >~ ~ _ P i T P i - E ( T )
q)
Pi TPi
- ~o(E(T))
i=1
Since e > 0 was arbitrary, it follows that qg(T) = qg(E (T)), as required. It remains to prove (1). To this end, consider the GNS representation of (E(g2), ~p). There is a Hilbert space H, a norm one vector x 6 H and a ,-representation 7r of s on/2(7-/) such that for all T in s =
An elementary argument shows then that x must be an eigenvector for each 7r(D), D ~ 79 with an eigenvalue qg(D). Hence, for any T in E(g2),
~o(D1TD2) = (Tr(D1TD2)x, x) -- (Tr(D1)rc(T)Tr(D2)x, x) = (Tr(T)Tv(D2)x 7r(D1
x)--(Tr(T)(qo(D2)x) 9o(D1)x)
= qo(D2)qo(D1)(rc(T)x, x ) = qo(D2)qo(D1)qo(T). The converse, uniqueness implies paving [4, (3.6)], can be proved in a similar spirit. One way is to show first that paving is implied by the relative Dixmier property, which says that for T 6 E(s
79(T) := conv{U*TU: U ~ D, U unitary} M 79 ~- 0. This immediately implies that E (T) belongs to 79(T). This is done in very much the same way as the argument presented above. On the other hand, if E ( T ) is not in 79(T), then E(T) can be separated from 79(T) by a functional qg. Heuristically, in view of the balanced
336
K.R. Davidson and S.J. Szarek
nature of 79(T), q9 may be assumed to be positive and with more work, an extension of a pure state on D. By construction, q) is different form the canonical extension obtained by composing with the conditional expectation E. There has been a lot of work on Problem 1.18 in the intervening years, mostly yielding partial and related results. In particular, it was shown by Berman, Halpern, Kaftal and Weiss [30], cf. [43,98], that the answer is positive for matrices with nonnegative entries. A major progress was achieved by Bourgain and Tzafriri. Their work was motivated principally by applications to local structure of L p-spaces, see [98] in this collection for more details. First, in [42], they showed that for Problem 1.18, there exists a diagonal projection P with IIP A P II ~< 89IIA II and rank P ~ 6n, where 6 is a universal positive constant. A closely related result was obtained earlier by Kashin [ 102]. It then clearly follows by iteration that there is a decomposition of identity P1, P2 . . . . . Pk verifying the conditions (i) and (ii)of Problem 1.18 such that k = O(logn). Then in [44], they obtained by far the strongest results to date. Problem 1.18 is solved in the affirmative when the absolute values of entries of the matrix A are relatively small, specifically O(1/(logn) l+~) for some 7/> 0 (Theorem 2.3). Their solution also applies to the cases of Hankel and Laurent matrices with certain regularity properties But the major accomplishment is a saturation result that follows. As earlier, we identify s with the algebra 79. Similarly, the power set P(N) --= {0, 1}r~ is thought of as a subset of s In particular, a C N is associated with the sequence (aj) ~ s the indicator function of a , and with the diagonal projection P,, in 79. THEOREM 1.1 9 (Bourgain-Tzafriri). Given ~ > 0 and T ~ /2(~2), one can find a positive measure v supported on the w*-compact set
K - K(r, of total mass
Ilvll ~
c N: IIP (r - E(r))P
[I <
c
CE -2 f o r which
fK aj dr(or)/> 1 f o r all j >~ 1, where C is a universal numerical constant.
For clarity, we state also the finite dimensional version of Theorem 1.19 from which the theorem easily follows by a diagonal argument. PROPOSITION 1.20. There is a universal constant c > 0 so that given ~ > O, n ~ N and a matrix A E 9J'(n with zero diagonal, there exist a finite sequence of nonnegative weights (ti) with E i ti = 1 and diagonal projections (Pi) such that
IleiAPi II ~ EIIAII f o r all i >~ 1
and
E
ti Pi ~ c~ 2 I. i
Let us point out first that the last condition in this proposition implies by trace evaluation that the rank of at least one of the Pi's is at least ce2n, thus recovering the result from [42]
Local operator theory, random matrices and Banach spaces
337
mentioned above with the best possible dependence on ~. Note also that, except for that optimal dependence, it is enough to prove Proposition 1.20 for some ~ E (0, 1). We wish to emphasize that Theorem 1.19 is very close to implying the affirmative answer to the Kadison-Singer problem. Indeed, if we somewhat carelessly apply a (pure) state q9 to both sides of the assertion of Theorem 1.19, we "obtain" (2)
~O(fKadV ) -- fKcP(a)dv ~ 1.
Hence there is a a E K such that qg(a) = qg(Po) > 0. Thus qg(Po) = 1 because q9 is multiplicative on 79 and Po is idempotent. We now proceed as in the derivation of the original Kadison-Singer problem from the paving problem. As a 6 K, one has IIP~(T - E(T))P~II <~~IITII a n d s o -
E(T))I
:
(T -
E(T )P )I
IITII
for all e > 0. Therefore (p(T) = (p(E(T)), whence qg(T) is determined by its diagonal part. The weak point of this "argument" lies in the fact that the integral fK a d r ( a ) makes only weak* sense, while in equality (2) we implicitly used weak convergence. The argument would work if the measure v was atomic, though. Still, Theorem 1.19 shows that there is an abundance of diagonal projections P~ verifying [[P~ (T - E(T))P~ 11~< ellT[[. That abundance just isn't formally strong enough to guarantee that the collection of such a ' s will intersect every ultrafilter. We show now a simple example to that effect. Let d E 1~ and let Id be the set of all words of length 2d in the alphabet {A, B } consisting of d A's and d B's. So n := #Id = (2d)!/(d!) 2. Next, for s = 1, 2 . . . . . 2d, let as be the set of those words in Id whose sth letter is A. Clearly, Z'd :-- {al, 0"2 . . . . . CrZd } provides a cover of Id. It is not a minimal cover, but d + 1 sets are required. Let r[t be the hereditary subset of the power set 79 (Id) generated by Zd, namely X~ := {a: a C as for some 1 ~< s ~< 2d}. Then every subcover, or partition, of Id consisting of elements of r ~ must have at least d + 1 -- O(logn) elements. On the other hand, it is easily seen that 1 ~-~,2~l Xo, -- 89Thus the set of projections {Po" a E X~} verifies the condition in the assertion of Proposition 1.20, but does not verify the condition of Problem 1.18 with k independent of n. An infinite example verifying the condition of Theorem 1.19, but not that of the infinite variant of Problem 1.18 is routinely obtained by identifying N with ~ d Id and setting Z'--{aCI~:
aAIdEr~foralld~l}.
Of course, this is just a combinatorial contraption. There is no a priori reason why r produced above would correspond to an actual operator T E s163 The methods of [42-44] are quite sophisticated. Without going into details, we mention that the first step in finding large subset a C { 1, 2 . . . . . n} for which IIP~ A P~ I[ is small involves a random procedure. The first approximation is a = {j: ~j = 1 }, where
K.R. Davidson and S.J. Szarek
338
~n is a sequence of independent Bernoulli selectors. These are random variables satisfying P(~j = 1) = 1 - P(~j = 0) - - ~ for j = 1,2 . . . . . n for some properly chosen E (0, 1). The norm of P~ A P~ is then the maximum of a random process, which is analyzed using decoupling inequalities (see, e.g., [106]), majorized by a more manageable maximum of a Gaussian process, and estimated via metric entropy and Dudley's majoration (see [112, (12.2)]). This works if the entries of the matrix A are rather small, such as the O ( 1 / ( l o g n ) l+u) bound mentioned earlier. Such random subsets also yield a partition of {1,2 . . . . . n }. In the general case, one only obtains an estimate on the norm of P~ A P~ as a map from s to ~ . In the context of Proposition 1.20, one has instead a weighted s The final step uses the Little Grothendieck Theorem [ 137]. Alternatively, some of the steps may be done by using the measure concentration phenomena (cf. Section 2.2) already employed in [103] or majorizing measures, cf. [165]. See [98] in this collection for more details on some of the above arguments and related issues. We conclude this section by commenting on the type of examples that need to be analyzed in hope of further progress. To be a potential counterexample, a matrix A -- (aij) ni j__ 1 (meaning a sequence of n • n matrices that together provide a counterexample) must have the following features: (i) ]lAB] must be much smaller than ]](]aij ]) ]], or otherwise we could apply the argument that works for nonnegative entries. (ii) ]aij] do not admit a (uniform) O ( 1 / ( l o g n ) l+u) bound. (iii) On the other hand, the substantial entries of A must be sufficiently abundant, or otherwise one could avoid them by the same combinatorial argument that works for nonnegative entries. (iv) The combinatorial structure of the substantial part of A must be quite rigid to distinguish between the conditions from Proposition 1.20 and Problem 1.18. One structure that comes to mind is related to adjacency matrices of Ramanujan graphs (see, e.g., [116]). Let B = (bij) be such a matrix corresponding to a d-regular graph on n vertices. Then []B ]] = d, and it is achieved on the eigenvector (1, 1 . . . . . 1), while all the remaining eigenvalues are bounded by 2~/d - 1. So the n • n matrix ( 89(bij - k ~ n ) / ~ / d - 1) is of norm at most 1, and appears to enjoy the features (i)-(iv), some of which are admittedly vague. The question would then be to determine whether matrices obtained this way from various constructions of Ramanujan graphs can be paved. It seems at the first sight that new techniques are required for any kind of answer. ~1, ~2 . . . . .
1.5. Hyper-reflexivity If r is an operator algebra contained in s then L a t A denotes the lattice of all of its invariant subspaces. Dually, given a collection 12 of subspaces, Alg/2 denotes the algebra of all operators leaving each element of 12 invariant. The algebra ,A is reflexive if Alg Lat,4 = A. There is a quantitative version of reflexivity which has proven to be a powerful tool when it is available. Notice that if PL is the projection onto an invariant subspace L of .,4, then for any operators T E s and A 6 .A,
II
II- II
A)P li
-
All.
Local operator theory, random matrices and Banach spaces
339
Hence the inequality
fls
"-- sup
IIPL TPL LI
dist(T, A).
LE/Z
The algebra A is hyper-reflexive if there is a constant C such that dist(T, A) <~ Cfls
for all T E/3(7-/).
We also say that an operator A is reflexive or hyper-reflexive if the unital WOT-closed algebra W ( A ) generated by A is reflexive or hyper-reflexive respectively. In finite dimensions, the question of which operators are reflexive was solved by Deddens and Fillmore [69] in terms of the Jordan form. Evidently, every reflexive subalgebra of 9Ytn is also hyper-reflexive; but there is no a priori estimate of the constant even in two dimensions. For example, it is very easy to show that the 2 • 2 diagonal matrices 792 are hyper-reflexive with constant 1. Similarity by an operator S will preserve hyper-reflexivity, but can change the constant by as much as the condition number [IS I] IIS-1ll 9Thus it may not be too surprising that the hyper-reflexivity constant of St 792S t | increases to +e~ with t it
when St - [0 1 ]. There are two classical situations in which the existence of hyper-reflexivity has played a key role. The first is nest algebras. A nest A/" is a chain of subspaces containing {0} and 7-/ which is closed under intersections and closed spans. The nest algebra is T(A/') = AlgA/" is the algebra of all operators with an upper-triangular form with respect to this chain. In particular, the algebra T~ of n x n upper triangular matrices with respect so some basis is the prototypical finite dimensional example. An early result of Ringrose [ 145] shows that T(A/') is reflexive, as well as the fact that LatT(A/') --A/'. Arveson [13] showed that nest algebras are hyper-reflexive with constant 1, so that one obtains the distance formula: THEOREM 1.21 (Arveson). Let jV" be a nest, and let A be an arbitrary operator in 13(7-[). Then dist(A, T ( H ) ) -
sup
[IPAPNII.
NcA/"
An easy proof [144], cf. [58], can be based on the well-known matrix filling lemma of Parrott [ 130] and Davis-Kahane-Weinberger [68] valid for matrix or operator entries:
inf xcs
A
B
-max{ll[A
n]ll,
o
In particular, Lance [108] used this distance formula to establish that close nests are similar. Likewise, this fact was a key ingredient in the first author's Similarity Theorem [54] stating that two nest algebras are similar if and only if there is a dimension preserving isomorphism between the underlying nests. The search for generalizations of this will lead to an open question below.
340
K.R. Davidson and S.J. Szarek
The second context in which this concept occurs is the von Neumann algebras. Christensen [52] showed that many von Neumann algebras (specifically all whose commutant do not have certain type H1 summands) are hyper-reflexive. This is closely related to derivations. Recall that a derivation is a map 8 of an algebra .A into an .A-bimodule A4 satisfying the product rule 8(ab) -- aS(b) 4- 6(a)b. The derivation is inner if there is an element x e .A4 such that 8(a) = ax - xa. Christensen showed: THEOREM 1.22. For a v o n Neumann algebra 92 contained in s the following are equivalent: (i) Every derivation 8 o f 92~ into s is inner. (ii) There is a constant C such that dist(X, 92) ~< C l l S x l ~ , l l f o r all X ~ s The second condition actually asserts that 92 is hyper-reflexive. The invariant subspaces of any C*-algebra correspond to projections in the commutant. Every von Neumann algebra is spanned by its projections. Thus von Neumann's famous double commutant theorem, which asserts that (92~)~ = 92, shows that every von Neumann algebra is reflexive. Moreover, the convex hull of the symmetries 2P - I, for P projections in 7)(92), is the whole unit ball of the self-adjoint part of 92. Now a simple calculation shows that
l[(2P -
I)X-
X ( 2 P - l)ll -
2llpxP"
- P'xP]l
2max{llexp•
Ile xell}.
Thus since every operator is the sum of its real and imaginary parts, 118xl~'ll % 4
sup IIP'xPII"
P e7)(9~)
So (ii) is equivalent to hyper-reflexivity. For all injective von Neumann algebras, the constant is at most 4 [52]. And for abelian von Neumann algebras, the constant is no more than 2 [ 146]. Even for the 3 x 3 diagonal matrices, the constant is greater than 1. Indeed, it is exactly ~/-3/2 [63]. So unlike the nest case, most examples involve non-trivial constants, not exact formulae. Now we turn our consideration to the class of woy-closed reflexive algebras which contain a masa, known as CSL algebras. The terminology is short for commutative subspace lattice algebra because any invariant subspace is, in particular, invariant for the masa, and therefore corresponds to a projection in the masa. So the orthogonal projections onto all of these invariant subspaces commute with each other. These algebras were introduced in a seminal paper by Arveson [14]. The finite dimensional versions occur in many contexts, and are also called incidence algebras or digraph algebras in other parts of the literature. A masa in ~)T~n is unitarily equivalent to the diagonal algebra 7)n. Any algebra containing Dn is determined by the standard matrix units Eij which belong to the algebra. Moreover, since this is an algebra, the set of matrix units is transitive in the sense that if Eij and Ejk belong, then so does Eik. Thus one may associate a directed graph to the algebra with n vertices representing the
Local operator theory, random matrices and Banach spaces
341
standard basis, and including a directed edge from node j to node i if Eij lies in the algebra. In this way, we obtain the graph of a transitive relation. Not all CSL algebras are hyper-reflexive. Davidson and Power [67] constructed finite dimensional examples with arbitrarily large distance constants. This is more subtle than the easy example mentioned earlier. However, Larson [109] showed that this construction was part of a general mechanism for producing examples of this type. On a positive note, Davidson and Pitts [65] showed that if there is a dimension preserving lattice isomorphism between two CSL algebras, then the two lattices are approximately unitarily equivalent. This was established for nests by Andersen [3], and was a key step in obtaining similarity invariants. Pitts [ 142] was able to obtain good perturbations results for the algebras without using hyper-reflexivity in spite of the fact that hyper-reflexivity was used in an important way in the nest case. But other possible extensions of results for nests to this more general context have been hampered by the lack of hyper-reflexivity. For this reason, attention has been focussed on an important finite dimensional case (see [58,59]). The algebra ,An = T~ | Tn represents the subalgebra of ~lY~n2 consisting of n x n upper-triangular matrices with n x n upper-triangular matrix entries. The problem becomes: PROBLEM 1.23. Is there is a finite upper bound to the hyper-reflexivity constants of all of the algebras ,An ? This is a typical situation where one understands well the one dimensional case, but not the higher dimensional or multivariable case. Similar difficulty appears when studying, e.g., multi-indexed orthogonal expansions. A few other examples of hyper-reflexivity are known and are worth mentioning. The unilateral shift generates the analytic Toeplitz algebra, and the first author [57] showed that it is hyper-reflexive. Then with Pitts [66], he showed that the algebra generated by the left regular representation of the free semigroup on n letters is also hyper-reflexive. Popescu [143] generalized this to a wider family of semigroups. Recently, Bercovici [26] used predual techniques to establish hyper-reflexivity constants for the large class of algebras with property Xo,• This property, which we do not define precisely, allows approximation of any weak-, continuous functional on the algebra by a sequence of rank one functionals tending to infinity in a very strong sense. These notions arose in using predual techniques to establish reflexivity for single operators beginning with the celebrated theorem of Scott Brown [49] that every subnormal operator has invariant subspaces. Olin and Thomson [128] showed that subnormal operators are reflexive. And we mentioned above that the prototypical subnormal operator is hyper-reflexive. Many subnormal operators fit into Bercovici's criterion, but the general question of hyper-reflexivity for subnormal operators remains open.
2. Random matrices In this chapter, we shall present a selection of classical and not-so-classical results on asymptotic spectral properties of random matrices that are related one way or the other to
K.R. Davidson and S.J. Szarek
342
the geometry of Banach spaces. As was the case with other topics, these connections come in several flavors. First, facts about random matrices are being applied to the geometry of Banach spaces. Second, the methods of the Banach spaces yield new results or offer an alternative perspective on random matrices. Finally, there are connections via vague analogies between the fields.
2.1. The overview Random matrices appeared in Banach space theory in an explicit way some time in the 70s, for example in [23,25] and [101], even though one can claim that their spirit was already present, e.g., in probabilistic proofs of Dvoretzky theorem, cf. [81] in this collection. Let us quote here a result from [23]: THEOREM 2.1. Let q E [2, oc) and f o r some m, n E N, let A = ( a i j ) be an m x n random matrix whose entries aij -- aij (co) are independent, mean-zero real random variables with [aij [ <<.1 f o r all i, j. Then
Ell A
~ ~qm
II
~
K m a x i m 1/q n 1/21
where E denotes the expectation and K -= K (q) is a numerical constant depending only onq.
This is a very typical statement as far as the asymptotic theory of finite dimensional Banach spaces is concerned. We have an estimate giving the correct rough asymptotic order (see the introduction for terminology) and involving a universal numerical constant. Usually we do not know, and often do not really care about, the optimal value of that constant, i.e., the precise asymptotic order. This could be viewed as an unsatisfactory situation from the point of view of other related fields. We shall return to this issue later on. On the other hand, in spite of the "asymptotic" qualification, we do have above an inequality valid for any m, n, a crucial detail for applicability to fields like geometry of Banach spaces, computational complexity and approximation theory. On the other hand, random matrices were of interest to statisticians at least since the 20s, and to theoretical physicists at least since the 50s, see [121]. Perhaps the most celebrated result coming from the latter area is the Wigner's Semicircle Law [184,185] which says that, under some weak regularity assumptions, the spectra of large symmetric random matrices are, in a sense, virtually deterministic and their spectral densities are, after proper normalization, asymptotically semicircular. More precisely, Wigner proved a somewhat weaker statement: THEOREM 2.2. For each n E N, let A -- A (n) (co) be an n x n random matrix whose entries a~; ) (co)" I2 --+ IR are symmetric real random variables satisfying f o r each n ~ 1 and l<.i,j~n, (i) Elaij 12 --- 1/n.
aij =
(ii) a~; ) are independent f o r 1 <. i <. j <. n. (n) (iii) a{~.) - - a j i .
Local operator theory, random matrices and Banach spaces
343
(iv) For each m ~ 1, there is a tim < CXDindependent of n so that E[aij Im ~ tim. Then, f o r any v E N, lim ETr(A v) -- fR xv d#(x),
(3)
n---->oo
where Tr is the normalized trace in the respective dimension and # is the measure on I[{ whose density w is given by
w(x) --
1 x/4_x 2 TY 0
i f x E [ - 2 , 2], if x ~ [ - 2 , 2].
The measure # above is often referred to as the standard semicircular distribution. Theorem 2.2 says in effect that, for large n, the random measure l // # n - - # n (CO) "-- -- Z (~)~j(A), n j=l
where 6x is the Dirac measure at x and ~.l(B) ~> ..- ~> ~.n (B) are eigenvalues of B e 9J~)]a, is approximately standard semicircular when considered as a measure on ~2 x • (which is the same as convergence in distribution in the free probability sense, as explained in Section 2.4). In fact, as was implicitly assumed by physicists and proved later (cf. [12]), a much stronger statement holds: for large n the random measures # , (co) are "close" to # with probability close to 1 (convergence in probability). In particular, #n (co), the empirical measure associated to the spectrum of A (n), is nearly deterministic. We emphasize that this is far from being a formal consequence of Theorem 2.2. It is possible in principle that the assertion of the theorem holds while, for any co E ~2, A (co) is a multiple of identity: just consider a scalar random variable ~ whose law is standard semicircular and a random matrix A :-- ~ I. However, we do have THEOREM 2.3. In the notation and under the hypotheses of Theorem 2.2 one has, f o r any interval I C ~, lim n--->oo
#{eigenvalues of A (n) contained in I }
= #(Z)
almost surely.
n
There are similar results for the asymptotic distribution of singular values of random non-selfadjoint matrices; e.g., when the hypothesis (ii) in Theorem 2.2 is replaced by (ii) a~; ) are independent f o r 1 <, i, j <, n and the hypothesis (iii) is dropped. The limiting distribution is then a quartercircle law (singular values being necessarily nonnegative), supported in the interval [0, 2] and given by the density ?1 ~ / 4 - x 2 Similarly, one may consider (see [118,179]) large rectangular matrices such that the ratio of the sides is roughly fixed, cf. Theorem 2.13, and many other "ensembles" (cf. the remarks preceding Theorem 2.7).
344
K.R. Davidson and S.J. Szarek
The example preceding Theorem 2.3 notwithstanding, the implication Theorem 2.2 =~ Theorem 2.3 is in fact a formal consequence of rather standard local theory techniques, even though this was not the way how it was historically shown. We shall sketch the argument later in this chapter, after Theorem 2.7; see also Corollary 2.12 and Theorem 2.17. As a bonus, one obtains fairly strong estimates on the probabilities implicit in the almost sure part of Theorem 2.3. That proof, and many other arguments presented here, while being a folklore among Banach space theorists, are not exactly a common knowledge in the random matrix circles. As opposed to the techniques emphasized here, standard tools used in that area involved the moment method (roughly, working directly - via a heavy duty combinatorics- with the moments involved in the limit in (3), cf. the remarks following Problem 2.18), the more precise Stieltjes transform method introduced in [118] (cf. [132]) and later adopted in and developed by the free probability approach, or, in the case of classical random matrices, analyzing the explicit formulae for joint densities of eigenvalues or singular values, see (8), which leads to the so-called orthogonal polynomials method, see, e.g., [121]. In other directions, it has been determined that much weaker assumptions on regularity of the entries suffice: one needs just a little bit more than the existence of second moments, with uniform estimates (more precisely, a Lindeberg type condition), and the symmetry hypothesis may be replaced by Eaij = 0 (see, e.g., [ 131,82]). We point out that assertions of Theorems 2.1 and 2.3 are not comparable. On one hand, Theorem 2.3 does not say anything about the norms of matrices. It is consistent with its assertion that o(n) largest (in absolute value) eigenvalues are far outside the interval [ - 2 , 2], the support of w. On the other hand, as indicated earlier, Theorem 2.1, being "isomorphic" in nature, does not give the precise order of the norm (even when q -- 2 and m = n), and it does not address the question of the asymptotic distribution of the eigenvalues (or singular values in the non-selfadjoint case). Again, common strengthenings of results of the two kinds have been obtained (see, e.g., [46,80,18]), but we do not know of any really satisfactory argument that encompasses simultaneously the two aspects of the picture. In the opposite direction, precise estimates were obtained on the (very small) probability that a specific eigenvalue (or singular value) of a random Gaussian matrix is far away from its theoretical value predicted by the corresponding Semicircle Law result. A sample such large deviation result, motivated by questions in geometry of Banach spaces and numerical analysis, is Theorem 2.4 below. Before stating the theorem, a few words about the level of generality of our discussion. More often than not, we shall concentrate on the central Gaussian case, and most of the arguments will be specific to that case. In particular, throughout this chapter G = G (n) will stand for an n x n random matrix whose entries gij, 1 <~ i, j <<,n, are independent identically distributed Gaussian random variables following the N(0, 1/n) law. Similarly, A = A (n) will usually stand for a Gaussian selfadjoint matrix verifying the hypotheses of Theorem 2.2. However, both the results and the methods employed are representative of much more general setting. It would be a useful project to work out in detail the consequences of general concentration results for product measures (see [166,167] and their references) and related "probability in Banach spaces" tools to the setting of random matrices. As indicated earlier, very few papers produced in the random matrix circles employ those methods and, to our knowledge, no vigorous research in this direction was attempted. It is
Local operator theory, random matrices and Banach spaces
345
clear that some consequences would be just a formal repetition of the arguments sketched in this section. Chances are that, to get best results, one may need to apply those tools rather creatively. The payoff can be substantial since important part of research in random matrices turns around "universality" of the limit laws, i.e., their independence of the particular probabilistic model involved, the "symmetries" of that model being the only meaningful parameter (the uniw;rsality conjecture). THEOREM 2.4 ([161,162]). Let G = G (n) be as above and S1 ~ $2 ~ "'" ~ Sn be its singular values. Then
(
dl
"7-0 Sn-d+l > ~-n
~
exp(-cfi2d2),
(
d)
s~-d+l <~~-n
<~ (C~) a2
f o r 1 <~ d <~ n, fl >~ [3o and ot ~ O. Above, c, C and flo are universal positive constants. Apart from the precise values of the constants, the estimates are optimal except possibly f o r the f o r the first one when n - d -- o(n). Analogous results can be obtained in the complex case and, for eigenvalues, in the selfadjoint case. The optimality of the theorem means here that there are similar lower estimates with c, C replaced by other positive universal constants. However, being an isomorphic result, the Theorem doesn't detect smaller deviations from the values predicted by the Quartercircle or Semicircle Law. Much more precise results of the same nature were obtained very recently in the case when d is asymptotically a fixed proportion of n [22]. One of the concepts employed there is the free (noncommutative) entropy [ 178], arrived at by following on the ideas sketched in Section 2.4 below. It is quite likely that the methods generalize to the case d = o(n), but probably not to the edge of the of the spectrum, i.e., the case when d is nearly equal to n; cf. Theorem 2.8 below and the comments following it. It would be potentially useful to clarify the relationship between Theorem 2.4 and these results. To complete the overview, we shall mention that, besides the large deviation results mentioned above and the results in the spirit of Wigner's Semicircle Law (the so-called global regime), there is a large body of research dealing with micro-local analysis of the spectrum {)Vl(A(co)) . . . . . )vn (A(co))}, and particularly the gaps )vj -- )v j + 1 in that spectrum (the local regime). These are natural features to consider in view of the original physical motivation, the eigenvalues being interpreted there as energy levels. Further attention to this direction was brought by the largely experimental results [ 127,77] relating the properly defined distribution of such gaps and the corresponding gaps between zeros of the Riemann ~" function; cf. [126,148]. In what follows, we shall state a result in the spirit of [122], the fundamental work concerning the local regime. It deals with a complex analogue of A (n) called the Gaussian unitary ensemble or GUE (defined more precisely in item (iii) the next section), for which Theorems 2.2 and 2.3 hold with the same limiting semicircle distribution. We need to introduce first some notation: given B ~ 93tsa with eigenvalues )vl (B) ~> 9.. ~> )vn (B) and x E R, we define 6B(x) to be the length of the interval ()vj-1 (B),~,j(B)] containing x, with the convention that 6 ~ ( x ) = oo if x > ~1 (B) or X ~ )vn(B ) and 68(x) -- 0 if x is a multiple eigenvalue.
K.R. Davidson and S.J. Szarek
346
THEOREM 2.5. There exists a probability density cr supported on [0, cx~) such that when1 ~ / 4 - x 2 the density of the standard semicircle distribuever x E ( - 2 , 2) and w(x) - ~-y tion, then, f o r any s ~ O,
(
s)/0
lim 79 6GUE(X) <<, n~ nw(x)
--
a(u) du.
We point out that the scaling of 6GVE(X) by the quantity n w ( x ) in Theorem 2.5 was to be expected: by Theorem 2.3, for large n and a short interval 2- containing x, the number of eigenvalues of the random matrix that belong to 2- will be, with probability close to 1, approximately n f l w ( u ) d u ~ nw(x)12-1, and so the gaps should average about 1 / ( n w ( x ) ) . Added in proof There was some very recent progress in the direction of the universality conjecture, i.e., generalizing results from classical ensembles of random matrices to those whose distributions of the entries are more generic. For example, large deviation results related to Theorem 2.4 and [22] were obtained in [87]. In the language of Theorem 2.7 below, a version of the inequality (5) for the extreme eigenvalues )~l, )~n was shown in [ 105]. Both of these papers use the measure concentration phenomenon and specifically techniques of Talagrand [ 166,167], thus addressing some elements of the program sketched in this survey. Particularly [87] is a tour de force. Some reasonably general results related to Theorem 2.8 below were obtained in [154]. Finally, some initial progress on the topics relevant to Theorem 2.5 was achieved in [97].
2.2. Concentration of measure and its consequences In this section we shall sketch some applications of the measure concentration phenomenon, well known and widely applied in local theory of Banach spaces (see the article [ 150] in this collection), to the subject of random matrices. Most of these applications have been a folklore among some of the experts in the former field, but to the best of our knowledge, they haven't been presented anywhere in a systematic fashion. One of possible starting points is the Gaussian isoperimetric inequality [39], which we shall present here in the functional form (see [ 136] for the last assertion). Recall that a function F defined on a metric space (X, d) is called Lipschitz with constant L if IF(x) - F(x')l <~ L d ( x , x ~) for all x, x ~ E X. Let y -- Vn be the standard Gaussian measure on ~ n with density (27r) -n/2 e -Ixl2/2, where I" I is the usual Euclidean norm. As usual, ~ ( t ) "-- ~'1 ((-cx~, t]) is the cumulative distribution function of the N (0, 1) Gaussian random variable. THEOREM 2.6. Let F be a function on ~n which is Lipschitz with constant L with respect to the Euclidean metric and let M = MF be the median value of F with respect to Vn. Then, f o r any t > O, 79 (F ~ M + t) ~ 1 - ~ ( t ) < e x p ( - t 2 / 2 L 2 ) .
(4)
One has the same upper estimate e x p ( - t 2 / 2 L 2) if the median M is replaced by the expected value fR, F dyn.
Local operator theory, random matrices and Banach spaces
347
For future reference we point out here that, for a convex function, its median with respect to a Gaussian measure does not exceed the expected value; see [71,107] or Corollary 1.7.3 in [76]. The relevance of Theorem 2.6 to random (Gaussian) matrices is based on the elementary and well-known fact, described in Section 1.2 of this article, that spectral parameters like singular values (respectively eigenvalues in the selfadjoint or unitary case) are Lipschitz functions with respect to the matrix elements; see [35,33] for related more general results. In particular, for each k c {1. . . . . n}, the k-th largest singular value sk(X) (respectively 17 the eigenvalue ,kk(X)) is Lipschitz with constant 1 if X (Xjk)j,k=| is considered as an - -
element of the Euclidean space R n2 (respectively the submanifold of R n2 corresponding to the selfadjoint matrices). If one insists on thinking of X as a matrix, this corresponds to considering the underlying Hilbert-Schmidt metric. Accordingly, in the context of applying Theorem 2.6, the Lipschitz constant of sk (G (n)) is 1/~/-n (because of the variances of the entries being l / n , which corresponds to the identification G = G (n) = 1/x/~ X). Respectively, for a Gaussian selfadjoint matrix A = A (n) verifying the hypotheses of Theorem 2.2, the Lipschitz constant of )~k(A(~)) is v/2/n. The additional 2 is a consequence of the same variable appearing twice, in the jkth and kj th position. The above comments, and hence the results below, carry essentially word for word to the following often considered variants, all Gaussian unless explicitly stated otherwise. (i) A variant of A = A (") in which variances of the diagonal entries are assumed to be 2In rather than l / n , called often the Gaussian orthogonal ensemble or GOE. This is in fact the same as v/2 times the real part of G (n) . Note for future reference that GOE can be represented as Y + A, where Y is a (diagonal) Gaussian random matrix independent of A, and so many results for GOE transfer formally to A. (ii) The complex non-selfadjoint case, all the entries being independent and of the form g + ig', where g, g' are independent real N(O, 1/2n) Gaussian random variables. (iii) The complex selfadjoint case: formally the same conditions as in Theorem 2.2 (except for the obvious modification in the symmetry condition), but the abovediagonal entries are as in (ii) while the diagonal entries are real N(0, 1/n)'s. Again, this is the real part of the matrix in (ii) times v/2, and is frequently referred to as the Gaussian unitary ensemble or GUE. (iv) Rectangular, real anti-symmetric or complex anti-selfadjoint matrices. (v) Orthogonal or unitary matrices distributed uniformly on SO(n), respectively U (n). It is also easily seen that in the first three cases above the Lipschitz constants are respectively 2 x / ~ , 1 / v / ~ and 1/v/-n. The anti-symmetric/anti-selfadjoint/rectangular cases are treated the same way, and there are equally useful results for orthogonal/unitary matrices cf. [124,86]. There appears to be no easily available exposition of the unitary case, even though all ingredients are available, cf. [176]. Still, a word of caution is needed. As noted in Section 1.1, eigenvalues are not very regular functions of general (non-normal) matrices. Combining the above remarks and Theorem 2.6 we get THEOREM 2.7. Given n E N, there exist positive scalars Sl, $2 Sn such that the singular values of the n x n real Gaussian random matrix G = G (n) satisfy .
79(Isk(G)- ski ~> t) < 2 e x p ( - n t 2 / 2 )
.
.
.
,
(5)
348
K.R. Davidson and S.J. Szarek
f o r all t ~ 0 and k -- 1, 2 . . . . . n. This holds, in particular, if sk's are the medians or the expected values of sk(G). Similar results hold f o r the eigenvalues of the real symmetric Gaussian random matrix A -- A (n) with - n t 2/4 in the exponent on the right side of (5). Moreover, the corresponding deterministic sequence ~.l, )~2 . . . . . )~n can be assumed to be symmetric, i.e., )~k = --)~n-k+l f o r k = 1, 2 . . . . , n. Likewise, related results hold f o r other Gaussian random matrices, in particular those described in (i)-(iii) above.
A shortcoming of the above result is that it doesn't see the possible relationships between the deterministic sequences sl, s2 . . . . . Sn (or)~1, ~.2 . . . . . )~n) for different n's. Still, it allows to formally deduce statements in the spirit of Theorem 2.3 from the corresponding Theorem 2.2 like results. Consider, as an illustration, the ensemble A = A(n). Once we know that there is a rough estimate on, say, EIIAII (e.g., of the type of Theorem 2.1), the analogue of (5) for A implies via an elementary calculation ETr(A v) - ~ x v d#(n)(x) <~
Ci)
(6)
for v 6 N, where #(n). ~1 Z j n= I 3)~j is the deterministic measure involving the expected values or medians of the eigenvalues of A and Cv is a constant depending only on v. Combining this inequality with Theorem 2.2 we deduce that #(n) __+ #, the semicircular distribution from Theorem 2.2, weakly as n --+ ec. The estimate on the right-hand side of (5) and the Borel-Cantelli lemma imply then the assertion of Theorem 2.3 for our ensemble A (n) or for any ensemble for which a result of the Theorem 2.7 variety holds. We emphasize that this is independent from how the matrices A(n) are stochastically related for different n's. We shall present even stronger results (Corollary 2.12, Theorem 2.17) in the next two sections. Another consequence of Theorem 2.7 is that, in the normalization we use, fluctuations of singular values or eigenvalues of n x n (Gaussian) random matrices are at most O(n-1/2): if t >> n-l/2, the exponent in (5) becomes "large negative". We need to point out that, in all likelihood, this is not an optimal result, and certainly not for the full scale of parameters. For example, it is conjectured that fluctuations of eigenvalues in the bulk of the spectrum (i.e., neither k = o(n) nor n - k = o(n)) are of the order O(n-1), and the conjecture is supported by the large deviation results quoted in the preceding section (see Theorem 2.4 and the paragraph following it). There are several partial results in that direction of varying degrees of generality and strength (see, e.g., [16,17]). In the Gaussian case, an improvement to the O(n -1/2) result "nearly" follows from the approach presented and the results stated in this article. For example, if we knew that the differences between the consecutive deterministic )~j's from Theorem 2.7 were of order 1/n (which we "almost" do, cf. Theorem 2.5), we could argue that, for a )~k(A) to be 0 or more away from its central value )~k, approximately the same would have to be true for con neighboring eigenvalues and so, by Theorem 1.8, the square of the Hilbert-Schmidt deviation from the most likely spectral picture would be of order t 2 = 03n; substituting this value, and L 2 = 2In, into an estimate of type (4) we would obtain a meaningful estimate whenever O/n -2/3 was large. Another promising approach would be in exploiting the representation of G (n) (which works also _
Local operator theory, random matrices and Banach spaces
349
for A (')) found in [152]. A related fact which is worth pointing out is that there are two sources contributing to the quantity in (6): the deviation of the Aj(A)'s from their central values k j ' s , and the difference between the deterministic measures #(n) and the limit semicircular distribution #, and both of them are of interest. For the very edge of the spectrum, i.e., the largest or smallest eigenvalues and perhaps a few adjacent ones and in particular for the norm of the matrix, the fluctuations are, in some cases, known to be of order n -2/3. A sample result (see [170]) is THEOREM 2.8. There exists an increasing function (p on IR such that the largest eigenvalue A I(GOE) o f the n • n Gaussian orthogonal ensemble satisfies lim 79(A1(GOE) <~ 2 + rn -2/3) --99(r)
(7)
t/----> o o
f o r r ~ JR, the convergence being uniform on compact subsets o f ]t{. Similar result (with a different 99) holds f o r the Gaussian unitary ensemble.
The reader will notice that exactly this size of fluctuations is predicted by the semicircle law itself: if we choose s > 0 so that 1/27r f22_,, x/4 - x 2 dx -- 1/n, then s ~ (3rc/(Zn)) 2/3. On the other hand, it is easy to see that (7) does not accurately predict the order of the probabilities without passing to the limit. For example, if r ~> n 1/6, which corresponds to t ~> 1 in Theorem 2.7, (5) yields the correct asymptotic order of l o g 7 9 ( k l ( G O E ) <, 2 + t), which, because of a difference in scaling, is inconsistent with the behavior that would have been suggested by (7). In view of possible applications in local theory, it would be potentially useful to recover the asymptotics on the probabilities involved, given by the arguments of [170], for the full range of the parameters, a similar project to the one suggested in the paragraph following Theorem 2.4 in the context of large deviation results. The proof of Theorem 2.8, and similarly, the proofs of Theorems 2.5, 2.4 and the other large deviation results, uses the explicit formulae for joint densities of eigenvalues (respectively singular values), which are of the form
l~j
l<~k<~n
where /3 = 1 or 2 depending on whether the context is real or complex and c(fl, n) is the normalizing numerical coefficient. Accordingly, one can not expect that it generalizes much beyond the Gaussian and some other classical random matrices. However, there is a strong circumstantial evidence that at least Theorem 2.5 (see [133,70]) and Theorem 2.8 are much more universal. Concerning the latter, it has been shown in [153] that the moments E Tr(A v), where A = A (') are real symmetric matrices satisfying any of our variance and independence assumptions and such that the laws of the appropriately normalized matrix elements are uniformly sub-Gaussian, exhibit, as n, v --+ oo, an asymptotic behavior which is consistent with the assertion of the theorem. A concentration result for the norm in similar degree of generality but going in a somewhat different direction can be found in [45]. As suggested earlier, it would be an interesting project to figure out relevant consequences of measure concentration phenomenon so successfully exploited in local theory.
350
K.R. Davidson and S.J. Szarek
The phenomenon signalized in Theorem 2.8 has not been exploited in local theory of Banach spaces, and so it is not known whether, conversely, any of the methods of that area are relevant here. It would be extremely interesting to find a general concentration principle which would imply the small fluctuations of the norm and/or eigenvalues in the bulk of the spectrum. Techniques and ideas that may be relevant here (and, more generally, when investigating the "local regime") include hypercontractivity, logarithmic Sobolev or Poincar6 type inequalities, transport of measures and various concepts of information theory, see [21,38,110,111,129,168].
2.3. Norm of a random matrix and the Slepian-Gordon lemma In this section we shall present a simple argument giving exact asymptotics for extreme singular values (basically norms) and eigenvalues of some symmetric Gaussian random matrices (in particular all real Gaussian matrices considered here). The argument is based on the well-known Slepian's lemma from probability and its generalization due to Gordon. For greater transparence we shall state the relevant special cases of both variants. LEMMA 2.9. Let ( X t ) t ~ T and (Yt)tcT be two finite families of jointly Gaussian mean zero random variables such that
(a)
IIX, - Xt, II2 ~ IIYt - Yt, l[2 f o r t , t' ~ T.
Then E max Xt ~< E max Yr. toT
(9)
t~T
Similarly, if T = U s e s Ts and
(b)
IIXt
(c)
IIXt - Xt, ll2 ~ IIYt - Yt' 112
-
Xt, ll2 ~ IIYt - Yt'll2
if t ~ Ts, t' ~ Ts, with s =/: s', if t, t' ~ Ts for some s,
then one has E max min Xs,t ~ E max min Ys,t. soS t~Ts
soS tCTs
REMARK 2.10. Lemma 2.9 is usually stated with an additional hypothesis IlXtl[2 = IIYtll2 for t 6 T, and yields the stronger assertion for any k E R, 79(maxt~T Xt > )~) <~ 79(maxt~T Yt > k); (9) is then an immediate consequence. Analogous comment applies to the second part of the lemma. Let us also point out that when all Ts are singletons, the second part of the lemma (the Gordon version) reduces to the first (the Slepian version), which in our formulation coincides with a result of Fernique [75].
Local operator theory, random matrices and Banach spaces
351
A typical context in which L e m m a 2.9 shall be applied is the probability space (~N, VN) with linear functionals as random variables. For u E R N, set Zu := {., u}; for future reference we point out that the map ~N ~ U ~ Zu E L2(]t~N, }IN) is an isometry. If K C R N, one has the gauge, or the Minkowski functional of the polar of K given by maxuEK Zu = I[" 11/~~ The quantity E maxu E/~ Zu may also be expressed in terms of the mean width of K. These relationships provide a link between Gaussian processes and convexity or geometry of Banach spaces. It is also clear that in this context T could b e any bounded set, as long as we replace max and min by sup and inf. In particular, if qJ" T --+ T is a surjective contraction (possibly nonlinear) between subsets of two Euclidean spaces, then Eli. liT~ ~ Eli 9I1~o. In our setting, the norm [1. lifo is the operator norm on n x n real matrices identified with N n2 For such a matrix X and u v E R n we have o
<Xu, v} - tr(X(v | u)) - <X, u | V}tr = Z , | where, as usual, v | u stands for the rank one
matrix (Ujl)k)j,k=l, that is, the matrix of
the map x --+ (x, v)u and {X, Y)tr " - tr(XY T) is the trace duality, (sometimes referred to as the Hilbert-Schmidt scalar product), which can also be thought of as the usual scalar product on R n2" note t h a t - as opposed to the remainder of the p a p e r - we use here the standard, i.e., not normalized trace. Accordingly IlXll-
max
{Xu, v } =
U, u E S n - I
max
z.|
lg, u E S n-1
The Gaussian process Xu,v " - Zu| u, v E S n-1 is now going to be compared with Yu,v := Z(u,v), where (u, v) is thought of as an element of R n x R n - R en. In view of prior remarks, to show that the Slepian's lemma applies, one only needs to verify that, for U , 1), bl t
V I E S n-1
(~o) an elementary exercise. On the other hand, for (x, y) E ]Rn x R n , Z(u,v)(x, y) = {x, u) + (y, v) whence max
u,I)ES n-I
Z(u,v)(x, y ) = Ixl +
lYl
(this is just saying that II-II(uxv)o - I1" IIuo + II" IIvo) and so the assertion of L e m m a 2.9 translates to
n'/2Ell
II 2
txl dyn<X
352
K.R. Davidson and S.J. Szarek
By comparing with the second moment of l. I, the last integral is easily seen to be < n 1/2 9 In fact, it equals v ~ F (--U-) n + l / F (~). n Note also that I I2 is distributed according to the familiar x2(n) law. The same argument, applied just to symmetric tensors u | u, allows us to analyze X I(GOE), the largest eigenvalue of the Gaussian orthogonal ensemble. The case of )~1 (A(n)) can then be deduced formally (see the comment after the definition of GOE). In combination with Theorem 2.6 and the remark following it, the above shows: 9
THEOREM 2.1 1. Given n ~ I~I consider the ensembles of n x n matrices G, A and GOE. If the random variable F equals either lIGI], X1 (A) or X1 (GOE), then MF < E F < 2 , where MF stands for the median of F. Consequently, for any t > O, 7~(F ~ 2 + at) < 1 - ~ ( t ) < e x p ( - n t 2 / 2 ) ,
(11)
where rr - 1 in the case of llGll and V ~ f o r XI(A) or XI (GOE).
The beauty of the above result lies, in particular, in the inequalities being valid for all n rather than asymptotically. However, Theorem 2.8 shows that asymptotically (11) is not optimal 9We need to mention that the gist of Theorem 2.11 and the argument given above can be extracted from the work of Gordon [83-85]; the same applies to Theorem 2.13 that follows. The relevance of the approach to the "standard" results in random matrices was noticed and publicized by the second author. Similarly as in the preceding section, it is possible to derive from (11) (via the BorelCantelli lemma) results on convergence in probability, for example lim X1 (A ~n)) -- 2
almost surely.
//---+ cx3
Indeed, (11) implies that l i m s u P n ~ )~1 (A(n)) ~< 2 almost surely. The reverse inequality for liminf is even easier: by Theorem 2.3, for any ~ > 0, l i m n ~ 79(3k Xk(A (n)) E [2 -~, 2]) -- 1 and so, for n large enough, the median X1 of X1 (A (n)) is ~> 2 - E. We now get the conclusion by appealing to (5). It is now rather routine to deduce the following more precise version of Theorem 2.7. COROLLARY 2.12. Given n ~ N, set ~,k "-- F -1 ( ~ ) , k - 1. . . . . n, where F is the cumulative distribution function of the Semicircle Law described in Theorem 2.2 (i.e., the n measure #(n) . _ _ n1 Y~k=l 6~k is the 'best' approximant of the semicircle distribution among measures with n atoms). Then lim max [Xk(A(n)) - ~-kl--0 n ~ 1~
Local operator theory, random matrices and Banach spaces
353
An argument just slightly more involved than the proof of Theorem 2.11 allows an analysis of the extreme singular numbers of rectangular Gaussian matrices with independent entries. It has been known for quite a while, cf. [118,179], that as the size of such matrices (appropriately normalized, with entries not necessarily Gaussian) increases with the ratio of the sides approaching/~ E (0, 1), then, as in Theorem 2.3, the empirical measures corresponding to the singular values converge in distribution, or almost surely, to a deterministic measure supported on the interval [1 - ,,/-fl, 1 + v/-fl] and often referred to as the M a r c h e n k o - P a s t u r distribution (or, more recently, in the fi'ee probability context, the free Poisson distribution, cf. the next section). Somewhat later it was determined [80,152, 187,20] that, under appropriate assumptions on the distribution of the entries, the extreme singular values do converge almost surely to the endpoints of the interval above. Here we present the following special but elegant fact. THEOREM 2.13. Given m, n E I~ with m <~ n, p u t 1~ -- m / n and consider the n • m random matrix F whose entries are real, independent Gaussian random variables following the N(O, I / n ) law. Let the s i n g u l a r v a l u e s be sl ( F ) >~ . . . >~ sin(F). Then (12)
1 - w/~ < E s m ( f ) < Ms,(F) < Esl ( F ) < 1 + and consequently, f o r any t > 0,
max{79(s, ( F ) >~ 1 + ~
+ t), V ( s m ( r )
,)1
<~ 1 -
< 1 - q~ (t) < e x p ( - n t 2/2). The proof of the upper estimate in (12) is essentially the same as that of the first assertion of Theorem 2.11. For the lower estimate, we consider the same families: Xu,v "-- Zu| and Y,,v " - Z(,,v) w i t h u E S m-1 , y E S n-1 , but then we set T, -- {(u, v)" y E S n-I }. The hypothesis (c) is satisfied since (10) is an equality if u -- u', and so we may use the second part of Lemma 2.9 to obtain
nl/2E
max uEsm-1
min yES n- 1
(f u, v) <~fR Ixldym(x)- fR Ixldgn(x). m
n
The quantity on the right does not exceed ~ v/-n. (This actually requires some work. What is clear is that the difference between the two expression --+ 0 as m, n --+ ~ . ) Since for any F , max
min ( f u, v) ~ - s m ( F ) ,
btCsm-1 t ~ E S n - I
the first inequality in (12) follows. It remains to appeal again to Theorem 2.6. The arguments of this section can be modified to treat other classes of real matrices, even if the outcomes may be less elegant. It would be nice and potentially very fruitful to find an approach to the complex case(s) that is based on similar ideas. As complex (or symplectic) matrices can be thought of as real matrices with a special structure, the
354
K.R. Davidson and S.J. Szarek
problem at hand is equivalent to considering simple expressions in real Gaussian matrices with matrix coefficients. For example, the complex analogue of G (n) ((ii) on our list) can be identified with a real matrix
~/~
GI
G
] ,(El0] - ~
0
1
|
E0 1] ) 1
0
|
,
(13)
where G and G f are independent copies of G (n). Accordingly, the proper context for the question appears to be that of operator spaces; more about related issues in the next section. In another direction, it is quite clear that variations of the methods of this sections can be applied to a wide range of questions including random factorizations, and particularly, estimating virtually any reasonable norm applied to Gaussian matrices. For example, one could consider for the operator norm with respect to underlying norms on ]1~n different than the Euclidean norm. One just needs to replace S n-1 by the spheres corresponding to the norms in question. This has been noticed early on, see [50,25,84]. However, more often than not, one only gets this way an approximate asymptotic order up to a constant like in Theorem 2.1, rather than the precise asymptotics obtained in this section. We mention here an important question which perhaps did not receive enough publicity in the Banach space circles. PROBLEM 2.14. Show the existence and find lim rt----> o o
Emax~c{-1,1}, (A(n)~, ~) n
The quantity in the numerator, which is close to and of the same order as the norm of A (n) considered as an operator from ~ to ~ , is related to spin glass theory [169, 123]. Even the existence of the limit is not clear. However, it is generally believed that it does exist and equals approximately 1.527. Mimicking the proof of Theorem 2.11 yields x/g-/Tr ~ 1.596 as an upper estimate for every n, and hence for the lim sup, just slightly worse than the best known rigorous upper estimate of 2 - x/e/4 ,~ 1.588 for the latter, due to B. Derrida. The best rigorous lower estimate is 4/Jr ,~ 1.273 [ 1]; we suspect that this one is most susceptible to attack due to improvements in technology in the intervening years. A related, possibly simpler, problem would be calculating the exact asymptotic order of, say, the norm of G (n) considered as an operator from ~pn to ~q. n To the best of our knowledge, this is known only for very special values of p and q. Or, perhaps more appropriately, one could consider the analogous question for the matrices F from Theorem 2.13.
2.4. Random matrices and free probability A significant recent development was the emergence, due largely to efforts of D. Voiculescu, of the area of free probability, and the realization of its connections to random matrices. Even though, a priori, free probability seems relevant only to the macroscopic features of the asymptotic spectral pictures of random matrices, it is hard to overestimate its effect
Local operator theory, random matrices and Banach spaces
355
on clarifying the subject. Several books on the topic appeared in recent years or are in the works and, accordingly, we shall only sketch here the basic ideas and mention a few results and problems that are relevant to the remainder of our discussion. A noncommutative probability space is a pair (A, qg), where A is an algebra, usually over C, with unit I and a state qg. A state is a linear functional verifying qg(1) -- 1. In the classical case, we have an algebra of (measurable) functions on a probability space and the expectation. There are also Banach algebra, C*-algebra etc. probability spaces where the algebra in question is endowed with the appropriate additional structure, and the state q9 respects that structure. Elements of A are thought of as random variables, and the distribution of a random variable x is the linear functional #x on the algebra of C[X] of complex polynomials in variable X defined by # x ( p ) := qg(p(x)). In the C*-algebra context, the distribution of a normal element is actually represented by a measure supported on the spectrum of x (by the Gelfand-Neumark theorem). In the classical context, this measure is necessarily the law of x. One defines similarly joint distributions as functionals on the algebra of polynomials in the appropriate number of noncommuting variables. The convergence in distribution is the w e a k - , convergence: #x,, (p) ~ #x (P) for all polynomials p. This can be defined even if the xn's belong to different probability spaces. A typical statement concerning convergence in distribution is Theorem 2.2. Three important examples of probability spaces are: (i) The C*-algebra 93tn of n x n matrices with the normalized trace Tr; the distribution 1 n of A 6 9Jtn is represented by the measure # a -- n ~-~j--1 6~j(a). (ii) The *-algebra A (n) of random n x n matrices A (co) on some sufficiently rich classical probability space ($2, r , P ) verifying EllA(co)ll p < ~ for all p < cx~, with q9 = ETr. (iii) An operator algebra A c L(7-/) with a vector state qg~(X) := (X~, ~), where ~ 6 7-/ is a norm one vector. The fundamental concept of the theory is that of freeness (or free independence). It is modelled after that of classical independence, which may be restated as follows: commuting subalgebras A l , A2 are independent if q g ( a l a 2 ) = 0 whenever aj E A j with qg(aj) -- 0 for j = 1, 2. By analogy, one says that a family ( A j ) j E J are free if for any reduced product a - - a l a 2 . . . am where neighboring elements come from different A j ' s , one has qg(a) = 0 whenever qg(ak) -- 0 for k = 1 . . . . . m. A family of random variables (or sets of such) is free (respectively, *-free) if the algebras (respectively, *-algebras) generated by them are free. An early result of free probability was a free central limit theorem (CLT): if (Xj)jEN is a sequence of free random variables normalized by qg(Xj) -- O, qg(X2) -- 1, j E 1~, and
satisfying some mild technical conditions such as supjcl ~ Iqg(x~)l < cx~for all k ~ N, then as N --+ cx~, xl+...+xx converges in distribution to the standard semicircular distribution. The mysterious appearance of that distribution in this context was elucidated only after Theorem 2.16 below was proved. It is not clear from the above discussion that nontrivial free random variables actually do exist. A prototype of free algebras is as follows. Let G l, G2 be discrete groups and G = G I * G 2 their free product. The group algebras ~ j = C ( G j ) , i.e., the formal linear combinations of elements of the corresponding group, considered as subalgebras of C ( G ) are free with respect to the state which assigns to an element of C ( G ) the coefficient of the
356
K.R. Davidson and S.J. Szarek
unit e. Equivalently, one may consider C(G) as a canonical subalgebra of L(g2(G)) with the vector state corresponding to ~ = ~e, the unit vector supported at e. However, a more useful model is provided by the creation/annihilation operators on the full Fock space. Given Hilbert space 7-/, the corresponding Fock space f(7-/) is defined as the orthogonal 9 7_/| of all tensor powers of 7-/. The 0th power is identified with CI2, direct sum Y~k>~0 where 12 is a fixed unit vector, which is thought of as an empty tensor product and is called the vacuum vector. The corresponding vector state ~0s2 is referred to as the vacuum state. Given h 6 7-/, one defines a shift operator g(h) by g ( h ) r / = h | r/for elementary tensors r/. We have FACT 2.15. In a probability space (L(f(7-/)), qgs2) the following hold (i) If (~"[j)jEJ is a family of orthogonal subspaces ofT-t, then the corresponding family ({t~(h): h E 7-[j })jEJ of subsets of L(gt'(7-/)) is ,-free. (ii) If h E 7-[ is a norm one vector, then s(h) := g(h) + g(h)* has the standard semicircular distribution. The crucial link between random matrices and free probability is provided by the following result of Voiculescu ([ 176], see also [ 177]): THEOREM 2.16. For each n E N, let (A j(n)~J j
EJ
be independent copies of the random n • n
matrix A (n), considered as elements of the noncommutative probability space (A (n) , E Tr) defined above. Then, as n --+ oo, (A~))jEJ converges in distribution to (Sj)jEJ, a family of free random variables, each of which has the standard semicircular distribution. The same is true for GOE, GUE and any ensembles of random matrices verifying the hypotheses of Theorem 2.2. The fact that, for any j ~ J, A j(n) converges in distribution to sj is of course equivalent to Theorem 2.2. The new ingredient is the asymptotic freeness of large random matrices. It is worthwhile emphasizing that this asymptotic freeness, in combination with the free CLT and the infinite divisibility of the Gaussian distribution, does imply that the only possible limit distribution is semicircular. As we indicated, free probability in general and Theorem 2.16 in particular open new vistas on the subject of random matrices. For example, if P is a polynomial in m noncommuting variables, it follows that, for any v E N, lim ETr(P(A(1 n) . . . . . A(mn))v)
- - (fl$2 ( P ( s 1
.....
Sm)V),
(14)
t/---+ o o
where Aj(n) and S j a r e as in Theorem 2.16. In particular, if the polynomial P is formally selfadjoint (the variables themselves are assumed to be selfadjoint here), this leads to a measure concentrated on the spectrum of the bounded selfadjoint operator P(Sl . . . . . Sm) as the asymptotic spectral distribution of the random matrix P (A (1n) . . . . . A (mn)).Asymptotic means a priori in the sense of Theorem 2.2, but proceeding as in Section 2.2, one deduces: THEOREM 2.17. In the notation of Theorem 2.16, let P be a selfadjoinf polynomial in m noncommuting variables and let n ~ N. Then there exist scalars )~1, )~2. . . . . )~n such that, for k = 1 . . . . . n and t >>.O,
357
Local operator theory, random matrices and Banach spaces
7 (l k
A ~ ) ) ) - ~'kl >~ t) < C e x p ( - c p n min{t 2, t2/d})
(~5)
where d is the degree of P, c e > 0 depends only on P and C is a universal constant. One necessarily has maxl<~k~
~
9 o
.
PROBLEM 2.18. In the notation of Theorem 2.17, does )~j converge to the maximum of the spectrum of # e ? An affirmative answer would imply the almost sure convergence of the eigenvalues )~k(P(A(I n) . . . . . A ~ ) ) to the same quantity and, more importantly, the fact that, asymptotically, (A(1n), . . . , A~ )) ar~ equivalent to (sl, . .., Sm) not only in distribution as in (14) or Theorem 2.17, but also in ~he C*-algebra sense. Problem 2.18 is of similar nature as Problem 2.14: in both cases w~: do know, for somewhat similar reasons, the order of growth of the quantity in question, bm we do not know exact asymptotics. However, the former seems to be much more accessible than the latter because of direct applicability of spectral methods. This is because, in the Euclidean case, []B[[ ~ (Tr(B~)) 1/~ if B is an n x n matrix and v is an even number such that v/log n is large, and so the problem reduces to properly
358
K.R. Davidson and S.J. Szarek
estimating E T r ( P ( A I n) , . . . , A(mn))v) for such v, n, in principle a straightforward combinatorial question. This approach, already present in the original Wigner's paper [ 184], was exploited by a number of authors, perhaps in the most sophisticated way in [38,89]. Another possibility would be to somehow use the ideas behind Theorem 2.7 or 2.11 to show by an a priori argument that )~k's of Theorem 2.17 vary slowly with k. This could be interesting for other reasons as the distribution of quantities like sl (G (n)) - s2(G (n)) are of interest in the theory of computational complexity, cf. [ 162]. (It is also conceivable that to efficiently study the gaps in the spectrum in the spirit of Theorem 2.5, one needs to consider functions of matrices and not only matrices as such.) For the quantity mentioned here, the needed information can probably be extracted from the methods of [ 171 ]. But, to our knowledge, no careful examination of the relevant arguments was made. A nontrivial but possibly accessible test cases for any approach would be to find the exact asymptotic behavior of II(G(~))~II for v ~> 2 or IIU + V + U* + V*ll, where U, V are independent and uniformly distributed on U(n) or SO(n). The predictions given by the free probability are v/(V + 1)v+l/vv ,~ ~ for large v, and 2V/-3. Both were actually determined before the era of free probability, see [182,104]. The latter reference treat the random walk on the free group, see [ 178] a clarification of the connection. A known (asymptotic) estimate on [l(G(n))~[[ is v + 1 [19].
2.5. Random vs. explicit constructions It is a quite frequent occurrence that existence of mathematical objects possessing a certain property is shown via nonconstructive methods (the probabilistic method). Roughly speaking, one produces a random variable whose values are those objects and then proves that the property in question is satisfied with nonzero (cf. [2]), or close to 1, probability. Two fields where this principle has been successfully applied are combinatorics and analysis, particularly harmonic analysis and local theory of Banach spaces. Many developments in the latter area are described elsewhere in this collection [ 117,81,41 ], and some spillovers to local operator theory were mentioned in this survey. Here we shall introduce only the details needed to address some philosophical aspect of the issue. We start by recalling a remarkable result of Kashin [101], cf. [160,164], motivated by questions in approximation theory, which roughly asserts that the space s is an orthogonal sum (in the s sense) of two nearly Euclidean subspaces. More precisely: THEOREM 2.19. Given n -- 2m e N, there exist two orthogonal m-dimensional subspaces El, E2 C ]1~n s u c h that 1
1
IIx 112 ~ - - 7 IIx II1 ~ IIx 112 for all Xi E E i, i -- 1, 2. ~/n
(16)
Moreover, for large n, this holds for nearly all decompositions E1 • E2 with respect to the Haar measure on the Grassmann manifold Gn,m. The existence of such a decomposition was surprising because, when considered on the entire space IRn, the ratio between the ~1 and e e norms varies between 1 and v/ft.
Local operator theory, random matrices and Banach spaces
359
Because of the wealth of examples of various objects (e.g., random Banach spaces) arising from or related to the one above, it would be of interest to have, for starters, an explicit example of a Kashin decomposition. Having explicit Ek's would likely lead to more natural examples of finite dimensional Banach spaces with various extremal properties than the ones obtained by the probabilistic method. Indeed, an explicit example must have an explicit reason, and this should presumably be reflected by a presence of some additional structure. It may also conceivably lead to some useful algorithms. It may be the right place to comment here on what exactly we mean by an explicit construction. We shall not give a general definition, but, in the case at hand, admissible descriptions of E1 9 E2 would include an algorithm yielding, for a given n, a basis of E1 or the matrix of the orthogonal projection onto El. That algorithm would need to have a reasonable worst case performance, preferably polynomial in n. The worst case requirement is needed to exclude a strategy that would involve choosing E1 @ E2 at random and then somehow efficiently verifying whether it satisfies (16). We note that, at least at the first sight, even checking (16) for a given Ek appears to be an exponentially hard problem. To the best of our knowledge, the largest explicit subspace E of g'~, for which an assertion of type (16) holds is of dimension approximately ~ as opposed to m/2 in Theorem 2.19 - a long way to go. This follows, e.g., from the construction in [ 147], using finite fields and difference sets (i.e., the so called finite geometries), and giving, for an even integer p ~> 4, an exact Ap-set. This leads to another question: finding explicit exact Ap-sets for other values of p; for probabilistic results see [40,165] and cf. [41]. An example of a probabilistic construction followed by an explicit one is the work on approximation of quasidiagonal operators mentioned in Section 1.3 [159,175,181]. The explicit approach used representations of groups with property T. It has been observed very recently by Benveniste and the second author [24] that the argument of [ 175] can be refined to yield explicit matrices poorly approximable by reducible matrices (cf. [95]), also mentioned in Section 1.3. Their argument suggests specific subspaces of g~n as candidates for E from the previous paragraph (a rather "long shot", but nevertheless a good starting point). A somewhat similar example comes from another field, involving random graphs. Following the seminal work of Erd6s [73], this became a powerful technique to show existence of graphs with various extremal properties. As some questions about graphs are very practical optimization problems (like design of a network), it was important to have explicit solutions. This is particularly relevant because verifying that a given large graph has some required property is sometimes computationally not feasible. For some important questions, explicit solutions were found, see [120,116] and much earlier work [119]. The constructions were based on properties of some arithmetic groups, sophisticated tools from arithmetic geometry or, again, the property T. Some of the graphs in question (Ramanujan graphs) were already mentioned at the end of Section 1.4. The final topic we propose to analyse is suggested by the preceding section on free probability. Theorem 2.16 can be interpreted as an assertion that two large random matrices are nearly free, in the sense of the pattern of the moments of monomials in the two matrices with respect to the normalized trace being approximately the same as it would have been the case if the two matrices were free. In fact, the same is true for, roughly speaking, one fixed and one random matrix [ 177]. A natural problem is to come up with explicit matrices
360
K.R. Davidson and S.J. Szarek
having the same property. (It cannot happen that two finite matrices are exactly free, except for the trivial cases.) Some examples to that effect were produced in [176] and [151 ]. While they are not fully satisfactory in some respects, it is nevertheless conceivable that they may also constitute a step towards an explicit Kashin decomposition (even though they do not work directly). The ideas from [37], involving representations of symmetric groups, may also be relevant here. This and the preceding two paragraphs do broadly suggest some of the areas of mathematics that might be pertinent to other explicit constructions.
Acknowledgments The first author was partially supported by a grant from the Natural Science and Engineering Research Council (Canada). The second author was partially supported by grants from the National Science Foundation (USA). The authors thank Professor L. Pastur for helpful comments on the second part of the article.
References [1] M. Aizenman, J.L. Lebowitz and D. Ruelle, Some rigorous results on the Sherrington-Kirkpatrick spin glass model, Comm. Math. Phys. 112 (1987), 3-20. [2] N. Alon, Probabilistic proofs of existence of rare events, Geometric Aspects of Functional Analysis (198788), Lecture Notes in Math. 1376, Springer, Berlin (1989), 186-201. [3] N.T. Andersen, Similarity of continuous nests, Bull. London Math. Soc. 15 (1983), 131-132. [4] J. Anderson, Extensions, restrictions, and representations of states on C*-algebras, Trans. Amer. Math. Soc. 249 (1979), 303-329. [5] J. Anderson, Pathology in the Calkin algebra, J. Oper. Theory 2 (1979), 159-167. [6] J. Anderson, Extreme points of positive linear maps on B(H), J. Funct. Anal. 31 (1979), 195-217. [7] J. Anderson, A C*-algebra A for which Ext(A) is not a group, Ann. Math. 107 (1978), 455-458. [8] J.R. Angelos, C.C. Cowen and S.K. Narayan, Triangular truncation andfinding the norm ofa Hadamard multiplier, Lin. Alg. Appl. 170 (1992), 117-135. [9] C. Apostol, C. Foia~ and D. Voiculescu, Some results on nonquasitriangular operators. VI, Rev. Roum. Math. Pures et Appl. 18 (1973), 1473-1494. [ 10] C. Apostol, C. Foia~ and D. Voiculescu, On the norm-closure of nilpotents. II, Rev. Roum. Math. Pures et Appl. 19 (1974), 549-577. [11] C. Apostol and N. Salinas, Nilpotent approximations and quasinilpotent operators, Pacific J. Math. 61 (1975), 327-337. [ 12] L. Arnold, On the asymptotic distribution of the eigenvalues of random matrices, J. Math. Anal. Appl. 20 (1967), 262-268. [13] W.B. Arveson, Interpolation problems in nest algebras, J. Funct. Anal. 3 (1975), 208-233. [14] W.B. Arveson, Operator algebras and invariant subspaces, Ann. Math. 100 (1974), 433-532. [15] E. Azoff and C. Davis, On distances between unitary orbits of self-adjoint operators, Acta Sci. Math. (Szeged) 47 (1984), 419-439. [16] Z.D. Bai, Convergence rate of expected spectral distributions of large random matrices. I. Wigner matrices, Ann. Probab. 21 (1993), 625-648. [17] Z.D. Bai, B. Miao and J. Tsay, Remarks on the convergence rate of the spectral distributions of Wigner matrices, J. Theoret. Probab. 12 (1999), 301-311. [18] Z.D. Bai and Y.Q. Yin, Necessary and sufficient conditions for almost sure convergence of the largest eigenvalue ofa Wigner matrix, Ann. Probab. 16 (1988), 1729-1741. [19] Z.D. Bai and Y.Q. Yin, Limiting behavior of the norm of products of random matrices and two problems of Geman-Hwang, Probab. Theory Related Fields 73 (1986), 555-569.
Local operator theory, random matrices and Banach spaces
361
[20] Z.D. Bai and Y.Q. Yin, Limit of the smallest eigenvalue of a large-dimensional sample covariance matrix, Ann. Probab. 21 (1993), 1275-1294. [21] D. Bakry and M. Ledoux, L~vy-Gromov's isoperimetric inequality for an infinite-dimensional diffusion generator, Invent. Math. 123 (2) (1996), 259-281. [22] G. Ben Arous and A. Guionnet, Large deviations for Wigner's law and Voiculescu's non-commutative entropy, Probab. Theory Related Fields 108 (1997), 517-542. [23] G. Bennett, V. Goodman and C.M. Newman, Norms of random matrices, Pacific J. Math. 59 (1975), 359365. [24] E.J. Benveniste and S.J. Szarek, Property T and irreducibility of matrices, Manuscript in preparation. [25] Y. Benyamini and Y. Gordon, Randomfactorization of operators between Banach spaces, J. Analyse Math. 39 (1981), 45-74. [26] H. Bercovici, Hyper-reflexivity and the factorization of linear functionals, J. Funct. Anal. 158 (1998), 242-252. [27] I.D. Berg, An extension of the Weyl-von Neumann theorem to normal operators, Trans. Amer. Math. Soc. 160 (1971), 365-371. [28] I.D.Berg, On approximation of normal operators by weighted shifts, Michigan Math. J. 21 (1974), 377383. [29] I.D. Berg and K.R. Davidson, A quantitative version of the Brown-Douglas-Fillmore Theorem, Acta Math. 166 (1991), 121-161. [30] K. Berman, H. Halpern, V. Kaftal and G. Weiss, Matrix norm inequalities and the relative Dixmier property, Int. Eq. Oper. Theory 11 (1988), 28-48. [31] R. Bhatia, Perturbation Bounds for Matrix Eigenvalues, Pitman Research Notes in Mathematics Series 162, Longman Scientific and Technical Pub. Co., London (1987). [32] R. Bhatia, Analysis of spectral variation and some inequalities, Trans. Amer. Math. Soc. 272 (1982), 323-332. [33] R. Bhatia, Matrix Analysis, Graduate Texts in Math. 169, Springer-Verlag, New York (1997). [34] R. Bhatia, C. Davis and E Koosis, An extremal problem in Fourier analysis with applications to operator theory, J. Funct. Anal. 82 (1) (1989), 138-150. [35] R. Bhatia, C. Davis and A. McIntosh, Perturbation of spectral subspaces and solution of linear operator equations, Lin. Alg. Appl. 52/53 (1983), 45-67. [36] R. Bhatia and J. Holbrook, Short normal paths and spectral variation, Lin. Alg. Appl. 94 (1985), 377-382. [37] P. Biane, Representations of symmetric groups and free probability, Adv. Math. 138 (1998), 126-181. [38] S.G. Bobkov and F. G6tze, Exponential integrability and transportation cost related to logarithmic Sobolev inequalities, J. Funct. Anal. 163 (1) (1999), 1-28. [39] C. Borell, The Brunn-Minkowski inequality in Gauss space, Invent. Math. 30 (1975), 207-216. [40] J. Bourgain, Bounded orthogonal systems and the A (p)-set problem, Acta Math. 162 (1989), 227-245. [41] J. Bourgain, A p sets in analysis: results, problems and related aspects, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 195-232. [42] J. Bourgain and L. Tzafriri, Invertibility of "large" submatrices with applications to the geometry of Banach spaces and harmonic analysis, Israel J. Math. 57 (1987), 137-224. [43] J. Bourgain and L. Tzafriri, Restricted invertibility of matrices and applications, Analysis at Urbana II, London Math. Soc. Lect. Notes 138, E.R. Berkson, N.T. Peck and J.J. Uhl, eds, Cambridge Univ. Press (1989), 61-107. [44] J. Bourgain and L. Tzafriri, On a problem of Kadison and Singer, J. Reine Angew. Math. 420 (1991), 1-43. [45] A. Boutet de Monvel and M. Shcherbina, On the norm of random matrices, Math. Notes 57 (1995), 475484. [46] B.V. Bronk, Accuracy of the semicircle approximation for the density of eigenvalues of random matrices, J. Math. Phys. 5 (1964), 215-220. [47] L.G. Brown, R.G. Douglas and EA. Fillmore, Unitary equivalence modulo the compact operators and extensions of C*-algebras, Proceedings of a Conference on Operator Theory, Halifax, NS 1973, Lect. Notes in Math. 345, Springer-Verlag, Berlin (1973), 58-128. [48] L.G. Brown, R.G. Douglas and EA. Fillmore, Extensions of C*-algebras and K-homology, Ann. Math. 105 (1977), 265-324.
362
K.R. Davidson and S.J. Szarek
[49] S.W. Brown, Some invariant subspacesfor subnormal operators, Int. Eq. Oper. Theory 1 (1978), 310-333. [50] S. Chevet, S~ries de variables al~atoires gaussiennes gt valeurs dans E ~ F. Application aux produits d'espaces de Wiener abstraits, S6minaire sur la G6om6trie des Espaces de Banach (1977-1978), Exp. No. 19, 15 pp., l~cole Polytech., Palaiseau (1978). [51] M.D. Choi, Almost commuting matrices need not be nearly commuting, Proc. Amer. Math. Soc. 102 (1988), 529-533. [52] E. Christensen, Perturbations of operator algebras II, Indiana Univ. Math. J. 26 (1977), 891-904. [53] M. Dadarlat, On the approximation of quasidiagonal C*-algebras, J. Funct. Anal. 167 (1999), 69-78. [54] K.R. Davidson, Similarity and compact perturbations ofnest algebras, J. Reine Angew. Math. 348 (1984), 72-87. [55] K.R. Davidson, Almost commuting Hermitian matrices, Math. Scand. 56 (1985), 222-240. [56] K.R. Davidson, The distance between unitary orbits of normal operators, Acta Sci. Math. (Szeged) 50 (1986), 213-223. [57] K.R. Davidson, The distance to the analytic Toeplitz operators, Illinois J. Math. 31 (1987), 265-273. [58] K.R. Davidson, Nest Algebras, Pitman Research Notes in Mathematics Series 191, Longman Scientific and Technical Pub. Co., London (1988). [59] K.R. Davidson, Finite dimensional problems in operator theory, The Gohberg Anniversary Collection, Operator Theory: Advances and Applications 40, BirkhaUser Verlag, Basel (1989), 187-201. [60] K.R. Davidson, C*-Algebras by Example, Fields Institute Monograph Series 6, Amer. Math. Soc., Providence, RI (1996). [61] K.R. Davidson, Polynomially bounded operators, NATO ASI Proceedings, Samos, Greece, August 1996, Operator Algebras and Applications, Kluwer Academic, Dordrecht (1997), 145-162. [62] K.R. Davidson, D.A. Herrero and N. Salinas, Quasidiagonal operators, approximation, and C*-algebras, Indiana Univ. J. Math. 38 (1989), 973-998. [63] K.R. Davidson and M. Ordower, Some exact distance constants, Lin. Alg. Appl. 208/209 (1994), 37-55. [64] K.R. Davidson and V.I. Paulsen, Polynomially bounded operators, J. Reine Angew. Math. 487 (1997), 153-170. [65] K.R. Davidson and D.R. Pitts, Approximate unitary equivalence of completely distributive commutative subspace lattices, Int. Eq. Oper. Theory 22 (1995), 196-211. [66] K.R. Davidson and D.R. Pitts, Invariant subspaces and hyper-reflexivity for free semigroup algebras, Proc. London Math. Soc. 78 (1999), 401-430. [67] K.R. Davidson and S.C. Power, Failure ofthe distance formula, J. London Math. Soc. 32 (1984), 157-165. [68] C. Davis, W.M. Kahan and W.E Weinberger, Norm preserving dilations and their applications to optimal error bounds, SIAM J. Numer. Anal. 19 (1982), 445-469. [69] J.A. Deddens and EA. Fillmore, Reflexive linear transformations, Lin. Alg. Appl. 10 (1975), 89-93. [70] P. Deift, T. Kriecherbauer, K.T.-R. McLaughlin, S. Venakides and X. Zhou, Uniform asymptotics for polynomials orthogonal with respect to varying exponential weights and applications to universality questions in random matrix theory, Comm. Pure Appl. Math. 52 (11) (1999), 1335-1425. [71 ] V.A. Dmitrovskff, On the integrability of the maximum and the local properties of Gaussian fields, Probability Theory and Mathematical Statistics 1 (Vilnius, 1989), Mokslas, Vilnius (1990), 271-284. [72] K. Dykema, On certain free product factors via an extended matrix model, J. Funct. Anal. 112 (1993), 31-60. [73] E Erd6s, Graph theory and probability, Canad. J. Math. 11 (1959), 34-38. Graph theory and probability //, Canad. J. Math. 13 (1961), 346-352. [74] R. Exel and T.A. Loring, Determinants and almost commuting matrices, Proc. Amer. Math. Soc. 106 (1989), 913-915. [75] X. Fernique, Des r~sultats nouveaux sur les processus gaussiens, C. R. Acad. Sci. Paris S6r. A 278 (1974), 363-365. [76] X. Fernique, Fonctions Al~atoires Gaussiennes Vecteurs Al~atoires Gaussiennes, Publications Centre de Recherches Math6matiques (1997). [77] P.J. Forrester and A.M. Odlyzko, Gaussian unitary ensemble eigenvalues and Riemann ~ function zeros: a nonlinear equation for a new statistic, Phys. Rev. E (3) 54 (1996), R4493-R4495. [78] P. Friis and M. RCrdam, Almost commuting self-adjoint matrices- a short proof of Huaxin Lin's theorem, J. Reine Angew. Math. 479 (1996), 121-131.
Local operator theory, random matrices and Banach spaces
363
[79] P. Friis and M. Rc~rdam, Approximation with normal operators with finite spectrum, and an elementary proof of the Brown-Douglas-Fillmore theorem, Preprint (1998). [80] S. Geman, A limit theorem for the norm of random matrices, Ann. Probab. 8 (1980), 252-261. [81] A.A. Giannopoulos and V.D. Milman, Euclidean structure infinite dimensional normed spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 707-779. [82] V.L. Girko, Theory of Random Determinants, Mathematics and its Applications (Soviet Series) 45, Kluwer Academic, Dordrecht (1990) (Translated from the Russian). [83] Y. Gordon, On Dvoretzky's theorem and extensions of Slepian's lemma, Israel Seminar on Geometrical Aspects of Functional Analysis (1983/84), II, Tel Aviv Univ., Tel Aviv (1984). [84] Y. Gordon, Some inequalities for Gaussian processes and applications, Israel J. Math. 50 (1985), 265-289. [85] Y. Gordon, Majorization of Gaussian processes and geometric applications, Probab. Theory Related Fields 91 (1992), 251-267. [86] M. Gromov and V.D. Milman, A topological application of the isoperimetric inequality, Amer. J. Math. 105 (1983), 843-854. [87] A. Guionnet and O. Zeitouni, Concentration of the spectral measure for large matrices, Electron. Comm. Probab. 5 (2000), 119-136 (electronic). [88] U. Haagerup and S. Thorbj~rusen, Random matrices with complex Gaussian entries, Preprint, Odense University (1998), available at ftp://ftp.imada.sdu.dk/pub/papers/pp-1998/07.ps.gz. [89] U. Haagerup and S. ThorbjCrusen, Random matrices and K-theory for exact C*-algebras, Doc. Math. 4 (1999), 341-450. [90] D. Hadwin, Triangular truncation and normal limits of nilpotent operators, Proc. Amer. Math. Soc. 123 (1995), 1741-1745. [91] D.A. Herrero, Approximation of Hilbert Space Operators. I, Research Notes in Math. 72, Pitman Books, London (1982). [92] D.A. Herrero, Normal limits ofnilpotent operators, Indiana Univ. Math. J. 23 (1974), 1097-1108. [93] D.A. Herrero, Unitary orbits of power partial isometries and approximation by block-diagonal nilpotents, Topics in Modern Operator Theory, Fifth Int. Conf. on Oper. Theory, Timi~oara and Herculane (Romania 1980), Operator Theory: Advances and Applications 2, Birkhatiser-Verlag, Basel (1981), 171-210. [94] D.A. Herrero, A trace obstruction to approximation by block diagonal operators, Amer. J. Math. 108 (1986), 451-484. [95] D.A. Herrero and S.J. Szarek, How well can an n • n matrix be approximated by reducible ones?, Duke J. Math. 53 (1986), 233-248. [96] A.J. Hoffman and H.W. Wielandt, The variation of the spectrum of a normal matrix, Duke J. Math. 20 (1953), 37-39. [97] K. Johansson, Universality of the local spacing distribution in certain ensembles of hermitian Wigner matrices, Preprint available at http://xxx, lanl. gov/ab s/math- ph/0006020. [98] W.B. Johnson and G. Schechtman, Finite dimensional subspaces of L p, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 837-870. [99] R.V. Kadison and I.M. Singer, Extensions ofpure states, Amer. J. Math. 81 (1959), 383-400. [ 100] J.E Kahane, Une indgalitd du type de Slepian et Gordon sur les processus gaussiens, Israel J. Math. 55 (1) (1986), 109-110. [101] B.S. Kashin, The widths of certain finite-dimensional sets and classes of smooth functions, Izv. Akad. Nauk SSSR Ser. Mat. 41 (1977), 334-351 (Russian). [102] B.S. Kashin, A property of bilinearforms, Soobshch. Akad. Nauk Gruzin. SSR 97 (1980), 29-32 (Russian). [103] B.S. Kashin, Some properties ofmatrices of bounded operators from the space l~ into l~, Izv. Akad. Nauk Armyan. SSR Ser. Mat. 15 (1980), 379-394 (Russian). English translation: Soviet J. Contemporary Math. Anal. 15 (1980), 44-57 (1982). [104] H. Kesten, Symmetric random walks on groups, Trans. Amer. Math. Soc. 92 (1959), 336-354. [ 105] M. Krivelevich and V.H. Vu, On the concentration of eigenvalues of random symmetric matrices, Preprint available at http://xxx.lanl.gov/abs/math-ph/0009032. [106] S. Kwapiefi, Decoupling inequalities f or polynomial chaos, Ann. Probab. 15 (1987), 1062-1071. [ 107] S. Kwapiefi, A remark on the median and the expectation of convex functions of Gaussian vectors, Probability in Banach Spaces 9 (Sandjberg, 1993), Progr. Probab. 35, Birkhafiser, Boston, MA (1994), 271-272.
364
K.R. Davidson and S.J. Szarek
[108] E.C. Lance, Cohomology and perturbations of nest algebras, Proc. London Math. Soc. 423 (1981), 334356. [ 109] D.R. Larson, Hyperreflexivity and a dual product construction, Trans. Amer. Math. Soc. 294 (1986), 7988. [ 110] M. Ledoux, Isop~rim~trie et in~galit~s de Sobolev logarithmiques gaussiennes, C. R. Acad. Sci. Paris S6r. I Math. 306 (2) (1988), 79-82 (French). [111] M. Ledoux, Concentration of measure and logarithmic Sobolev inequalities, S6minaire de Probabilit6s, XXXIII, Lecture Notes in Math. 1709, Springer, Berlin (1999), 120-216. [ 112] M. Ledoux and M. Talagrand, Probability in Banach Spaces. Isoperimetry and Processes, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) 23, Springer-Verlag, Berlin (1991). [113] H. Lin, Almost commuting selfadjoint matrices and applications, Operator Algebras and Their Applications (Waterloo, ON, 1994/1995), Fields Inst. Commun. 13, Amer. Math. Soc., Providence, RI (1997), 193-233. [114] T.A. Loring, K-theory and asymptotically commuting matrices, Canad. J. Math. 40 (1988), 197-216. [ 115] T.A. Loring, Lifting Solutions to Perturbing Problems in C*-Algebras, Fields Institute Monograph Series 8, Amer. Math. Soc., Providence, RI (1997). [116] A. Lubotzky, R. Phillips and P. Sarnak, Ramanujan graphs, Combinatorica 8 (1988), 261-277. [ 117] P. Mankiewicz and N. Tomczak-Jaegermann, Quotients offinite-dimensional Banach spaces; random phenomena, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), to appear. [118] V.A. Mar~enko and L.A. Pastur, Distribution of eigenvalues in certain sets of random matrices, Mat. Sb. (N.S.) 72 (1967), 507-536 (Russian). [ 119] G.A. Margulis, Explicit constructions of expanders, Problemy Pereda6i Informacii 9 (1973), 71-80 (Russian). English translation: Problems Inform. Transmission 9 (1973), 325-332. [120] G.A. Margulis, Explicit group-theoretic constructions of combinatorial schemes and their applications in the construction of expanders and concentrators, Problemy Peredachi Informatsii 24 (1988), 51-60 (Russian); English translation: Problems Inform. Transmission 24 (1988), 39-46. [ 121] M.L. Mehta, Random Matrices, 2nd edn, Academic Press, Boston, MA (1991). [122] M.L. Mehta and M. Gaudin, On the density ofeigenvalues of a random matrix, Nuclear Phys. 18 (1960), 420-427. [123] M. M6zard, G. Parisi and M.A. Virasoro, Spin Glass Theory and Beyond, World Scientific Lecture Notes in Physics 9, World Scientific Publishing, Teaneck, NJ (1987). [124] V.D. Milman and G. Schechtman, Asymptotic Theory of Finite Dimensional Normed Spaces. With an Appendix by M. Gromov, Lecture Notes Math. 1200, Springer Verlag, Berlin (1986). [125] L. Mirsky, Symmetric gauge functions and unitarily invariant norms, Quart. J. Math. (2) 11 (1960), 50-59. [126] H.L. Montgomery, The pair correlation of zeros of the zeta function, Analytic Number Theory (Proc. Sympos. Pure Math. XXIV, St. Louis Univ., St. Louis, MO, 1972), Amer. Math. Soc., Providence, RI (1973), 181-193. [127] A.M. Odlyzko, On the distribution of spacings between zeros of the zeta function, Math. Comp. 48 (1987), 273-308. [128] R.E Olin and J.E. Thomson, Algebras of subnormal operators, J. Funct. Anal. 37 (1980), 271-301. [ 129] E Otto and C. Villani, Generalization ofan inequality by Talagrand and links with the logarithmic Sobolev inequality, J. Funct. Anal. 173 (2) (2000), 361-400. [130] S.K. Parrott, On a quotient norm and the Sz.-Nagy Foia~ lifting theorem, J. Funct. Anal. 30 (1978), 311328. [131] L.A. Pastur, Spectra of random selfadjoint operators, Uspehi Mat. Nauk 28 (1) (1973), 3-64 (Russian). [ 132] L.A. Pastur, A simple approach to the global regime of the random matrix theory, Mathematical Results in Statistical Mechanics, S. Miracle-Sole, J. Ruiz and V. Zagrebnov, eds, World Scientific, Singapore (1999), 429-454. [ 133] L.A. Pastur and M. Shcherbina, Universality of the local eigenvalue statistics for a class of unitary invariant matrix ensembles, J. Statist. Phys. 86 (1-2) (1997), 109-147. [134] V.I. Paulsen, Completely Bounded Maps and Dilations, Pitman Research Notes in Mathematics Series 146, Longman Scientific and Technical Pub. Co., London (1986). [ 135] C.M. Pearcy and A.L. Shields, Almost commuting matrices, J. Funct. Anal. 30 (1978), 311-328.
Local operator theory, random matrices and Banach spaces
365
[136] G. Pisier, Probabilistic methods in the geometry of Banach spaces, C.I.M.E. Varenna 1985, Lecture Notes in Math. 1206, Springer Verlag, Berlin (1986), 167-241. [ 137] G. Pisier, Factorization of Linear Operators and Geometry of Banach Spaces, CBMS Regional Conference Series in Mathematics 60, Amer. Math. Soc., Providence, RI (1986). [138] G. Pisier, A polynomially bounded operator on Hilbert space which is not similar to a contraction, J. Amer. Math. Soc. 10 (1997), 351-369. [ 139] G. Pisier, Similarity Problems and Completely Bounded Maps, Lect. Notes in Math. 1618, Springer-Verlag, Berlin (1996). [140] G. Pisier, Operator spaces and similarity problems, Proc. Inter. Congress Math., Vol. I (Berlin, 1998), Doc. Math., Extra Vol. I (1998), 429-452. [141] G. Pisier, Operator spaces, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), to appear. [142] D.R. Pitts, Close CSL algebras are similar, Math. Ann. 300 (1994), 149-156. [143] G. Popescu, A generalization of Beurling's theorem and a class of reflexive algebras, J. Operator Theory 41 (1999), 391-420. [ 144] S.C. Power, The distance to upper triangular operators, Math. Proc. Camb. Phil. Soc. 88 (1980), 327-329. [145] J.R. Ringrose, On some algebras of operators, Proc. London Math. Soc. 15 (1965), 61-83. [146] S. Rosenoer, Distance estimates for von Neumann algebras, Proc. Amer. Math. Soc. 86 (1982), 248-252. [147] W. Rudin, Trigonometric series with gaps, J. Math. Mech. 9 (1960), 203-227. [148] Z. Rudnick and P. Sarnak, Zeros of principal L-functions and random matrix theory, A Celebration of John F. Nash, Jr., Duke Math. J. 81 (1996), 269-322. [149] N. Salinas, Homotopy invariance of Ext(A), Duke Math. J. 44 (1977), 777-794. [150] G. Schechtman, Concentration, results and applications, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), to appear. [151] D. Shlyakhtenko, Limit distributions of matrices with bosonic and fermionic entries, Free Probability Theory (Waterloo, ON, 1995), Fields Inst. Commun. 12, Amer. Math. Soc., Providence, RI (1997), 241252. [152] J.W. Silverstein, The smallest eigenvalue of a large-dimensional Wishart matrix, Ann. Probab. 13 (1985), 1364-1368. [ 153] Ya.G. Sinai and A.B. Soshnikov, A refinement of Wigner's semicircle law in a neighborhood of the spectrum edge for random symmetric matrices, Funk. Anal. i Prilozhen. 32 (1998), 56-79 (Russian). English translation: Funct. Anal. Appl. 32 (1998), 114-131. [154] A. Soshnikov, Universality at the edge of the spectrum in Wigner random matrices, Comm. Math. Phys. 207 (3) (1999), 697-733. [155] V.S. Sunder, Distance between normal operators, Proc. Amer. Math. Soc. 84 (1982), 483-484. [ 156] S.J. Szarek, The finite dimensional basis problem with an appendix on nets of Grassmann manifolds, Acta Math. 151 (1983), 153-179. [157] S.J. Szarek, Nets of Grassmann manifold and orthogonal groups, Proceedings of Banach Space Workshop, University of Iowa Press (1982), 169-185. [158] S.J. Szarek, On almost commuting Hermitian operators, Proc. Seventh GPOTS (Lawrence, KS, 1987), Rocky Mountain J. Math. 20 (1990), 581-589. [159] S.J. Szarek, An exotic quasidiagonal operator, J. Funct. Anal. 89 (1990), 274-290. [160] S.J. Szarek, On Kashin's almost Euclidean orthogonal decomposition of g~]~,Bull. Acad. Polon. Sci. S6r. Sci. Math. Astronom. Phys. 26 (1978), 691-694. [161] S.J. Szarek, Spaces with large distance to gn and random matrices, Amer. J. Math. 112 (1990), 899-942. [162] S.J. Szarek, Condition numbers of random matrices, J. Complexity 7 (1991), 131-149. [163] S.J. Szarek, Metric entropy of homogeneous spaces, Quantum Probability (Gdafisk, 1997), Banach Center Publ. 43, Polish Acad. Sci., Warsaw (1998), 395-410. [164] S.J. Szarek and N. Tomczak-Jaegermann, On nearly Euclidean decomposition for some classes of Banach spaces, Comp. Math. 40 (1980), 367-385. [165] M. Talagrand, Sections of smooth convex bodies via majorizing measures, Acta Math. 175 (1995), 273300. [166] M. Talagrand, New concentration inequalities in product spaces, Invent. Math. 126 (1996), 505-563.
366
K.R. Davidson and S.J. Szarek
[167] M. Talagrand, Concentration of measure and isoperimetric inequalities in product spaces, Inst. Hautes l~tudes Sci. Publ. Math. 81 (1995), 73-205. [168] M. Talagrand, Transportation cost for Gaussian and other product measures, Geom. Funct. Anal. 6 (3) (1996), 587-600. [ 169] M. Talagrand, Huge random structures and meanfield models for spin glasses, Proc. Inter. Congress Math., Vol. I (Berlin, 1998), Doc. Math., Extra Vol. I (1998), 507-536. [170] C.A. Tracy and H. Widom, Level-spacing distributions and the Airy kernel, Comm. Math. Phys. 159 (1994), 151-174. [171] C.A. Tracy and H. Widom, On orthogonal and symplectic matrix ensembles, Comm. Math. Phys. 177 (1996), 727-754. [172] D. Voiculescu, A non-commutative Weyl-von Neumann theorem, Rev. Roum. Math. Pures et Appl. 21 (1976), 97-113. [173] D. Voiculescu, Asymptotically commuting finite rank unitary operators without commuting approximants, Acta Sci. Math. (Szeged) 45 (1983), 429-431. [174] D. Voiculescu, A note on quasidiagonal operators, Topics in Operator Theory, Constantin Apostol Memorial Issue, Operator Theory: Advances and Applications 32, Birkhatiser-Verlag, Basel (1988), 265-274. [175] D. Voiculescu, Property T and approximation of operators, Bull. London Math. Soc. 22 (1990), 25-30. [ 176] D. Voiculescu, Limit laws for random matrices and free products, Invent. Math. 104 (1991), 201-220. [177] D. Voiculescu, A strengthened asymptotic freeness result for random matrices with applications to free entropy, Inter. Math. Res. Notices 1 (1998), 41-63. [178] D. Voiculescu, K.J. Dykema and A. Nica, Free Random Variables. A Noncommutative Probability Approach to Free Products with Applications to Random Matrices, Operator Algebras and Harmonic Analysis on Free Groups, CRM Monograph Series 1, Amer. Math. Soc., Providence, RI (1992). [179] K.W. Wachter, The strong limits of random matrix spectra for sample matrices of independent elements, Ann. Probab. 6 (1978), 1-18. [180] S. Wassermann, A separable quasidiagonal C*-algebra with a nonquasidiagonal quotient by the compact operators, Math. Proc. Cambridge Phil. Soc. 110 (1991), 143-145. [181] S. Wassermann, C*-algebras associated with groups with Kazhdan's property T, Ann. Math. 134 (1991), 423-431. [182] R. Wegmann, The asymptotic eigenvalue-distribution for a certain class of random matrices, J. Math. Anal. Appl. 56 (1976), 113-132. [183] H. Weyl, Der asymptotische Verteilungs gesetz der Eigenwerte linearer partieller Differentialgleichungen, Math. Ann. 71 (1912), 441-479. [ 184] E.P. Wigner, Characteristic vectors of bordered matrices with infinite dimensions, Ann. Math. 62 (1955), 548-564. [185] E.P. Wigner, On the distribution of the roots of certain symmetric matrices, Ann. Math. 67 (1958), 325327. [186] I.M. Wils, Stone-Cech compactifications and representations of operator algebras, Ph.D. thesis, Catholic Univ. of Nijemgin (1968). [187] Y.Q. Yin, Z.D. Bai and P.R. Krishnaiah, On the limit of the largest eigenvalue of the large-dimensional sample covariance matrix, Probab. Theory Related Fields 78 (1988), 509-521.
CHAPTER
9
Applications to Mathematical Finance
Freddy Delbaen Department of Mathematics, E.T.H., Ziirich, Switzerland E-mail: delbaen @math. ethz. ch
Walter Schachermayer Department of Mathematics, Vienna University of Technology, Vienna, Austria E-mail: WSchach @statl, bwl. univie, ac. at
Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Strategies and arbitrage possibilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. The fundamental theorem of asset pricing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. The continuous case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Changes of num6raire and a related Banach space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Weighted norm inequalities and closedness of a space of stochastic integrals . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Abstract We give an introduction to the theory of Mathematical Finance with special emphasis on the applications of Banach space theory. The introductory section presents on an informal and intuitive level, some of the basic ideas of Mathematical Finance, in particular the notions of " N o Arbitrage" and "equivalent martingale measures". In section two we formalize these ideas in a mathematically rigorous way and then develop in the subsequent four sections some of the basic themes. O f course, in this short handbook-contribution we are not able to give a comprehensive overview of the whole field of Mathematical Finance; we only concentrate on those issues where Banach space theory plays an important role.
HANDBOOK OF THE GEOMETRY OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 367
369 374 379 382 384 387 390
This Page Intentionally Left Blank
Applications to mathematical finance
369
1. Introduction The field of Mathematical Finance has undergone a remarkable development since the seminal papers by E Black and M. Scholes [4] and R. Merton [40], in which the famous "Black-Scholes Option Pricing Formula" was derived. In 1997 the Nobel prize in Economics was awarded to R. Merton and M. Scholes for this achievement, thus also honoring the late E Black. The idea of developing a "formula" for the price of an option actually goes back as far as 1900, when L. Bachelier [2] wrote a thesis under the supervision of H. Poincar6 with the title "Th6orie de la spdculation". His aim was precisely to obtain such a formula to be used on the stock market in Paris, which was also booming at the previous fin de si~cle. Bachelier's contribution was remarkable in several respects: firstly he had the innovative idea of using a stochastic process as a model for the price evolution of a stock. He had an almost mystic belief in the "law of probability": Si, dt l'dgard de plusieurs questions traitdes dans cette dtude, j ' a i compard les rdsultats de l'observation ~ ceux de la thdorie, ce n'dtait pas pour vdrif ier des formules dtablies p a r les mdthodes mathdmatiques, mais pour montrer seulement que le marchd, ~ son insu, obdit ~ une loi qui le domine: la loi de la probabilitd. For a stochastic process (St)0~
(1)
Indeed, a moment's reflection reveals that the option is worthless (at time T) if ST <~ K, and is worth the difference between ST and K if ST > K: in the latter case the holder of the option will exercise the option to buy one unit of stock at a price K and can immediately resell it at the present market price ST to make a profit ST - K. (In practice, a cash settlement is often made.)
370
F. Delbaen and W. Schachermayer
Hence we know, conditionally on the random variable St, what the option will be worth at time t = T. But what we really want to know is what the option is worth today, i.e., at time t = 0. To determine this quantity Bachelier simply took the expected value, i.e., Co " - E [CT] -- E [(ST -- K)+]
(2)
which leads to a simple explicit formula for Co involving the normal distribution function. The approach of formula (2) has been used in actuarial mathematics for centuries and is based on a belief in the law of large numbers: if the stochastic model (St)o<
(3)
dSt = # S t dt + a St dBt,
where (Bt)o<~t<~T is a standard Brownian motion starting at B0 --0. Using It6's formula one quickly verifies that the solution to (3) is given by the process
St - So exp
#-
-~
t + o Bt ,
(4)
which is called geometric Brownian motion. If returns are defined through the logarithm, then the price process St gives returns that are stationary and are independent when taken over disjoint time periods. It is a well known result that the only L6vy processes with this property and continuous paths are Brownian motions with drift. This immediately leads to (3) or (4). The rationale of (3) is that the drift as well as the noise term driving the process St is proportional to St; in other words, the difference between Brownian motion with drift, i.e., St = So + at + b Bt and geometric Brownian motion with drift as defined in (4) is similar to the difference between the methodology of calculating linear or compound interest. Clearly, in the short run there is little difference between the linear and the exponential
Applications to mathematical finance
371
point of view (in the deterministic as well as in the stochastic case) while, in the long run, the difference is very noticeable; it also is obvious, that the exponential point of view is economically more meaningful. The model (4) of geometric Brownian motion today became the standard reference model to describe the price evolution of a stock; although promoted by Samuelson, it now is often called the Black-Scholes model or even the Black-Scholes world. It has many very appealing properties but its match to reality is not very good in many aspects (as is very well-known to practitioners): there are several key properties observed in real financial data (e.g., heavy tails, volatility clustering etc.) which are not captured by this model. Many alternative models have been proposed (e.g., in the early work of B. Mandelbrot [38]). Hence, turning back to question (1) raised a b o v e - whether Brownian motion is the good m o d e l - it is generally agreed that geometric Brownian motion with drift is economically more reasonable than Bachelier's original choice (note, however, that in the short run there is little difference), but the question whether (geometric) Brownian motion is a "good model", cannot be answered with a simple yes or no: it depends on the context and purpose of the modeling. We now turn to question (2) pertaining to the economic soundness of the pricing argument based on the law of large numbers: it was the merit of Black and Scholes [4] and Merton [40] to have replaced this argument by a "no arbitrage" argument, which is of central importance to the entire theory. The basic principle underlying this idea is nicely explained in the subsequent little story quoted from a paper of Dupire [20]: Imagine you are offered a strange bet. A coin is tossed; you win $60 if it comes up heads and lose $40 f o r mils. Would you accept the bet ? It looks attractive: the expectation is (0.50 x 60) + (0.50 x - 4 0 ) = 10, a healthy bias. A more thorough examination may lead to second thoughts, however. Perhaps the pain o f losing $40 would overweight the pleasure o f making $60. What if it were $60 000 and $40 000? What if the coin were unbalanced, or tossed by a manipulator who can get tails three times out o f four? In other words, your decision would depend on both preferences (risk aversion) and expectations, and it may well be wiser to refuse this apparently attractive proposition. Let us now set the stage f o r a very similar situation. You are in a casino, at a simplified roulette table. There are only two possible outcomes: red or black. There is no zero and it is possible to play with the casino on a 1:1 basis. The proposed bet is: win $60 if red comes up and lose $40 f o r black. Would you accept ? At first sight, this seems equivalent to the previous bet, so you would decline the invitation. But the fact that it is possible to play with the casino makes a real difference. You can hedge the bet and build up a certain gain, regardless o f the outcome - an arbitrage, in other words. Just accept the bet and, at the same time, put a stake o f $50 on black with the casino. I f red does come up, you make $60 through the bet and lose $50 with the casino. If black wins, you lose $40 through the bet and make $50 with the casino.
372
F. Delbaen and W. Schachermayer
In either case, you are better off by $10: the expected, potential gain has been converted into a certain one. Most important, this is true regardless o f the real probabilities o f red versus black, which could very well be 10: 90.
Summing up: an arbitrage- which is the concept crystallized in the second part of the above story - is a riskless way of making a profit with zero net investment. This intuitive notion will be mathematically formalized in Section 2 below. An economically very reasonable assumption on a financial market consists of requiring that there are no arbitrage opportunities. (To convince yourself that this assumption is indeed reasonable, just think for a moment, what would happen in a liquid financial market offering arbitrage opportunities.) The remarkable fact is that this simple and primitive "principle of no arbitrage" allows already to determine a unique option price in the BlackScholes model: to be slightly more formal, we assume that in our financial market there are two traded assets, the stock S = (St)0<~t<~T whose price process is given by (4) above and a riskless bond B whose price process (Bt)o<~t <~T evolves deterministically with a constant interest rate r, i.e., Bt -- Boe rt . For (mainly notational) simplicity we assume r = 0. We denote by (ft)O<~t<~T the (right continuous, saturated) filtration generated by the process S (which is defined on some stochastic base (1-2, ,T, I?)). We then have that every 9t'r-measurable random variable f (satisfying an appropriate integrability condition) can uniquely be written as
f -- c +
f0 T H u d S u - - c + ( H . S ) T
(5)
for some constant c and a predictable process (Ht)o
Applications to mathematical finance
373
Ht to rewrite this as an integral on S martingale. It is trivial, using (3) and letting Ht -- ~-g, (remember that # = 0):
f--
/0 T-H, dB, --
/~, cr dS, S. -
H. dS..
(6)
Hence, admitting the above cited martingale representation theorem, we get the representation (5) for an arbitrary f E L2(X2, f , 1?) where c = E l f ] . Hence we essentially find the same methodology as used by Bachelier and by actuaries: calculate prices by taking expectations. The reasoning behind this procedure however is now completely different; the argument involving the strong law of large numbers has been replaced by the (economically much more compelling) no arbitrage argument. What happens in the case # -76-0? Here the crucial trick is to pass from the original measure 17 to an equivalent measure Q such that under Q the process S becomes a martingale. The determination of the Radon-Nikodym density of the measure Q with respect to the original measure IP is the theme of the Cameron-Martin-Girsanov theorem (see, e.g., [43]): under the above assumptions the measure Q << 17 under which S is a martingale is unique, and its density is given by dQ ( # 1#2 ) d1? - e x p ---Brcr - 2 ~ - j T .
(7)
dQ Note that i--Y > 0 1?-a.s. so that the measures Q and 1? are equivalent, i.e., have the same nullsets. Now apply the above reasoning to the process S defined on (s 5c, (ft)0~
c = EQ[f],
(8)
where E~[. ] denotes expectation taken with respect to Q. Using again the principle of no arbitrage we find that c is the only possible price for f not allowing arbitrage opportunities (in the mathematical formalization of the no arbitrage principle given below we shall see that this principle remains unchanged under an equivalent change of measure; the only thing that matters are the sets of measure zero). For a European call option f = (ST - K)+ the constant c in (8) can be explicitly calculated yielding the celebrated Black-Scholes option pricing formula. Summing up the above methodology: we have passed from the original measure 17 to a so-called risk-neutral measure Q, which in the Black-Scholes model turned out to be unique, and then have used this measure to calculate the price of a derivative f via (8); we have claimed that this is the unique arbitrage-free price for f . The above presentation was deliberately informal and should only motivate the questions which are dealt with in full mathematical rigor in the subsequent sections. For example: What happens for more general stochastic processes (St)0~
374
F. Delbaen and W. Schachermayer
The latter question is the theme of the so-called Fundamental Theorem of Asset Pricing w h i c h - roughly speaking- states that a process S = (St)0~
2. Strategies and arbitrage possibilities As mentioned in the introduction, agents can use investment strategies, that are time dependent. This leads to the use of stochastic integrals. Of course this requires then some regularity on the trajectories of the stochastic processes used to model prices. Also as we will see, not all investment strategies can be used. We will first give some definitions and then we will give the interpretations. As stated in the introduction the main ingredient in arbitrage theory is stochastic integration. In this section we will give a mathematical formulation of the objects referred to in the introduction. All processes are defined on a stochastic
Applications to mathematical finance
375
basis (s ( f t ) t >~0, P) where f t describes an increasing family of o--algebras on I2. The probability measure l? is defined on the o--algebra f e c = V t ~>0f t . o-algebra f t can be seen as the mathematical description of the information available at time t. As common in stochastics, we will also assume that the filtration ~ satisfies the usual hypotheses, i.e., the filtration is right continuous, meaning that f t = ['-'ls>t f s , and the o--algebra f 0 contains all the null sets of foc. To avoid trivialities at time t = 0 we assume that f 0 only consists of the null sets and their complements, set, the set of natural numbers or even a finite set. By taking R+ as the time set we aim for the most general situation. It is obvious that models with finite time horizon can be imbedded in a model where the time set is R+. The time an agent wants to buy or sell a financial instrument can be a random variable, prices coincide. Of course such random times can but it may only depend on past information and hence it must be a stopping time. DEFINITION 2.1. A stopping time T is a random variable T :s -+ R+ U {+co} such that for each t ~> 0, we have that {T ~< t} E Ut. If T1 ~< T2 are two stopping times then the stochastic interval ~T1, T2~ is the set
~T1,T2~- {(t, co) It
9 R+. T~ (co) ~
(9)
Other intervals ~T1, T2~, ~T,, T2[[, ~T1, T2~ are defined in an analogous way. The smallest o--algebra on R+ x s containing the stochastic intervals of the form ~T1, T2]] ( T1 and T2 are stopping times) is called the predictable o--algebra 72. A process H :R+ • s --> R d is called predictable if it is measurable with respect to the predictable o--algebra. Mathematically a process S, modeling the price evolution of d stocks, is best described by a d-dimensional semi-martingale S : R + x I2 --+ Rd. The set R+ is the set of times where trading can take place. As in the introduction we also suppose that there is a riskless bond (Bt)0~
F. Delbaen and W. Schachermayer
376
If S is a d-dimensional c?adl~g process then we can define the stochastic integral in an elementary way: f
( H . S)~ = ] Hu dSu J[o,oo] d
(10)
n
- ~Zfj(
Tj+! _
_
T;) .
S i
S i
(11)
i--1 j----1
The indefinite integral defines a cadl?ag process. f
( H . S)t -- I
H, dSu
(12)
,t] _
--Z
S i fj(T;+,At
_
S TjAt)" i
(13)
i=1 j = l
The notion of stochastic integral has an immediate financial interpretation. If the trader decides to hold the position (fj) during the time period ~Tj, Tj+l]] then the value of the portfolio is easily seen to be described by the process ( H . S). The final value of the process ( H . S ) ~ is then given by ( H . S ) ~ = limt__,~(H. S)t = (H. S)Tn+~. Stochastic integration is a natural tool in financial modeling. Predictable processes serve as strategies. DEFINITION 2.3. A c?~dl?ag adapted process S ' R + • ~2 --+ R d is a semi-martingale if for every T < cx~ the mapping | --+ L0(S2, U ~ , P); H --+ ( H 9S)T is continuous for the sup-norm topology on G ~ . This means that if suPt,~o IHtn (co) l --+ 0 then ( H n 9S)T ~ 0 in probability. If S is a semi-martingale then the notion of stochastic integral ( H . S) can be extended to all bounded predictable integrals H. This is the subject of the general stochastic integration theory, see, e.g., [ 17,32,42]. Exactly as in the deterministic, one-dimensional case, the notion of integral can be extended to some unbounded integrands H. See [32,42,45] for details. If ( H . S) can be defined, then we say that H is S-integrable. Integrands that are S-integrable can be seen as general strategies. If the limit ( H 9S ) ~ - l i m t ~ ( H 9S)t exists, then we can say that ( H 9S ) ~ is the final outcome of the strategy H. As observed in [29] not all S-integrable processes H can be accepted as reasonable strategies (from the financial point of view). Indeed we should avoid so-called doubling strategies. The most classical example of such a doubling strategy can be constructed using Rademacher functions. EXAMPLE 2.4 (Doubling strategy (discrete time)). On (S-2, U ~ , P) we take (gn)n>/1 t o be a sequence of independent variables, so that P[en - 4-1] - 1 which we may think of as the gain or loss of a risky investment revealed at time n. As filtration we take (Un)n>>O where Un = a (el . . . . . en). Sn is defined as el -+-...-+- en. The reader can easily change the
Applications to mathematical finance
377
definition in such a way that (.Tt)0~
Lt -- exp(Wt - 89 is easily seen to be the unique solution of the stochastic differential equation dLt = Lt dWt, with initial condition L0 -- 1. Hence we find that
1 -
Lt
f]0,t]
( - L u ) dW, = (H . W)t.
-
-
In both the discrete and the continuous example, we get that ( H 9S) is a martingale but not a uniformly integrable martingale. To exclude the p h e n o m e n o n described by the above examples (where in both cases it was crucial that the process ( H 9S)t was not uniformly b o u n d e d from below) one m a y adopt the following concept: DEFINITION 2.6 ([30] and [11]). An S-integrable predictable process H : R + x $2 --+ R d is called a-admissible for some a ~> 0 if, (i) ( H 9S) ~> - a i.e., the process always remains above the ("budget") level - a , (ii) l i m t ~ ( H S)t = ( H . S ) ~ exists. We call H admissible if it is admissible for some a ~> 0. It is clear that the set of a-admissible integrands forms a cone but not necessarily a vector space. We define: K = { ( H . S)oc I H admissible}, c
-
(14)
h L+InL --(X-L+)nL
interpretation, it is simply the cone of all outcomes of admissible strategies (starting with an initial wealth zero). The set C has a similar interpretation. It is a philanthropic version of K in the sense that after having obtained ( H . S)oc the investor can give away a nonnegative amount of m o n e y thus making the gain uniformly bounded. Before continuing, let us point out that in the case of non-locally b o u n d e d processes it m a y happen that K is reduced to {0}. Indeed, for an easy example take St = 0 for t < 1 and St - X for t ~> q 1 where X is a standard normal variable. The filtration (.Tt)0~
378
F. Delbaen and W. Schachermayer
predictable integrand H we have that H1 -- constant. It follows that ( H . S ) ~ = ( H . S)l is of the form )~X where )~ ~/1~ and H is only admissible if H1 = 0 which yields ( H . S ) ~ = 0. In this example we obtain that C - - L + . We now have enough material to define the No Arbitrage properties. Roughly speaking these properties say that it is impossible to make money out of nothing (no money pump, no arbitrage, no free lunch ...). Because of doubling strategies we have to restrict to admissible integrands. DEFINITION 2.7 ([11]). The semi-martingale S satisfies the No Arbitrage (NA) property if K A L + = {0} or, equivalently, C fq L + -- {0}. The idea of this notion is that, by using admissible strategies, the investor can only make money if he/she is willing to face the possibility of also losing money. A standard and very intuitive example of (NA) is given by: PROPOSITION 2.8. If there is aprobability measure Q "- IP underwhich S is a martingale, then S satisfies (NA). PROOF. Let H be admissible and Q a probability measure, Q ~ P; then ( H . S) is a Qlocal martingale by Emery's theorem (see [25]). Since ( H . S) is bounded below it is a Q-super martingale and hence E Q [ ( H 9S)e~] ~ 0. It follows from the assumption Q "~ P that P [ ( H . S ) ~ < 0] > 0 as soon as P [ ( H - S ) ~ > 0] > 0 i.e., (NA). [] The fundamental theorem of asset pricing is the appropriate converse of Proposition 2.8. It is easily seen that in the case of a finite set #2, the cone K is a vector space and an easy separation argument yields the existence of a risk neutral probability Q such that S becomes a martingale (compare the more general arguments below). However in the general case, it is not clear that the sets K or C are closed (in an appropriate topology). Also in a general context it can be shown that it is impossible to obtain a martingale measure Q for a process S. This is of course related to the integrability properties of the process S. The generalizations we need are described in the following definition. DEFINITION 2.9. If (S-2, (~t)O~t,~) is a filtered probability space, then a process S :IR+ x #2 ~ IR (or IRd) is called a local martingale if there is a non-decreasing sequence of stopping times 0 ~< To <~ Tl ~< T2 <~ ... such that Tn --+ ~ and for each n we have that the stopped process S Tn is a uniformly integrable martingale. We recall that S T is defined as the process S f = StAT. In case the time interval is finite, e.g. [0, 1], we require that limP[Tn = 1] = 1. A standard example of a uniformly integrable (!) process S that is a local martingale and not a martingale is given by the inverse of a Bessel process in 3 dimensions (see, e.g., [43]). The process is defined as follows. We start with a 3-dimensional Brownian motion X : [ 0 , 1] x S2 --+ R 3. The filtration is the standard filtration generated by X. We suppose that X0 = (1, 0, 0). The Bessel process R is defined as R = IlSll where II II is the Euclidean norm on IR3. By the well known property of Brownian motion in 3 dimensions
Applications to mathematical finance
379
we have that R -r 0 and S -- 1/R is therefore well defined. It can be checked that S is a local martingale (roughly speaking because 1/ll R II is harmonic on IR3\{(0, 0, 0)}), that (St)o<<.t<<.1is a uniformly integrable set (in fact bounded in any L p space with p < 3) and that E [St ] < 1. The latter shows that S cannot be a martingale. If we replace the filtration by the smaller filtration generated by S, then it can be shown that IF' is the only probability that turns S into a local martingale. The reader can check that Proposition 2.8 also holds for local martingales and hence we get that S satisfies (NA) (see [43] for details on Bessel process). For non-continuous processes the concept of local martingale is not yet sufficient as is shown in [15]. We need the even more general concept of sigma-martingale. These processes were introduced by Chou and Emery (see [25]). DEFINITION 2.10 ([15]). A process S ' R + x S-2 ~ IR defined on the filtered probability space (S-2, (~t)0~
3. The fundamental theorem of asset pricing We start this section with a generalization of the concept of No Arbitrage, which can be seen as a topological version of (NA). DEFINITION 3.1. With the notation of Section 2 we say that the semi-martingale S'IR+ x S2 --+ R d satisfies the No Free Lunch with Vanishing Risk (NFLVR) property if n L + -- {0} where d is the closure of C in Loc with respect to the norm II IIoo. There are different topological versions of (NA) pertaining to weaker topologies than the sup-norm (and therefore yielding stronger assumptions). We refer the reader to [ 11 ] for a discussion of the different concepts. THEOREM 3.2. A semi-martingale S satisfies (NFLVR) if and only if S satisfies the following two properties: (a) S satisfies (NA). (b) If en > 0 is a sequence of positive numbers tending to zero, if (Hn)n>>.l is a sequence of en-admissible strategies (i.e., H n 9S ~ - e n ) then (H n 9S)oc --+ 0 in probability.
380
F. Delbaen and W. Schachermayer
One may give examples showing that (a) does not imply (b) and (b) does not imply (a). For the former we defer the discussion to Section 4 below. For the latter we may take the Bessel process R in three dimensions (see [ 12]). We drop the details. The reader can now see why the property is called No Free Lunch with Vanishing Risk. The No Free Lunch with Vanishing Risk property is also equivalent to a boundedness property. In fact we have: THEOREM 3.3. A semi-martingale S satisfies (NFLVR) if and only if S satisfies the following two properties: (a) S satisfies (NA). (b) The set K = {(H-S)oc I n is 1-admissible} is bounded in probability. The boundedness property also has an immediate economic interpretation. If strategies are chosen so that the losses are bounded uniformly (i.e., in Lot), then the gains are bounded in probability (i.e., in L0). This can also be seen as some continuity property. Of course, the relation between items (b) of the two preceding theorems is immediately noticeable. In some cases the analytic definition (i.e., through a stochastic differential equation) may help. Property (b) can easily be checked in the case of continuous processes. We will see this later on. The fundamental theorem of asset pricing in its most general form can now be stated (see [11] and [15]). THEOREM 3.4. For a semi-martingale S :IR+ x s --+ IRd the following two properties are equivalent: (a) S satisfies the (NFLVR) property. (b) There is an equivalent probability measure Q ~ P such that under Q the process S is a sigma-martingale. If we assume that the semi-martingale S is locally bounded (respectively bounded) the term "sigma-martingale" in (b) may be replaced by the term "local martingale" (respectively "martingale "). In case the (NFLVR) property is satisfied we denote by l~/I[e the set of all equivalent sigma-martingale measures. The set M is the set of all absolutely continuous sigmamartingale measures. In case S is locally bounded NI -- M e, where the closure is taken with respect to the norm of L1 (I?), but with general case NI ~ M e. The proof of this theorem is lengthy and we cannot repeat it here. But we will give a sketch of the different points that relate the theorem to functional analysis. The first and most delicate step consists in proving that under the assumption of (NFLVR) the set C is already weak*-closed. Yan's theorem [52] (whose proof consists of a combination of a Hahn-Banach and an exhaustion argument) (see, e.g., [48]) then yields the existence of an equivalent probability measure Q such that, for all f 6 C, E q [ f ] ~< 0. Interpreting Q as a linear functional on L ~ , it separates C from the positive cone L + . In the case where S is locally bounded this already implies that S is a local martingale for the measure Q. In the general case one has to make an additional (nontrivial) step
381
Applications to mathematicalfinance
that consists in showing that, for e > 0, there is a measure Q0 such that Q0 is equivalent to P, that ]IQ - QoIJ ~< e and that under Q0 the process S is a sigma-martingale. See [ 15] for details and for an example that shows that one cannot do better than a sigmamartingale. As indicated above, the central point of the proof is the fact that C is weak*-closed. This is done using the Krein-Smulian theorem, also called the Banach-Diedonn6 theorem. This theorem says that a convex set C in the dual X* of a Banach space X is ~ (X*, X) (i.e., weak*-closed) if and only if C n (nBx,) is ~(X*, X) closed for each n ~> 1. If X = L1 and X* = L ~ we can, using the characterization of relatively weakly compact sets in L1 as uniformly integrable subsets of L1, make this even more precise. A convex set C C L ~ is weak*-closed if and only if, for each sequence ( f , ) ~ > l in C that is uniformly bounded and converges in probability to a function f , we have that f E C. Since in our context the set C is a cone we have to show the following fact. CLAIM 3.5. Let (Hn)n>~j be a sequence of 1-admissible integrands, let (fn)n>~l be a sequence in L ~ ~ , ~) such that - 1 <~ fn <<.(H n" S)~, which tends to f in probability; then under the assumption of (NFLVR) there is a 1-admissible integrand H such that f <~ ( H . S)~. To prove this claim we have to replace the integrands by a better choice. Let us define the following concept: _
_
m
DEFINITION 3.6. An element f E C1 is called maximal if g E C and g >~ f a.s. implies g - f . Here C1 denotes the set C1 - {h ] h 6 Co, h ~> - 1 } and the bar refers to the closure in the space L0 (i.e., with respect to convergence in probability). m
Our assumptions on (NFLVR) show that C1 is bounded in L0 and hence C1 has sufficiently many maximal elements. In fact every g E C1 is dominated by such a maximal element. To prove the claim we reduce it to the following" CLAIM 3.7. The maximal elements o f C1 are already in C1. So let us fix a sequence ( H n ) n ~ l such that H n is 1-admissible and such that ( H n 9 S ) ~ --+ f in probability, where f is a maximal element in C1. The first trick is to replace the probability P by an equivalent probability (still denoted by I?) such that the function supn~> 1 supt I(H n 9S)t] is in L2(]P). That this is possible follows from an upcrossing type lemma and the maximality of f . After an additional technical reduction, which we skip, the D o o b - M e y e r decomposition theorem now allows to decompose S into a martingale M and a predictable process of finite variation A, i.e., S - M + A. The D o o b Meyer decompositions of ( H n 9S) are then given by ( H " 9M) and ( H ~ 9A). In order to control the jumps of M and A we use the following generalization of an inequality of Stein.
382
F. Delbaen and W. Schachermayer
PROPOSITION 3.8. Let (.Tn)n=0 ..... N be a discrete time filtration on the probability space (12, ~ , I?), let ( fn)n--1 ..... U be adapted to (.Tn)n=O ..... N, let gn be the predictable projection, i.e., gn = E [fn I f ' n - l ] then we have
EI(ZJgn[q)P/q]I/P~2E[(~]fnlq)p/q]1/p whenever 1 ~ p <. q <~ cx~. In other words the mapping
L p (l N)
> L p (lqN),
(16)
(fn)n>~l t
> (E[fn I.T'n-1])n>~l
(17)
has norm less than or equal to 2.
The next step is to show that the convex hull of ( ( H n. M)~)n>>l and of ( ( H n. A)~)n>>l are both bounded (in a good sense). Since ( ( H n . S ) ~ ) n ~ l is bounded in L2(I?) both sets are at the same time bounded or unbounded. But if they are unbounded we are faced with the fact that ( H n 9A)t increases in a "linear" way whereas ( H n . M ) t , due to the orthogonality of its increments, grows only in a way related to the square root. This leads to a contradiction. Once the boundedness is proved, it is a straightforward track to find convex combinations K n of (Hn)n>~l that converge to a predictable process K such that (K 9S ) ~ = f . The latter technique is a combination of techniques of Memin (see [39]). The separation argument in the proof of the fundamental theorem can be exploited further. It yields to the following duality result (see [24,9]). THEOREM 3.9. A s s u m e that S is a locally bounded semi-martingale such that M e is not empty. Then f o r f ~ 0 we get sup E Q [ f ] : inf{ot I 3g ~ K with ot 4- g ~ f } : : or0.
(18)
Furthermore, if the quantities are finite, the infimum is a minimum and there is a maximal element g E K max with oto 4- g ~ f .
4. The continuous case
When the process S is continuous or, more generally, locally bounded, the assumption on S to be a semi-martingale turns out to be a necessary condition for the conclusion of the fundamental theorem of asset pricing to hold true. In fact we can prove the following result (see [ 11 ]), stated under a finite time horizon, say [0, 1]. THEOREM 4.1. I f the locally bounded process S = (St)o<~t <<1. is not a semi-martingale then there is a function f >~ 0, I?[f > 0] > 0, and a sequence o f simple strategies Hn such that (Hn . S) >>.- e n , en --+ O, (Hn . S) 1 ~
f .
Applications to mathematicalfinance
383
We recall that a strategy H is simple, if there is a finite sequence of stopping times
0 <, To <~T1 <.... <, Tn < cx~ as well as functions (gk)k>~0, gk being 9t'r~ measurable and H __ ~n--1 k=0 gk 2~rk,Zk+~" These strategies are most elementary and the definition of their
stochastic integral (H 9S) does not involve any calculus. The theorem is not true for processes with unbounded jumps, since in this case the boundedness condition on (H 9S) might imply that H ~ 0. Theorem 4.1 has some nice consequences for continuous processes. In particular fractional Brownian motion fails to be a semi-martingale (with the exception of Brownian motion, of course) and therefore allows, by Theorem 4.1, this kind of arbitrage. This implies that these models are useless in the context of pricing and hedging derivative securities by no-arbitrage arguments. This fact was known since a long time, but even after publication of the above theorem, people still continued to ask about fractional Brownian motion. Rogers [44] decided to settle the problem by giving a fairly explicit strategy H such that (H.S) l ~>0and~[(H.S)I>0]>0,(H.S)>/-1. The strategy is not simple (one can show that for simple strategies (NA) is satisfied), but it is "semi"-simple in the sense that restricted to each interval [0, t] with t < 1, it is simple. Fractional Brownian motion, however, has some nice long range dependencies, observed in financial data. In forthcoming work of Cheridito [5] the reader can find a way out to this dilemma. For continuous processes the concept of (NA) can be analyzed further. So let us assume that S is a continuous semi-martingale with respect to a filtered space (I2, (~t)0~
~u f olt h 2 d(M, M)u), the so-called Dol6ans-Dade or stochastic exponential of ( - h 9M),
384
E Delbaen and W. Schachermayer
also written as s 9M). Of course the existence depends on the finiteness of the integral f0 h] d(M, M)u. If the integral is not almost surely finite then there are essentially two possibilities. To distinguish them we introduce the stopping time
{If0 ' h u2 d(M , M)u -- e~ }.
T--inf t
(19)
On the set T ~ 1 we then have either f o h ] d ( M , M)u --cxz or we have, Ve > 0, fTT+ e hu2 d ( M , M)u = oc.
Technically the latter case is more difficult but it has a simple solution: there is a very special kind of arbitrage, namely there is H supported by ~T, 1~ and such that (H 9S) ~> 0 i.e., the outcome at every time is nonnegative, and iP[(H 9S)T+e > 0] > 0 for all e > 0. dt + d Wt where W is Brownian motion An example of this nature is given by dSt - -~
defined in its natural filtered space. A strategy of the form above is given by Kt -1/(~/71og(t-1)) and then we stop the process ( K . S) either when it reaches a level 1 or when it hits 0 for the first time after 0. The iterated logarithm theorem implies that, immediately after 0 the process (K 9S) is strictly positive! The next result is closely related to these arguments. It was shown in [12] and, independently and under slightly stronger hypothesis, in [37]. THEOREM 4.2. I f the continuous martingale S satisfies (NA) then there is a measure Q absolutely continuous with respect to iP and such that S is a local martingale under Q. The support o f the measure Q can be chosen to be equal to {L1 > 0} where Lt -- e x p ( hu dMu - I h2 d ( M , M)u) stopped when h2 d ( M , M)u hits oc, or what is the same: when L hits O.
fo
fo
fo
Even when S satisfies (NA) and fo h] d(M, M}u < ec a.s., the process L need not be a martingale, i.e., it can happen that E[L1] < 1. This means that the measure Q is not necessarily given by dS = L1 diP. See [47] and [ 16]. We also have the following: PROPOSITION 4.3. I f S is a continuous semi-martingale decomposed as dS = d M + h d ( M , M ) then the set KI = {(H. S ) ~ ] H is 1-admissible} is bounded in Lo(IP) if and only if f 0 h2, d ( M , M)u < cx) a.s.
The 3-dimensional Bessel process shows that this does not rule out arbitrage (compare
[13])~ 5. Changes of num~raire and a related Banach space Throughout this section we suppose that the process S modeling the price of d stocks is a d-dimensional semi-martingale that is locally bounded. We also suppose that the process
Applications to mathematical finance
385
S admits a local martingale measure. We repeat that the set of equivalent local martingale measures is denoted by M e, the notation M being reserved for the closed convex set (in fact the closure of M e) of absolutely continuous local martingale measures for S. Ka will denote the cone of outcomes of (a-)admissible strategies, K - - U a Ka. As seen before, maximal elements play an important role in the proof of the fundamental theorem. However, more can be said. If V = 1 + (H 9S) is a strictly positive process satisfying V~ > 0 a.s., then we might use V as a new money unit (say $ instead of CHF or Euro). The market 1 s ) . Economic interpretation is now described by the (d + 1)-dimensional process X - (V, leads us to conjecture that S satisfies (NA) if and only if X satisfies(NA): It does not matter whether you do the bookkeeping in $, CHF or in Euro! However since the admissible strategies may change, we have to be more careful. THEOREM 5.1. If S satisfies (NFLVR) then the following assertions are equivalent for a process V = 1 + (H . S), where H is admissible and V ~ > 0 a.s.: 1 s (1) X -- (V, V ) satisfies (NA), (2) there is an equivalent local martingale Q measure f o r the process S, such that V is a uniformly integrable Q-martingale, i.e., E q [ V ~ ] = 1 or equivalently E q [ ( H 9 S)~] =0, (3) (H . S ) ~ is maximal in K1 (and hence also in K). It is not true that in this case V is a uniformly integrable martingale for each element R in the set of equivalent local martingale measures M e, see [47] or [16] for this surprising fact. The above theorem also yields another proof of the theorems of Jacka [31 ] and AnselStricker [ 1]. THEOREM 5.2. If S satisfies (NFLVR) and f is a positive random variable, then the following conditions are equivalent: (a) f = ~ + (H . S ) ~ with (H . S ) ~ maximal in K, (b) there is Q E 1 ~ e such that E q [ f ] = sup{E R[f][R E M} < cx~. The Bishop-Phelps theorem now immediately implies THEOREM 5.3. If S satisfies (NFLVR) and continuous local martingale measures are already equivalent, then M is reduced to a singleton. PROOF. As the set of absolutely continuous local martingale measures M is a bounded, closed and convex subset of L1, the set { f I f attains its supremum on M} is a norm dense subset of L ~ . The set of all elements of the form c~ + ( H - S ) ~ , where the process ( H . S) is bounded is therefore dense in L ~ and because it is closed (this follows essentially from claim 3.5) we have that it equals L ~ . This implies that all elements of L ~ are constant on M, hence M can only be a one-point-set. [-1 Another application of Banach space theory is given by James' theorem on weakly compact sets. We state the result in its negative form, see [9] for details.
386
F. Delbaen and W. Schachermayer
THEOREM 5.4. Suppose S is continuous, satisfies (NFLVR) and suppose that all martingales with respect to (.Yt)o<<.t are continuous (i.e., all stopping times are predictable). Then we have (a) either M is a singleton, (b) or M is so big that it has no extreme points. It turns out that (a) occurs if and only if 1Vgis weakly compact and the proof uses James' theorem. In the case when S is only assumed to be locally bounded (and not necessarily continuous), the above theorem is false and the implications of M being weakly compact are not yet fully understood. Turning back to the theme of Theorem 5.1 above, we find that the maximal elements generate, in a natural way, a Banach space. THEOREM 5.5. The set K max of maximal elements forms a cone in L1 (~). This is not obvious. Indeed if g, f 6 K, E Q [ f ] = 0 and E R[g] = 0 for Q, R 6 M e then how can we find an element Q' ~ 1~]ie with E Q , [ f + g] = 0? The proof uses the numeraire Theorem 5.1 above. With the convex cone K max we may construct the vector space G = K max - K max. On G there is natural norm, if g = f - h where f, h ~ K max then we put Ilgll--inf{a I g - f - h, f, h E gamaX}. Surprisingly (G,
(20)
II II) is complete and
211gll = sup{llgllL,(Q) I Q ~ l~/i[e}.
(21)
Although the constructions of K max and of G make sense economically, the implications of this result in mathematical finance are not yet clear. Examples in [14] show that the space G can have different natures, it can be an L1 space but it can also contain a complemented L ~ space. Also the space G is different from the space of functions f so that sup{ IIf IIL I(Q) I Q ~ Me } < 00. Indeed take f 6 K1 \ K max (such elements exist) then f ~ G (this requires a proof!) and supQ~IVl[e IIflIL~(Q) ~< 1. A side result of the theory is the following THEOREM 5.6. I f f E K max then { Q I Q 619I[, E Q [ f ] - 0 } is a dense G~-set in M. Hence { Q I Q ~ Me, E Q [ f ] = 0} is a dense G~-set in M e.
(22)
387
Applications to mathematical finance
6. Weighted norm inequalities and closedness of a space of stochastic integrals In this section we investigate a topic initiated by the work of F611mer and Sondermann [27] and F611mer and Schweizer [26], to deal with the problem of hedging in incomplete markets, i.e., when there is no uniqueness of the equivalent martingale measure. To motivate the idea, first suppose for simplicity that S is an R-valued martingale under the original measure I? which we also assume to be bounded in L 2 ( ~ , .)c-, iP). We denote by G the subspace of L 2 ( ~ , f', iP) consisting of the functions of the form
f --
fo T Hu dSu
(23)
= ( H . S)T,
(24)
where H is a predictable process such that the stochastic integral makes sense in L 2, i.e., such that ((H- S)t)t~
II/0THu dS,
L2 (~,.)c',IP)
(25)
= IIHIIL2(S-2•
=
Ht2 (co) d(S)t (09) diP
.
(26)
Here 79 denotes the sigma-algebra of predictable subsets of s x [0, T] and diP d(S) denotes the finite measure on 79 induced by the quadratic variation process of S. As the space L2(s x [0, T], 79, diP d(S)) is obviously complete and can be identified via the above isometry with thesubspace G of L2(~2, .T, iP) we find that G is closed in L2(I2, .T, iP). We denote by G the space spanned by G and the constant functions which clearly again is closed in L 2. Note in passing that in the case when (St)0~
(27)
where c is a constant, g may be written as g - ( H . S)T while h is orthogonal to G. We may associate to g and h the martingales E[g I~'t]O<~t<~T= ((H. S)t)O<~t<~rand (E[h I bet ])0~t ~
388
E Delbaen and W. Schachermayer
So far we have used the simplifying assumption that the price process S - (St)0~
(28)
We have seen in Section 4 that under the condition of (NFLVR) the IRd-valued measure dA is absolutely continuous with respect to d(M) (taking values in the non-negative definite operators on 1Rd) and we may find an IRd-valued predictable process )~ = ()vt)0~
dAt -- d(M)t)~t
1?-a.s. for 0 ~< t ~< T.
(29)
A formal application of the Girsanov-theorem indicates that the probability measure Qmin defined via the density process
Lt := s dQ min d1?
9M)t,
and
(30)
:= LT" = C(-)v. M)7",
(31)
where C denotes the Dol6ans-Dade exponential of a local martingale, is a candidate for an equivalent local martingale measure for the process S (provided all the involved limiting procedures make good sense and L7" really defines a density of an equivalent measure). F611mer and Schweizer have called this measure the "minimal" martingale measure Qmin (if it exists). Another natural choice from the set l~/I[e (S) of equivalent local martingale measures for S is QOpt, the "variance-optimal" measure, which minimizes the L 2 (S2, bt-, 1?)-norm among all elements of ~/]~e(S) (again, provided it exists). In many cases, e.g., when S has independent increments, Qmin and QOpt coincide, but in general they are different. We now turn again to the theme of closedness of a (properly defined) space of stochastic integrals. To introduce the notion of predictable trading strategies appropriate in the present context, we call a predictable IRd-valued process L2-admissible if H . M as well as H . A make sense in L 2, i.e., if H . M is an L 2 (S-2, ~ , 1?)-bounded martingale and E [ IHu dAu I)2] < oo. The definition is chosen in such a way that we may define the stochastic integral H . S - H . M + H . A and that ( H . S)7~ is in L2(12, ~ , 17). Denote again by G the subspace of L 2 (12, ~ , 17) formed by the functions ( H . S)7", where H runs through the L2-admissible trading strategies.
(fo
Applications to mathematical finance
389
To formulate the theorem we recall that a measure Q .~ I? with density process Lt -E [ ...~ I fi't_], satisfies the reverse HOlder condition Rp, for 1 < p ~< oc, if there is a constant C such that E
~
I.Tt
1
~CE
~l.Tt
]'
forl~t~
(32)
The name stems from the fact that by HOlder's inequality the reverse inequality is always satisfied with C = 1. The reverse H61der condition is dual to the Muckenhaupt condition (see [ 18]). We now can formulate the central result of [ 10]: THEOREM 6.1. Let S = (St)o<<.t<.T be a continuous semi-martingale and suppose that there exists an equivalentprobability measure Q with ~dQ E L2(I2, f ' , 17) under which S is a local martingale. The space G is closed in L 2 (I-2, .T, 1?) if and only if there is an equivalent local martingale measure Q for S which satisfies the reverse HOMer condition R2. In this case the so-called "variance-optimal" measure Qopt, which has minimal L 2 (s .T, 1?)-norm among all equivalent local martingale measures, is well-defined and satisfies the reverse HOlder condition R2. There are some variants of the above theorem in terms of the process )~. M. A necessary condition for Theorem 6.1 to hold true (under the assumptions given there) is that the process )~. M is in BMO but this condition is not sufficient. In fact )~. M is in BMO is equivalent to the completeness of G with respect to the stronger norm defined via the maximal function II(H. S)ll "--Ilsup(H. M), IlcZ(s~,~-,?)-
(33)
t dQ min
On the other hand the condition that the minimal m e a s u r e ~min defined via o~ = s M) exists and satisfies the reverse H61der condition R2 is sufficient but not necessary for the theorem to hold true. In fact, this stronger condition is equivalent to the fact that G is closed and, in addition, that the (in general not orthonormal) projection re from L 2 ( ~ , f ' , P) onto G with ker(zr) -- M • is well defined and continuous. Here M • denotes the subspace of L 2 orthogonal to the space M generated by the constants and the stochastic integrals on the local martingale part M of the process S. If Jr is well-defined and bounded then Jr induces the F611mer-Schweizer decomposition as it splits a general contingent claim f E L2($2, f , 17) into the hedgeable part 7r(f) which may be written as Jr ( f ) - const + f o 14, dSu for an L 2 -admissible trading strategy H and the non-hedgeable part f - Jr ( f ) which induces a martingale strongly orthogonal to the local martingale part MofS. We refer to [10] for a detailed presentation of these topics for the above discussed closedness of G in LP for p = 2 and the case of Rd-valued continuous semi-martingales S. Extensions to the case 1 < p < cx~ as well as to the case of general Rd-valued semimartingales were obtained in [28,6] and [7].
390
F. Delbaen and W. Schachermayer
References [1] J.E Ansel and C. Stricker, Couverture des actifs contingents etprix maximum, Ann. Inst. Henri Poincar6 30 (1994), 303-315. [2] L. Bachelier, Th~orie de la speculation, Ann. Sci t~cole Norm. Sup. 17 (1900), 21-86. English translation: The Random Character of Stock Market Prices, E Cootner, ed., MIT Press, Cambridge, MA (1964), 17-78). [3] T. Bj6rk, Arbitrage Theory in Continuous Time, Oxford University Press (1998). [4] E Black and M. Scholes, The pricing of options and corporate liabilities, J. Political Econom. 81 (1973), 637-654. [5] E Cheridito, Arbitrage for fractional Brownian motion, ETH working paper (2000). [6] T. Choulli, L. Krawczyk and C. Stricker, On Fefferman and Burkholder-Davis-Gundy inequality for Emartingales, Probab. Theory Related Fields (1997). [7] T. Choulli, L. Krawczyk and C. Stricker, E-martingales and their applications in mathematical finance, Ann. Probab. 26 (1998), 853-876. [8] R.C. Dalang, A. Morton and W. Willinger, Equivalent martingale measures and no-arbitrage in stochastic securities market model, Stochastics Stochastics Rep. 29 (1990), 185-201. [9] E Delbaen, Representing martingale measures when asset prices are continuous and bounded, Math. Finance 2 (1992), 107-130. [10] E Delbaen, E Monat, W. Schachermayer, M. Schweizer and C. Stricker, Weighted norm inequalities and hedging in incomplete markets, Finance and Stochastics 1 (3) (1997), 181-227. [ 11 ] E Delbaen and W. Schachermayer, A general version of the fundamental theorem of asset pricing, Math. Ann. 300 (1994), 463-520. [ 12] E Delbaen and W. Schachermayer, The existence of absolutely continuous local martingale measures, Ann. Appl. Probab. 5 (4), (1995), 926-945. [ 13] F. Delbaen and W. Schachermayer, An inequality for the predictable projection of an adapted process, Sem. de Probabilites XXIX, Lecture Notes in Math. 1613, J. Azma, M. Emery, EA. Meyer and M. Yor, eds, Springer (1995), 17-24. [14] E Delbaen and W. Schachermayer, The Banach space of workable contingent claims in arbitrage theory, Annales de 1' I.H.P. 33 (1) (1997), 113-144. [15] E Delbaen and W. Schachermayer, The fundamental theorem of asset pricing for unbounded stochastic processes, Math. Ann. 312 (1998), 215-250. [16] E Delbaen and W. Schachermayer, A simple counterexample to several problems in the theory of asset pricing, Math. Finance 8 (1998), 1-12. [17] C. Dellacherie and P.A. Meyer, Probabilit~s et Potentiel. Th~orie des Martingales, Hermann, Paris (1980). [18] C. Dol6ans-Dade and EA. Meyer, In~galitis de normes avec poids, S6m. de Probabilit6 XIII, Lecture Notes in Math. 721, Springer (1979), 313-331. [19] D. Duffle and C.-E Huang, Multiperiod security markets with differential information; martingales and resolution times, J. Math. Econom. 15 (1986), 283-303. [20] B. Dupire, Model Art, Risk 6 (9) (1993). [21 ] A. Einstein, On the moment of small particles suspended in stationary liquid demanded by molecular-kinetic theory ofheat, Ann. Phys. 17 (1905). [22] A. Einstein, Zur Theorie der Brownschen Bewegung, Ann. Phys. 19 (1906), 371-381. [23] A. Einstein, Investigations on the Theory of the Brownian Movement, Dover, New York (1956). [24] N. E1Karoui and M.C. Quenez, Dynamic programming and pricing of contingent claims in an incomplete market, SIAM J. Control Optimiz. 33 (1995), 29-66. [25] M. Emery, Compensation de processus ~ variation finie non localement int~grables, S6m. de Probabilit6 XIV, Lecture Notes in Math. 784 (1980), 52-60. [26] H. F611mer and M. Schweizer, Hedging of contingent claims under incomplete information, Applied Stochastic Analysis, M.H.A.Davis and R.J.Elliott, eds, Gordon and Breach, London (1991), 389-414. [27] H. F611mer and D. Sondermann, Hedging of non-redundant contingent claims, Contributions to Mathematical Economics in Honor of G6rard Debreu, W. Hildenbrand and A. Mas-Colell, eds, North Holland, Amsterdam (1986), 205-223. [28] P. Grandits and L. Krawczyk, Closedness of some spaces of stochastic integrals, Seminaire de Probabilites XXXII (1998), 73-85.
Applications to mathematical finance
391
[29] J.M. Harrison and D.M. Kreps, Martingales and arbitrage in multiperiod securities markets, J. Econom. Theory 20 (1979), 381-408. [30] J.M. Harrison and S.R. Pliska, Martingales and stochastic integrals in the theory of continuous trading, Stochastic Process. Appl. 11 (1981), 215-260. [31] S.D. Jacka, A martingale representation result and an application to incomplete financial markets, Math. Finance 2 (1992), 239-250. [32] J. Jacod, Calcul Stochastique et Problbmes de Martingales, Lecture Notes in Math. 714, Springer, Berlin (1979). [33] I. Karatzas and S. Shreve, Methods of Mathematical Finance, Springer (1997). [34] A.N. Kolmogoroff, Grundbegriffe der Wahrscheinlichkeitsrechnung, Erg. Math., B.2 (1933). [35] D. Kreps, Arbitrage and equilibrium in economies with infinitely many commodities, J. Math. Econom. 8 (1981), 15-35. [36] D. Lamberton and B. Lapeyre, Introduction to Stochastic Calculus Applied to Finance, Chapman & Hall (1996). [37] S. Levental and A.S. Skorohod, A necessary and sufficient condition for absence of arbitrage with tame portfolios, Ann. Appl. Probab. 5 (1995), 906-925. [38] B.B. Mandelbrot, Forecasts of future prices, unbiased markets and "martingale" models, J. Business 39 (1966), 242-255. [39] J. Memin, Espaces de semi-martingales et changement de probabilitY, Z. Wahrscheinlichkeitsth. verw. Geb. 52 (1980), 9-39. [40] R.C. Merton, Theory of rational option pricing, Bell J. Econom. Manag. Sci. 4 (1973), 141-183. [41] M. Musiela and M. Rutkowski, Martingale Methods in Financial Modelling, Applications of Mathematics, Springer (1997). [42] E Protter, Stochastic Integration and Differential Equations, Springer, Berlin (1990),. [43] D. Revuz and M. Yor, Continuous Martingales and Brownian Motion, Springer, Berlin (1991). [44] L.C.G. Rogers, Arbitrage from fractional Brownian motion, Math. Finance 7 (1997), 95-105. [45] L.C.G. Rogers and D. Williams, Diffusions, Markov Processes and Martingales, Vol. 2: ItO Calculus, Wiley, Chichester (1987). [46] EA. Samuelson, Proof that properly anticipated prices fluctuate randomly, Industr. Manag. Review 6 (1965), 42-49. [47] W. Schachermayer, A counter-example to several problems in the theory of asset pricing, Math. Finance 3 (1993), 217-229. [48] W. Schachermayer, Martingale measures for discrete time processes with infinite horizon, Math. Finance 4 (1) (1994), 25-55. [49] A.N. Shiryaev, Essentials of Stochastic Finance. Facts, Models, Theory, World Scientific (1999). [50] M. Smoluchowski, Drei Vortriige iiber Diffusion, Brownsche Bewegung und Koagulation von Kolloidteilchen, Phys. Z. 17 (1916), 557-585. [51] C. Stricker, Arbitrage et lois de martingale, Ann. Inst. H. Poincar6 Probab. Statist. 26 (1990), 451-460. [52] J.A. Yan, Caracterisation d' une classe d'ensembles convexes de L 1 ou H 1, Seminaire de Probabilites XIV, Lecture Notes in Math. 784 (1980), 220-222.
This Page Intentionally Left Blank
CHAPTER
10
Perturbed Minimization Principles and Applications Robert Deville Department of Mathematics, University of Bordeaux, Bordeaux, France E-mail: deville@ceremab,u-bordeaux.fr
Nassif Ghoussoub Department of Mathematics, University of British Columbia, Vancouver, BC, Canada E-mail: [email protected]
Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Dentability and perturbed minimization principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. A general perturbed minimization principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. First examples of admissible classes of perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Martingales and admissible cones of perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4. Perturbed minimization and compactifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Perturbed minimization and differentiability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Smooth minimization principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Smooth norms and smooth bumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Generic differentiability of convex functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4. The existence of separating polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Non-smooth calculus in Banach spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. First order sub- and super-differentials for continuous functions . . . . . . . . . . . . . . . . . . . 4.2. The sub-differential of the sum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Second order differentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4. Stability of subdifferentials and superdifferentials . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5. Mean value theorems . . . . . . ..................................... 5. Application to critical point theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Palais-Smale minimizing sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. The mountain pass theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Applications to Hamilton-Jacobi equations in Banach spaces . . . . . . . . . . . . . . . . . . . . . . . .
395 397 397 399 401 403 405 406 407 409 412 414 415 418 419 422 422 423 423 424 426
6.1. The m a x i m u m principle for stationary first order Hamilton-Jacobi equations . . . . . . . . . . . . 6.2. The m a x i m u m principle for parabolic Hamilton-Jacobi equations . . . . . . . . . . . . . . . . . .
426 428
6.3. Uniqueness in second order Hamilton-Jacobi equations . . . . . . . . . . . . . . . . . . . . . . . .
430
6.4. The existence of viscosity solutions for Hamilton-Jacobi equations . . . . . . . . . . . . . . . . . . References
.....................................................
H A N D B O O K OF T H E G E O M E T R Y OF B A N A C H SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 393
431 433
394
R. Deville and N. Ghoussoub
Abstract Given a bounded below, lower semi-continuous function q~ on an infinite dimensional Banach space or a non-compact manifold X, we consider various possibilities of perturbing q~ by an element p of a reasonable class of functions A in such a way that for the new functional 4) - P, the minimization problem infx (q~ - p) is well-posed (i.e., every minimizing sequence is convergent). These perturbed minimization principles are quite powerful and have a wide array of applications in Banach space theory, potential theory, non-smooth analysis, non-linear analysis, the variational approach to differential equations as well as in the theory of viscosity solutions for Hamilton-Jacobi equations.
Perturbed minimization principles and applications
395
1. Introduction A Perturbed Minimization Principle is a statement of the following type:
Let A be a class of functions on a complete metric space X and consider a bounded below, lower semi-continuous function ~0:X --+ R U {+co}. Keeping in mind t h a t - in g e n e r a l - such a function does not necessarily attain its minimum, is it then possible to perturb ~0 by an element p of A in such a way that the new functional q) - p attains its minimum? a unique minimum? or that the minimization problem (X, ~0- p) is well-posed (i.e., every minimizing sequence is convergent). The problems involved in looking for such a statement revolve around the appropriate classes of functions A and the Banach spaces where it is possible. These choices are of course motivated by their scope of applicability albeit to differential equations, non-linear analysis or to Banach space theory. The first perturbed minimization principle is of course the celebrated: THEOREM 1.1 (Bishop-Phelps). Let C be a nonempty, closed, convex and bounded subset o f a real Banach space X. Then, the set M c o f all elements o f X* that attain their supremum on C is norm dense in X*. The connection with our topic is more clear if we state the following reformulation due to Br6ndsted and Rockafellar. COROLLARY 1.1. Let C be a nonempty closed convex subset o f X and let q) : C --+ R be lower semi-continuous, convex function that is bounded from below. Then, f o r any e > O, there exists x* ~ X*, Ilx* 11~< e such that q) + x* attains its minimum at some point xo ~ C.
Later, Ekeland [32] established the following very general and ultimately much more applicable theorem. THEOREM 1.2 (Ekeland). Let (X, d) be a complete metric space and consider a function q):X ~ ( - c o , +ec] that is lower semi-continuous, bounded from below and not identical to +ec. Let E > 0 and )~ > 0 be given and let x ~ X be such that q)(x) <, infx q9 + e. Then there exists xe ~ X such that:
(i) qg(xe) ~ qg(x). (ii) d ( x , x e ) <~ 1/)~. (iii) For each y 7~ xe in X, q)(y) > q)(xe) - e)~d(xe, y). Note that the class of perturbations in Ekeland's theorem is the space A of all Lipschitz functions, while in the Bishop-Phelps theorem, it is the much nicer space of continuous linear functionals, provided of course we limit ourselves to the problem of minimizing only convex functionals which is quite restrictive for most purposes. There are also other important differences between the 2 results. In Ekeland's theorem, the minimization problem becomes well-posed after perturbation, and the procedure can be localized around any given approximate minimum. This is not the case in the Bishop-Phelps theorem. However, we shall exhibit in this paper a whole range of admissible classes of perturbations that are
396
R. Deville a n d N. G h o u s s o u b
sometimes larger but more flexible than the class of continuous linear functionals but can also be smaller but more regular than the class of Lipschitz functions. First, let us show how Ekeland's result implies Bishop-Phelps', modulo the H a h n - B a n a c h Theorem. PROOF. Let q9 as in the above corollary. According to Ekeland's principle, there exists xo E C such that qg(x) ~> ~o(xo) - ~llx - xoll for all x 6 C. Consider in the Banach space X 9 R the subsets: C1 - {(x, t) E C @ R; t/> ~o(x)}, C2 - - { ( x , t) E X ~) R; t ~< ~p(xo) -
EIIx - xoll ].
They are both closed and convex and the interior of C2 is non empty and disjoint from C1. Using the H a h n - B a n a c h theorem, we can find G = (y*, c~) 6 X* E3 R = (X G R)* such that (y*, c~) r (0, 0), and m E R such that: (1) if x ~ C and t >~ qg(x), then y * ( x ) + at >~m; (2) if x 6 X and t <~ qg(xo) - ellx - x0ll, then y * ( x ) + at <<,m. Note that c~ = 0 is impossible. Otherwise, using condition (2), y * ( x ) <<,m for all x E X, which implies y* - - 0 and contradicts the fact that (y*, or) ~ (0, 0). The case c~ < 0 is not possible either: take x -- x0 and t large enough to get a contradiction in (1). Thus ot > 0. Replacing if necessary G by ~1 G , we can assume ot - 1. Since (x0, qg(x0)) 6 Cl n C2, m = y* (x0) + ~0(x0), (1) and (2) then imply: (3) for all x E C, y * ( x ) + qg(x) >~ y*(xo) + qg(x0); (4) for all x E X, y * ( x ) + qg(xo) - ellx - x0ll <~ y*(xo) + qg(xo). Relation (3) shows that y* + ~o attains its minimum on C at x0 and (4) shows that y* (x xo) ~< ellx - xoll for all x c X, whence Ily*[I ~< e. D To establish now the Bishop-Phelps theorem, let x* 6 X* and e > 0 be fixed. By the above, there exist y* ~ X*, Ily*ll ~< e, such that - x * + y* attains its minimum on C at some point x0 6 C, so z* = y* - x* attains its supremum on C and it is e-away from x*. We now describe an important improvement of Ekeland's theorem due to Borwein and Preiss [8]. Consider the class Q of all real-valued functions on X of the form 1 ~ q (X) -- -~ Z
, )2 # n d (x Vn. ,
n=l
where #n >~ 0, ZnC~__l # n - - 1 and where (Vn)n is some convergent sequence in X. The following result states that the Lipschitz perturbations obtained in Ekeland's theorem can be replaced by quadratic perturbations in the class Q. The relevance of this improvement becomes clear in the case when X is a Banach space admitting a differentiable norm (away from the origin), since the functions in the class Q are then differentiable everywhere. THEOREM 1.3 (Borwein-Preiss). Let qg : X --+ (-cx~, +c~] be a lower semi-continuous function that is bounded from below and not identical to +cx~ on the complete metric
Perturbed minimization principles and applications
397
space (X, d). Fix e > O, X > 0 and assume that u E X satisfies qg(u) < infx q9 + e. Then there exist q E Q and v~ ~ X such that: (i) qg(v~) < infx ~o + e.
(ii) d(u, v~) < X. (iii) For all x E X, qg(x) + 2eX-Zq(x) >~qg(v~) + 2eX-Zq(v~). In view of the three perturbed minimization principles discussed above, it is natural to inquire about the classes of functions on a complete metric space X that are suitable to be perturbation spaces for a minimization principle. In the next section, we shall isolate a fairly general condition on a class A of continuous functions on X that makes it eligible to be a perturbation space. We shall then say that A is an admissible class. We will see that the perturbations used in Ekeland's principle as well as the one of Borwein-Preiss, readily satisfy our condition. The same will hold for various spaces of smooth functions defined on a suitable Banach space. However, the proofs of the facts that, on certain Banach spaces, the spaces of linear functionals or cones of plurisubharmonic functions are admissible classes, are more involved. The rest of the paper contains various applications of these principles to Banach space theory, potential theory, non-smooth analysis, non-linear analysis, the calculus of variations as well as to the theory of viscosity solutions for Hamilton-Jacobi equations.
2. Dentability and perturbed minimization principles 2.1. A general perturbed minimization principle Let (X, d) be a metric space and let (A, 3) be a metric space of real valued functions defined on X. For any subset F of X, we shall denote by ,AF the class of functions in A that are bounded above on F. For f E AF, and t > 0, we denote by S(F, f, t) the following slice of F
S(F, f, t) = {x E F" f (x) > s u p f ( F ) - t}. DEFINITION 2.1. The space (X, d) is said to be uniformly A-dentable if for every nonempty set F C X, every f E AF, and every e > 0, there exists g 6 ~AF such that 6(f, g) <<,e and a non-empty slice S(F, g, t) of F such that diam S(F, g, t) ~< e. In the sequel, we associate to the metric space (X, d), the space X - X • R equipped with the pseudo-metric
~l((x, X), (y, #)) -- d(x, y). We also associate to (A, 3) the class A of functions of the form f "-- (f, - 1 ) which act on X via (f, - 1)(x, )~) -- f ( x ) - X. We equip A with the distance ~ ( f , ~,) - 3(f, g). DEFINITION 2.2. Say that a function qg" X --+ R U {+cx~} strongly exposes y in X if:
398
R. Deville and N. Ghoussoub
(i) qg(y)= inf{q)(x); x 6 X} and (ii) d(y, Yn) --+ 0 whenever q)(Yn) --+ q)(Y). We then say that y is a strong minimum for q). Note that, in particular, y is then a unique minimum for q). Denote by D(qg) the set {x ~ X; qg(x) < +c~}. We can now state the following general perturbed minimization principle. THEOREM 2.1. Let A be a class of bounded and continuous functions on a complete metric space (X, d) equipped with a metric 6. Suppose ,,4 satisfies the following conditions: (i) There is a K > 0 such that 6 ( f , g) >~ K sup{If(x) - g(x)l; x e X} f o r all f, g ~ A. (ii) (A, 6) is a completemetric space. (iii) The product space (X, d) is uniformly A-dentable. Then, f o r any lower semi-continuous function q):X --+ R U {+cx~} that is bounded below with D(~o) 7~ 0, the set {g E A; q9 - g attains a strong minimum on X}
is a dense G a subset G of,4. We shall then say that (,4, 6) is an admissible class of perturbations for the space (X, d). PROOF. We claim that the set /.gn = {g c A ; 3Xn e X with ( ~ o - g ) ( x n ) < inf{(~o - g)(x): d ( x , x n ) >~ I / n } } is an open dense subset of A. Indeed, ~n is open because of assumption (i). To see that/gn is dense, let g E A and e > 0. We need to find h E A, 6(h, g) < e, and Xn E X such that
(~o - h)(xn) < inf{(~o - h)(x); d ( x , x n ) >~ 1/n}.
(*)
To do that, note that the functional (g, - 1 ) in A is bounded above on the epigraph F of q) in ~', hence for any e' > 0, there exists a non-empty slice S - S(/t, F, t) of F with diameter less than e ' and determined by a function h = (h, - 1 ) ~ A verifying 6(h, g) < d. Take e ~ < min{e, I / n } and pick any (Xn,)~n) E S. For any x E X such that d ( x , x n ) 1/n, we have that (x, qg(x)) ~ F \ S, so that h (x) - q9(x) ~< sup h - t < h (Xn) - )~n. F
Since )~n >~ q)(Xn), we obtain that the function h verifies (,). Since A is a complete metric space, G - On>~l L/n is a dense G~-subset of A, by the Baire category theorem. We claim now that if g ~ G, then q) - g attains a strong minimum on X. Indeed for each n ~> 1, there exists Xn E X such that
(~o - g)(xn) < inf{(cp - g)(x); d ( x , x n ) >f 1/n}.
Perturbed minimization principles and applications
399
We necessarily have that d (Xp, Xn) < 1/n for each p > n. Indeed, if not then by the definition ofxn, we have (qg-g)(xp) > (q)-g)(xn). On the other hand, we willhaved(xn,Xp) >I 1/n >~ 1/p which gives, by the definition of Xp, that (q9 - g)(xn) > (q9 - g)(xp). This is clearly a contradiction. Thus (Xn)n is a Cauchy sequence converging to some x ~ 6 X and we claim that x ~ is a strong m i n i m u m for 99 - g. Indeed, since 99 is lower semi-continuous,
(~p - g ) ( x ~ ) <~ liminf(tp - g)(Xn) <~liminfinf{(~o - g)(x)" d(x,xn) >~ I/n} <~ inf{ (q9 - g)(x); x e X \ { x ~ } }. Moreover, let (y~) be a sequence in X such that (q9 - g)(yn) converges to ((p - g)(x~). Let us assume that (Yn) does not converge to x ~ . Extracting, if necessary, a subsequence, we can assume that there exists e > 0 such that for all n, d (y,, x ~ ) >~ e. Thus there exists an integer p such that d (Xp, y,) ~ 1/p for all n. Consequently
(cp - g ) ( x ~ ) <~ (q) - g)(Xp) < inf{(~o - g)(x); d ( x , x p ) > l / p } <~ (~o - g)(yn) for all n, which contradicts the fact that limn ((p - g)(y,) = (q9 - g)(x~).
2.2. First examples of admissible classes of perturbations One can easily recover Ekeland's result as well as the theorem of Borwein and Preiss [8] from T h e o r e m 2.1. For that, we m a y consider the cones r (respectively .A2) consisting of those functions f on (X, d) of the form
f(x)--~)~,~d(x,xn)
(respectivelyf(x)--Z)~nd2(x,x~))
n
n
with Xn >~ 0. We assume here, without loss of generality, that d is a b o u n d e d metric on X. Let us check that they are admissible cones. Indeed, suppose F is a closed subset of ~" - X • R such that/~ -- ( h , - 1 ) 6 .A1 is b o u n d e d above on F. For small r > 0, consider a point (x0, so) 6 F such that h(xo) - so > suPF/~ -- r 2. Let k be the functional in .A1 defined by/r -- (h - r d (, x0), - 1) and let S be the slice of F given by
s - { (s, ~) e F; /,(s, ~) > ~(xo, so) - r2}. It is easy to see that the d - d i a m e t e r of S is less than r, which means that X is .A1-uniformly dentable. A similar proof works for .A2. Note that Ekeland's result would then follow from T h e o r e m 2.2 and the triangular inequality. Note also that we could have used the space .A - Lip(X) of Lipschitz functions on X as an admissible space. We shall now investigate the possibility of having other classes of functions as perturbation spaces. For simplicity, we shall only deal, in the sequel, with the case where X is a Banach space.
400
R. Deville and N. Ghoussoub
A function b ' X --+ R is said to be a bump function on X if it has a bounded and nonempty support. PROPOSITION 2.1. Let ,A be a Banach space o f continuous functions on a Banach space
X satisfying the following properties: (i) For each g 6 A, IlgllA t> Ilgll~ = sup{lg(x)]; x E x}. (ii) A is translation invariant, i.e., if g E A and x ~ X, then rxg : X --+ R given by rxg(t) = g(x + t) is in A and II~xgllA = IlgllA. (iii) A is dilation invariant, i.e., if g ~ A and ot ~ R then g~ : X --+ R given by g~ (t) = g(ott) is in A. (iv) There exists a bump function b in A. Then A is an admissible space o f perturbations f o r the space X. PROOF. According to (ii) and (iv), we can find a bump function b in A such that b(0) r 0. Using (iii) and replacing b(x) by otl b(oQx) with suitable coefficients C~l, ot2 E R, we can assume that b(0) > 0, Ilbll~ < e and b(x) - 0 whenever Ilxll ~> e. Let now ~ - (g, - 1) E A be bounded above on a closed subset F of X and let (x0, so) F be such that g(xo) - so > sup F ~ - b(0) and consider the function h(x) -- b(x - xo) and k = (g + h , - 1 ) . By (ii), h 6 A and Ilhll.a - IlbllA < e, which implies that Ilg - ~:IIA < e. On the other hand, consider the following slice of F,
S--
{
]
(x, s) 6 F; k(x, s) > sup ~ . F
It is non-empty since it contains (xo, so). On the other hand, if (x, s) 6 F and IIx - xoll ~ e, then b(x - x0) = 0 and (x, s) cannot belong to S. It follows that the d-diameter of S is less than 2e. Consequently, A is an admissible family of perturbations. E] COROLLARY 2.1 (Localization). Assume A is a Banach space o f bounded continuous functions on X satisfying conditions (i)-(iv) above. Then, f o r some constant a > O, depending only on X and A, the following holds: I f 99 : X --+ R U { + ~ } is lower semi-continuous and bounded below with D(99) ~: 0 and if Yo ~ X satisfies qg(y0) < infx q9 + ae 2 f o r some e ~ (0, 1), then there exist g ~ A and xo ~ X such that (i) Ilx0 - y0ll ~< e, (ii) Ilgll,a ~< e, (iii) q9 + g attains its minimum at xo. PROOF. We can clearly assume that there exists a bump function b in A with b(O) - 1, 0 ~< b ~< 1 and such that the support of b is contained in the unit ball of X. Hypothesis (i) implies that M "-- IlbllA/> Ilbll~ - 1. Let a -- 1 / 4 M and suppose that e and Y0 are given. Define the function
Perturbed minimization principles and applications
401
We have qS(yo) < infx ~o - a~ 2 and qS(y) >~ infx ~o whenever Ily - yoll ~> e. From Proposition 2.1 and Theorem 2.1, we can find xo E X and k E A such that []kl[A ~< min{e/2, ae2/2} and q5 4- k attains its minimum at xo. The above conditions imply that ]Ix0 y0ll < e. Thus, the function g e ,A defined by g(x) -- - 2 a e Z b ( x--xo ----E- ) + k(x) satisfies claims (i), (ii) and (iii) of the corollary. E] REMARK 2.1. Again, we can recover Ekeland's minimization principle on Banach spaces from the (easily verifiable) fact that the space ~41 of all bounded Lipschitz functions on X equipped with the norm [
[If[lA, = sup{ If(x)l; x E X} + sup{ /
If(x) IIx -
f(y)l
; x=/=y}
yll
satisfies the conditions of Proposition 2.1 and hence it is an admissible space of perturbations. Note that, as an additional bonus, we get that the perturbation is also small in the uniform norm as well as in the Lipschitz norm. To derive an analogue of the Borwein-Preiss Theorem, we can consider the space .42 of all bounded Lipschitz functions f on X that also verify the following second order condition
I[fl[
_ supJ If(x + 2h) - 2 f ( x + h) + f ( x ) [ 9x, h E X } / h2
<ec.
The space .,2[2 equipped with the norm IIf ]IA2 = IIf [IA~ + Ilf II is also an admissible space of perturbations. Clearly, the above norms will correspond to the C 1 and C2-norms whenever the functions are differentiable. But since X is in general infinite dimensional, we have to deal with two types of difficulties: firstly, the appearance of various different and generally nonequivalent types of differentiability and secondly, the problem of admissibility of these spaces of differentiable functions which requires extra assumptions on the Banach spaces involved. We deal with some of these problems in Section 3.
2.3. Martingales and admissible cones of perturbations In this section, we briefly sketch the relationship between admissible cones of perturbations and the theory of balayage associated to such a cone. Let (s I7, P) be a probability space and let (17n)n be an increasing sequence of suba-fields of I7 (i.e., r n C 17n+1). Recall that a sequence (Fn)n of real-valued integrable random variables on s is said to be a martingale (respectively a submartingale) with respect to ( r n ) n if for each n E N, (i) Fn is 17n-measurable, and (ii) fA Fn dP -- fa Fn+l dP (respectively fa Fn dP <~fa Fn+l d P ) for every a E r n .
402
R. Deville and N. Ghoussoub
In other words, Fn = E[Fn+I; r n ] (respectively Fn <~ E[Fn+I; r n ] ) where the latter denotes the conditional expectation of Fn+l with respect to the a-field r n . To extend the above notions to a nonlinear setting, we proceed in the following way. Suppose that X is a complete metric space and let A be a convex cone of real valued continuous functions on X. DEFINITION 2.3. Say that a sequence (Fn) n of X-valued random variables on ~2 is an A-martingale if for any ~0 E A, the process (~p o Fn)n is a real-valued submartingale. As stated above, there is a close relation between the almost everywhere convergence of X-valued A-martingales and the admissibility of A as a cone of perturbations for a minimization principle on X. This program has been developed extensively in the upcoming monograph [48]. We shall here describe briefly a few - particular but i m p o r t a n t - examples that have been developed in Ghoussoub and Maurey [46,47]. If X is a Banach space, then one can define the Bochner integral of an X-valued random variable F as well as its conditional expectation, first by defining them for step functions and then by taking limits in the usual fashion and as in the real valued case. Therefore, one can say that a sequence (Fn)n of X-valued Bochner integrable random variables is a vector martingale if for all n E N, we have that almost surely E[Fn+I; r n ] = Fn, where both are regarded as X-valued random variables. It is easy to check that if X is a Banach space, the notion of X-valued martingales coincides with the notion of A-martingales where A is the cone of continuous convex (or linear) functionals on X. The convergence of uniformly bounded Banach space valued martingales does not always hold and it is closely related to the geometry of the space involved. Actually, one can show the following result. THEOREM 2.2 (Bourgain-Stegall). If C is a closed convex bounded subset of a Banach space X, then the following properties are equivalent: (i) Every C-valued martingale converges almost surely. (ii) The space X* of continuous linear functionals is an admissible space of perturbations on C. Such sets are then said to possess the Radon-Nikodym property. See for example [46]. Now, if X is a manifold or just a quasi-Banach space (as in the case of L p (0 < p < 1)) the notion of a vector integral is not well defined in general, and one can already see that the concept of an A-martingale is more appropriate. We first recall the relevant concepts. Let X be a vector space. A map x ~ llx II from X into R + is called a quasi-norm if (i) Ilx l] > 0 when x -r 0.
(ii) II~x II - I ~ 1 IIx II, for ~ ~ C, x ~ X. (iii) Ilxt + x211 ~< C(llxl II + IIx211) for all xl, x2 6 X. Here C/> 1. We call II II a p-norm for 0 < p ~< 1, if in addition it is p-subadditive, that is (iv) Ilxl +x211 p ~< Ilxt IIp + IIx211p f o r x l , x 2 E X. The Aoki-Rolewicz theorem asserts that every quasi-norm is equivalent to a p - n o r m for some p (0 < p ~< 1). A complete quasi-normed vector space X will be called a quasiBanach space. If the quasi-norm on X is also p-subadditive, we say that X is a p-Banach
Perturbed minimization principles and applications
403
space. For the basics about non-locally convex vector spaces, we refer the reader to the book [57]. Let A = {z 6 C, Izl < 1 } be the open unit disc. Denote by 0A or T the unit circle {z C; Iz[ = 1} and by z~ = A U 0A. DEFINITION 2.4. Let g : X --+ [ - c o , + c o ) be an upper semi-continuous function. It is said to be plurisubharmonic on X if for every x, y 6 X we have
g(x) ~
f0 zrg(x + ei~
dO 2re
We denote by PSH(X) the class of all such functions. LIPp(X) will denote the set of functions g on X satisfying for some K > 0, Ig(x) - g(Y)l ~< Kllx - y[I p for all x, y in X. We shall write PSHp(X) for PSH(X) A LIPp(X) and we shall equip it with the following norm Ilgllp -- max{Ig(O)l, sup{Ig(x) - g(y)l/llx - YlIP; x, y 6 X, x - r y}}. Here is the Plurisubharmonic perturbed minimization principle [47]. See also [45]. THEOREM 2.3 (Ghoussoub-Maurey). Let X be a p-Banach space for some p (0 < p <, 1). The following assertions are equivalent: (i) All bounded X-valued PSHp-martingales converge almost surely. (ii) The space PSHp(X) is an admissible class of perturbations for any closed bounded subset C of X: that is, for every bounded below, lower semi-continuous function q) on C, the set
{ g E PSHp (X); 99 - g attains a strong minimum on C } is a dense G~ in PSHp(X). Examples of quasi-Banach spaces that verify the above principle are Lebesgue spaces LP, Hardy spaces HP or the Schatten spaces C p where 0 < p ~< 1. Note that when 1 < p < ~ , the above mentioned spaces are all reflexive Banach spaces and therefore one can apply the stronger linear perturbation principle (Theorem 2.2).
2.4. Perturbed minimization and compactifications Suppose now that C is a subset of a compact metric space (K, d) a n d - for simplicity- let A be a Banach space of continuous functions on K equipped with a norm II II. A natural question is to identify the topological structure of C as a subset of K that insures that A is an admissible cone of perturbations on C. It turns out that the concept of As-sets - defined b e l o w - is the appropriate notion for this description.
404
R. Deville and N. Ghoussoub A
We start by associating to any subset L of K, the function ~/~ as follows:
A
{
j
qJL (x) -- sup h (x) - sup h; h 6 A and IIh II ~< 1 . L D E F I N I T I O N 2.5. With the above notations:
(i) A closed subset L of K is said to be A-convex in K if it is of the form
L-
N { x E K; hi(x) ~ ~,i} i6I
where the functions hi belong to A and )~i E R. (ii) A subset C of K is said to be an A~-set in K if K \ C is a countable union of compact A-convex sets (Kn)n. (iii) C is a strict A~-set in K if in the above representation, we have infx~c qJKn (x) > 0 for each n 6 N. Modulo the appropriate hypothesis, one can show in quite a general setting that A is an admissible space of perturbations on any strict A~-subset C of K. An interesting aspect of this approach is that the perturbed minimization principle also holds not just for lower semi-continuous functions but also for strict A~-functions: i.e., those whose epigraph in K x R is a strict A~-set. See examples below. Conversely, if one starts with a general complete metric space C and a Banach space A of continuous functions on C that forms an admissible vector space of perturbations, then one can find a compactification K of C in which C sits as a strict A~-set. Again, this program is developed in its full generality in the upcoming monograph [48] and we shall only state here the linear c a s e - studied extensively in [46] - where C is a closed convex subset of a Banach space X and A is the dual space X*. The case where A is the space of weakly continuous harmonic functions was considered in [49]. In the linear case, C is assumed to be a subset of a dual space Y* (usually the double dual X** of the Banach space X containing C. A-convexity coincides then with the usual convexity, meaning that A-convex compact sets are exactly the convex weak* compact subsets of Y*. In this case, A~-sets are called w*-H~-sets (i.e., their complement is a countable union of convex weak* compact sets) since A-convex compact sets are - in this setting - just intersections of hyperplanes determined by linear functionals in Y. Here is the main result of the linear theory [46]. THEOREM 2.4 (Ghoussoub-Maurey). Let K be a w*-compact convex set in some dual Banach space Y* and let C be any strict w*-H~ -subset of K. For any strict w*-H~-function llt on C, we have: (1) The set {y ~ Y; lp + y attains its maximum on C} is dense in Y. (2) If K is w*-metrizable, then the set {y ~ Y; ~ + y exposes C from below } is a dense G~ in Y: i.e., (Y, el Hi) is an admissible space of perturbations for (C, w*). (3) If C is norm separable then the set {y ~ Y; ~p + y strongly exposes C from below} is a dense G~ in Y: i.e., (Y, ]1 I]) is an admissible space of perturbations for (C, I] I])-
Perturbed minimization principles and applications
405
Moreover, if in (2) and (3), ~p is assumed to be concave then the minimum is (uniquely!) attained on an extreme point o f C.
Here are a few basic examples of strict w*-H~-sets and functions. PROPOSITION 2.2. Let K be a weak*-compact convex subset o f a dual Banach space Y*. Then (1) A countable intersection o f open half-spaces in K determined by functionals in Y is a strict w*-H~-set in K. (2) I f f is a functional in Y** that is in the first Baire class f o r the weak*-topology, then it is a strict w*-H~-function on K and any half-space determined by f in K is a strict w*-H~-subset o f K. (3) If 99 is convex, w*-lower semi-continuous and norm-Lipschitz on K, then lp = -cp is a strict w*-H~-function on K. (4) The supremum o f a bounded sequence o f strict w*-H~-function on K is also a strict w*-H~ -function. (5) I f C is a strict w*-H~-set in K and if K is weak*-metrizable then any relatively weak*-closed subset D o f C is a strict w*-H~-set in K and any weak*-lower semicontinuous function on C is a strict w*-H~-function. (6) I f C is a norm separable strict w*-H~-set in K, then any norm closed subset o f C is a strict w*-H~-set in K and any norm-lower semi-continuous function on C is a strict w*-H~-function. In particular, if K is norm separable, then any norm-lower semi-continuous function on K is a strict w*-H~-function.
Finally, we have the following compactification theorem [46]. THEOREM 2.3 (Ghoussoub-Maurey). Let C be a closed convex bounded subset o f a separable Banach space X. The following are then equivalent: (i) C has the R a d o n - N i k o d y m property. (ii) (X*, II II) is an admissible space o f perturbations f o r (C, II II). (iii) There exist a separable subspace Y o f X* and an isometric embedding T : X --+ Y* such that T (C) is a strict w*-H~ in Y*.
3. Perturbed minimization and differentiability Let again q9 be a function defined on a Banach space X with values in R U { + ~ }. In principle, we are only concerned here with the concepts of Frgchet and G~teaux-differentiability. However, since the arguments in both cases are similar, we shall avoid repetition, by working with the notion of differentiability associated with any bornology on X. Recall that a bornology fl is just a class of bounded subsets of X such that the topology r/~ on X* that corresponds to the uniform convergence of linear functionals in X* on the sets of/3 defines a locally convex topology on X*. Note that if/3 is the class of all bounded subsets (respectively all singletons) of X, then r/~ coincides with the norm (respectively weak*) topology on X*.
406
R. Deville and N. Ghoussoub
DEFINITION 3.1. Say that q9 is ~-differentiable at xo ~ D ( f ) with/3-derivative qg'(xo) =
p 6 X* if for any A 6 t3, lim t -1 [qg(x0 + th) - ~p(xo) - (p, th)] - 0 t--+0
uniformly for h E A. When 13 is the class of all bounded subsets (respectively all singletons) of X, the 13differentiability coincides with the classical notion of Fr~chet-differentiability (respectively GCtteaux-differentiab ility). We shall denote by cl~(x) the space of all real valued, bounded Lipschitz and /3differentiable functions g on X such that g~:(X, II II) ~ (X*, ~ ) is continuous, endowed with the norm IIg II~ = sup { Ig (x)l; x E x } + sup { IIgt (x)ll; x E x ] = IIg II~ + IIg' II~ . We shall use the notation c l ( x ) (respectively c l ( x ) ) when we are dealing with Fr6chet (respectively G~teaux) differentiability. Analogously, we can define the spaces C~ (X) (respectively C~ (X)) equipped with the C2-norm. 3.1. Smooth minimization principles Two immediate applications of Theorem 2.1 and Proposition 2.1 are the following [24]: THEOREM 3.1 (First order smooth minimization principle). Suppose X is a Banach space on which there exists a C1-Frechet smooth (respectively, G~teaux-differentiable) and Lipschitz continuous bump function. Then C~ (X) (respectively C1F(X)) is an admissible space of perturbations: i.e., for each e > 0 and f o r each lower semi continuous and bounded below function ~p: X --+ R U {+cx~ } such that q9 is not identically equal to +c~, there exists a C 1 (respectively, GCtteaux differentiable), Lipschitz continuous function p such that: (1) IIPlI~ = sup{Ip(x)l; x E X} < E, (2) IIP'll~ = sup{llP'(x)llx*; x E X} < e, and (3) q9 + p has a strong minimum at some point xo E X. THEOREM 3.2 (Second order smooth minimization principle). Suppose X is a Banach space on which there exists a C2-Frechet smooth bump function b with Lipschitz derivative. Then C2 (X) is an admissible space of perturbations: i.e., f o r each e > 0 and f o r each lower semi continuous and bounded below function qg : X --+ R U {+cx~} such that 99 is not identically equal to +cx~, there exists a C 2-function p on X with pt Lipschitz continuous such that: (1) IIPlI~ = sup{Ip(x)l; x ~ X} < e, (2) IIP'll~ = sup{llP'(x)llx*; x E X} < e, (3) IIP"II~ = sup{llP"(x)llB(x); x ~ X} < e, and (4) q9 + p has a strong minimum at some point xo E X.
Perturbed minimization principles and applications
407
REMARK 3.1. It is worth comparing the above with the result of Borwein and Preiss (Theorem 1.3). Indeed, if X is a Hilbert space or more generally an LP space (1 < p < ec), then the perturbation q in Theorem 1.3 can be taken to be of the form q ( x ) = 89 - wll 2 for some w (usually not equal to re). This quadratic perturbations is more explicit than the above, however it lacks the boundedness properties of the perturbations in Theorem 3.2.
3.2. Smooth norms and smooth bumps The existence of a smooth bump function is a key hypothesis in the smooth minimization principles stated above. From now on, we shall use the following notation: (H1) There exists a C 1-smooth, Lipschitz continuous bump function b: X ~ R. (H1G) There exists a Gfiteaux-differentiable, Lipschitz continuous bump function b : X - - + R.. (H2) There exists a C2-smooth bump function b : X ~ R with Lipschitz derivative. We begin this section by pointing out that these assumptions are necessary for the validity of the smooth minimization principles. PROPOSITION 3.1. Let qg: X --+ R be a (Lipschitz continuous)function on X satisfying 99(x) > Of o r all x E X and l i m l l x l l ~ qg(x) = 0. (1) I f there exists a C 1-smooth, Lipschitz continuous function g such that 99 - g has a global minimum attained at some point xo, then X satisfies (HI). (2) I f there exists a Gfiteaux-differentiable, Lipschitz continuous function g such that 99 - g has a global minimum attained at some point xo, then X satisfies (H1G). (3) I f there exists a C2-smooth function g with Lipschitz derivative such that 99 - g has a global minimum attained at some point xo, then X satisfies (H2). Our next result, due to Leduc [59], asserts that the existence of a smooth bump function can be characterized by the existence of a smooth function looking like a norm, but not convex in general. PROPOSITION 3.2. Let X be a Banach space. The following assertions are equivalent: (i) There exists on X a Lipschitz continuous function d : X --+ R +, which is C l-smooth on X\{0}, and a constant K > O, such that [Ixll ~< d ( x ) <~ Kllxll f o r all x E X, and d(s = s f o r all x E X and all s ~ O. (ii) X satisfies (H1).
PROOF. Let b : X --+ R be a C 1-smooth, Lipschitz continuous bump function, with support in the unit ball and such that 0 ~< b ~< 1 and b(0) > 0. The function d : X ~ R defined by d(0) - 0 and d ( x ) - ( f + ~ b ( t x ) dt) -1 for x ~ 0 satisfies the required properties. Conversely, let r :R --+ R be a continuously differentiable function such that r = 0 on (-cx~, 1] U [3, +cx~) and r(2) ~ 0. The function b : X --+ R defined by b(x) = r ( d ( x ) ) for x E X is a continuously Fr6chet differentiable bump on X. E] An immediate application of the above proposition (and a similar proof in the case of Gfiteaux-differentiability) is that the existence of a smooth norm implies the existence of a
R. Deville and N. Ghoussoub
408
smooth bump function. That is if X has an equivalent Fr6chet-differentiable norm (respectively, G~teaux-differentiable norm), then X satisfies (HI) (respectively, (HI G)). A consequence of the above result is that every reflexive space and every space with separable dual satisfy (HI), (because these spaces admit an equivalent Frechet-differentiable norm). Finally, we note that Haydon [52,54,53] constructed an example of a Banach space which satisfies (HI) (this space even admits a C ~ - s m o o t h Lipschitz continuous bump function), but has no equivalent Fr6chet-differentiable norm, thus showing that the converse is not true in general. On the other hand, for separable Banach spaces, a converse is available: THEOREM 3.3. Let X be a separable Banach space. The following conditions are then equivalent: (1) X satisfies (H 1). (2) There exists on X an equivalent Frdchet-differentiable norm. (3) The dual X* of X is separable. PROOF. (3) implies (2) is a classical result of Kadec and Klee. Since X* is separable, we let N* (x *) --
Ilx*II+ ~
2-n dist(x *, En ),
n
where (En)n is an increasing sequence of finite dimensional subspaces such that X* = Un En. It is easy to see that N* is weak*-locally uniformly convex: that is, a sequence (Xn*)n on the N*-unit sphere of X* norm-converges to x*, whenever it weak*-converges to x* and N* (x*) - 1. The predual norm N on X is then Frdchet differentiable on X \ {0} and its derivative is (norm to norm)-continuous from X \ {0} to X*. For more details, we refer the reader to the monograph [25]. The proof that (1) implies (3) uses the smooth minimization principle, so we shall provide a sketch. Let b be a C l-smooth Lipschitz continuous bump function on X. Denote U -- {x E X; b(x) ~ 0} and define f ' U --+ R by f ( x ) = 1/b(x). We claim that f is C 1smooth on U and that {ft (x); x E U} is norm dense in X*. Indeed, fix p E X* and e > 0. The function q)" X --+ R U {+oo} defined by qg(x) = - p ( x ) + f ( x ) if x E U and qg(x) -- cx~ otherwise, is lower semi continuous on X. By the first order smooth minimization principle, there exists a C 1-function g on X such that IIg~Iloc < e and q9 + g has a minimum at some point x0 E U. We have f ' ( x o ) - p + g'(xo), hence I[f'(x0) - Pll < e which proves the claim. On the other hand, since f t is continuous and U is separable, {f~ (x); x 6 U} is separable. Together with the claim, this implies that X* is separable. D REMARK 3.2. As noted above, if X is a Banach space which admits an equivalent G~teaux-differentiable norm, then there exists on X a Lipschitzian GSteaux differentiable bump function b such that b t is norm to weak* continuous. In particular, X satisfies (H1G). The class of Banach spaces satisfying (H1 G) is very large: Banach spaces which are weakly compactly generated (in particular, all reflexive Banach spaces and all separable Banach spaces) admit an equivalent G~teaux-differentiable norm. Therefore, they satisfy (H1G). On the other hand we shall see in Section 3.3 that goc does not satisfy (H1G).
Perturbed minimization principles and applications
409
We conclude this section by showing that the existence of a smooth bump function implies the existence of other bump functions enjoying special properties. PROPOSITION 3.3. Let X be a Banach space on which there exists a C 1-smooth Lipschitz continuous bump function b, then: (1) There exists a C 1-smooth Lipschitz continuous bump function bl such that b~ ~ 0 and bl (x) = 1 = maxx bl f o r x in a neighbourhood o f O. (2) There exists a C 1-smooth Lipschitz continuous bump function b2 which attains its strong maximum at O. (3) I f moreover X is separable and infinite dimensional, then there exists a C l - s m o o t h bump function b3 such that b~3(X) -- X*. PROOF. (1) By translation and replacing if necessary b by - b , we can assume that b(0) > 0. Let ~0: R --+ R be a C l - s m o o t h Lipschitz continuous function such that ~p(t) = 0 if t <~ 0 and ~0(t) = 1 if t >~ b(0)/2. The function bl = ~o o b has the required properties. (2) Apply the smooth minimization principle to a constant function to get a function g which attains a strong minimum at some point x0. Replacing b by b(. - x0), we can assume that x0 = 0. Let a > f (0) such that IIx II < 1 whenever f (x) < a. Let ~ : R --+ R be a Cl-smooth Lipschitz continuous function such that gr(t) = 0 if t/> a and 7r' (t) < 0 if t < a. The function b2 = ~ o b has the required properties. (3) The construction of b3, which is given in [2], requires a series of functions of the form x --+ cnbl (anx - Xn).Yn, where an, Cn > O, Xn C X and Yn E X* are suitably chosen. Observe here that under the assumptions of (3), there exists on X an equivalent norm I1.11 which is Frechet-differentiable on X\{0}. If b : X --+ R is of the form b(x) = qg(llx II), where q) is a C l bump function on R with support in (0, + e c ) , then, according to James theorem, the set {a.b'(x); a > O, x ~ X} coincides with X* if and only if X is reflexive. Thus the bump function b3 cannot be of this form in general. D For higher order smoothness, the situation is more delicate. Let us mention here that for 2 ~< p < + ~ , the norm of the LP spaces is C2-smooth and its derivative is Lipschitzian on the unit sphere. Therefore these spaces satisfy (H2) and the second order minimization principle can be applied to these spaces. On the other hand, if X satisfies (H2), then X is superreflexive of type 2 (see [25]). If one does not require a Lipschitz condition on the derivative, the situation is different: for instance, there exists on spaces of continuous functions on countable compact, an equivalent C ~ - s m o o t h norm [53], although these spaces are not reflexive. Actually one can even construct an equivalent analytic norm on these spaces. Finally, in spaces not containing co, the existence of a C ~ - s m o o t h bump function is equivalent to the existence of a polynomial P on X such that P (0) = 0 and P (x) >~ 1 whenever Ilx II = 1.
3.3. Generic differentiability o f convex functions We show in this section how the smooth minimization principle is an efficient tool to establish generic differentiability of convex functions. The following is a theorem of Ekeland and Lebourg [32,33,59].
410
R. Deville and N. Ghoussoub
THEOREM 3.4. A s s u m e X satisfies (HI), then it is an A s p l u n d space, i.e., every convex and continuous function ~p " X --+ R is differentiable on a dense set. PROOF. Fix xo E X and e > O. Set F ( x ) --
1 - q ) ( x ) + b2(m(x_xo) ) +cxz
if b ( M ( x - xo)) 7~ O, otherwise.
The function F is lower semi-continuous, and bounded below for M large enough. By the smooth minimization principle, there exists g" X --+ R, C 1-smooth, such that F + g attains its m i n i m u m at some point Xl where b ( M ( x l - xo)) ~ O. If M > 0 is large enough, then Ilx 1 - x0 II < e. Moreover, if we let
f (x) -- b 2 ( M ( x _ xo)) + g ( x ) ,
the function q) - f has a local m a x i m u m at x l, so q9 is superdifferentiable at the point X l and hence at each point of a dense subset D of X. Since q9 is convex and continuous, q) is subdifferentiable at every point (by the H a h n - B a n a c h theorem). Therefore, q) is differentiable at every point of D. D An immediate application of T h e o r e m 3.4 is: COROLLARY 3.1. A s s u m e that the norm o f X satisfies f o r every x ~ X,
lim sup h--,0
Ilx + h II + IIx - h II - 21Ix II
> 0.
(,)
Ilhll
Then X does not satisfy (HI) (i.e., there exists no bump b" X --+ R which is C l - s m o o t h and Lipschitz continuous).
PROOF. Indeed, a norm satisfying (,) is an example of a nowhere differentiable convex continuous function. This corollary allows to deduce that a certain number of spaces do not satisfy assumption (HI): indeed, the usual norm on the space C([0, 1]) satisfies
lim sup h~0
IIx + hll + IIx - h l l - 211xll
= 2
Ilhll
for every x 6 C ([0, 1]). Therefore C ([0, 1]) does not satisfy (HI). The same remark applies to the space g l (N) and to the space L 1([0, 1]). The fact that C ([0, 1]), g 1(N) and L 1([0, 1]) do not satisfy (H 1) also follows from Proposition 3.3, since these spaces are separable with non separable duals. D
Perturbed minimization principles and applications
411
PROBLEM 1. Is the converse of Theorem 3.4 true? The answer is yes for separable spaces, since the class of separable Asplund spaces and the class of separable Banach spaces satisfying (H 1) both coincide with the class of spaces with separable dual. Thus, any counterexample should be a non separable Banach space. Possible candidates for a counterexample is C ( K ) , where K is the compact space constructed by Kunen, or the long James space J(col). Here is the "Gfiteaux" counterpart of Theorem 3.4 which can be proved in a similar way. THEOREM 3.5. A s s u m e that X satisfies (H1G). Then every convex and continuous function ~p" X ---> R is G~teaux-differentiable on a dense set. Let X -- s (N) and ~0"X ---> R be the convex continuous function defined by ~o(x) -limsup ]xn]. Since for every h 6 X, ]]hl]- 1 and for every x 6 X, we have lim sup
~o(x + th) + ~p(x - th) - 2~0(x)
t-~O
= 2,
t
the function ~0 is nowhere Gfiteaux-differentiable. Therefore there exists no Lipschitz continuous bump function b" X --+ R which is Gfiteaux-differentiable at every point of s REMARK 3.3. Let X be a Banach space and let A1 be the Banach space of all bounded Lipschitz functions on X (equipped with the norm given in Remark 2.1). Recently, Bachir [4] proved that the supremum norm on the Banach space ~41 is generically Fr6chetdifferentiable on ,AI. His proof relies on the general perturbed minimization principle and on a duality result involving a new notion of conjugacy. THEOREM 3.6. A s s u m e that there exists a bump b" X --+ R which is differentiable at every point, Lipschitz continuous and such that, f o r some p ~ (1,2], sup lim sup x6X
b ( x + h) + b ( x - h) - 2b(x)
< +ec.
Ilhll p
h-+O
Then, f o r every convex and continuous function ~p" X ~ o f X such that f o r every x ~ D, ~p is differentiable and
lim sup h~O
IIx -4- hll-4-IIx - h l l -
211xll
R, there exists a dense subset G
< +o~.
IIhlIP
A consequence of this result is that on s 1 ~< q < 2, there is no bump function which is differentiable at every point, Lipschitz continuous and such that, for some p E (q, 2], sup lim sup x~X
h--+O
b ( x + h) + b ( x - h) - 2b(x)
< +oc.
]]h]lP
The following theorem follows from Theorem 3.6 with p -- 1.
R. Deville and N. Ghoussoub
412
THEOREM 3.7. Let X be a Banach space satisfying (H2) and let qg: X --+ R be a convex
continuous function. Then the set of points xo ~ X such that q9 is differentiable at xo and such that, for h small enough, Iqg(x0 + h) - qg(x0) - qg'(x0).hl ~< Cllhl] 2, is dense. In other words, a concave continuous function defined on a space satisfying (H2) has a non empty proximal subgradient on a dense set. For a study of proximal subgradients in Hilbert spaces, we refer to [ 15], and to [ 16] for a more extensive study with application to optimal control. PROBLEM 2. Does Theorem 3.7 remain true if the assumption "X is a Banach space satisfying (H2)" is replaced by the weaker hypothesis "there exists on X a C2-smooth bump function"? In particular, since there exists a C ~ - s m o o t h bump function on c0(N), is it true that a concave continuous function defined on co (N) has a non-empty proximal subgradient on a dense set? We conclude this section with an important open problem. Recall that Alexandroff's theorem states that a convex continuous function in R n has second order expansions almost everywhere, i.e., for almost every x ~ R n, there exists (p, Q) ~ R n • S(n) such that f (x + h) = f (x) + (p, h) + Q(h, h) + o(llhl12). PROBLEM 3. Does Alexandroff's theorem extend to an infinite dimensional setting? More precisely, if H is a separable Hilbert space and if f : H ~ R is a convex continuous function, does there exist (a dense set of) points x 6 H such that there exists (p, Q) 6 H x S(H) satisfying f (x + th) = f (x) + t.(p, h) + t2.Q(h, h) + o(t 2) for every h 6 H. Note that the convex continuous function f ( x ) = I[x+l[ 2 on s shows that we cannot hope to have f (x + h) = f (x) + (p, h) + Q(h, h) + o(llhll2), and the function defined by the same formula on t~2(F), with F uncountable, shows that it is necessary to assume that H is separable.
3.4. The existence of separating polynomials Let X be a real Banach space, let p >t 1 be a real number and let f be a real valued function defined on X. We say that f has a Taylor expansion of order p at the point x E X, if there is a polynomial P of degree at most n, where n = [p] is the integer part of p, verifying
[f(x + h ) -
f(x)-
P(h)[-o(IihliP).
We say that f is TP-smooth if it has a Taylor expansion of order p at every point. Note that if f is m-times Fr6chet differentiable on X, then from Taylor's theorem we have that f is T P-smooth for 1 ~< p ~< m. We say that a polynomial P on X is a separating polynomial if P (0) = 0 and P (x) >~ 1 for all x in the unit sphere of X. It is known (and easy to prove) that if X admits a separating polynomial, then X admits a C ~ - s m o o t h bump function.
413
Perturbed minimization principles and applications
Recall that the norm II. II on a Banach space X has modulus o f convexity o f p o w e r type p if there exists a constant C > 0 such that for each e E [0, 2], 6(e) "-- inf{ 1 -
x+y 2
9 x, y ~ x
Ilxll ~ 1; Ilyll ~ 1; IIx
- y ll/> e} ~
cE p
The following result is due to Deville et al. [30]. We include its proof since it uses the linear minimization principle of Bourgain-Stegall. THEOREM 3.8. Let p >~ 1 be a real number and let X be a Banach space. Assume: (1) There exists on X a T P-smooth bump function. (2) The norm o f X is uniformly convex with modulus o f convexity o f p o w e r type p. There exists then on X a separating polynomial o f degree <, [p]. In particular, there is a C~176 bump function on X. The following corollary improves a result of Bonic and Frampton [7]. COROLLARY 3.2. Let p >~ 1 be a real number which is not an even integer. Then there is no T P-smooth bump function on L p as long as it is infinite dimensional. PROOF OF COROLLARY 3.3. Assume there exists a TP-smooth bump on Lp. Since the modulus of convexity of the norm in L p is of power type p, Theorem 3.8 shows that there is a separating polynomial P on L p of degree less or equal than p. The polynomial Q defined by Q (x) = P (x) + P ( - x ) is separating, even and of degree less or equal than the degree of P. Since p is not an even integer, the degree of Q is strictly less than p. This contradicts the fact that there exists no separating polynomial of degree < p on LP. [3
For the proof of Theorem 3.8, we shall use the following elementary lemmas from Fabian et al. [37]" LEMMA 3.1. Let 6(e) be the modulus o f convexity o f the norm I1.11of X. Let x , h ~ X and f ~ X* such that f ( x ) - ]]xl[ - ]if I ] - 1, f ( h ) - - 0 a n d e <, ]]hl] ~< 2. Then"
IIx +hll~>l +6(2 ) . Consequently, if the norm has a modulus o f convexity o f p o w e r type p and if x, h E X, f E X* are such that f ( x ) - ]]xl] -J: 0, ]if I] - 1, f ( h ) - 0 and ]thai ~< 2]]xl], then:
IIx -+-hll- Ilxll/> Cllxlll-pllhll p. LEMMA 3.2. Let k ~ 2 be an integer and let F be a finite codimensional subspace o f a Banach space X. Assume that P is a polynomial on X o f degree <, k which is a separating polynomial on F. Then there is a separating polynomial o f degree <~ k on X.
R. Deville and N. Ghoussoub
414
PROOF OF THEOREM 3.8. Let b be a bump function on X with Taylor expansion of order p at any point and such that b(0) = 0. Let qg:X ~ R U {+cx~} be the function defined by: qg(x) = 1/b(x) 2 if b(x) ~ 0 and qg(x) = +cx~ otherwise. The function q0 - I1.11is lower semicontinuous, bounded below and identically equal to + ~ outside a bounded set. On the other hand, since X is uniformly convex, it has the R a d o n - N i k o d y m property. According to the Bourgain-Stegall minimization principle, there exists g E X* (actually a dense G~ in X* of such g's) such that q9 - II. II + g attains its minimum at some point x. So for every hEX: qo(x + h) - I I x + hll-4- g(x + h) >~~o(x) - I l x l l + g(x). Consequently"
qo(x + h ) - ~o(x) + g(h) ~ IIx + h l l -
Ilxll.
Next x r 0 because qg(0) = + ~ . Let f 6 X* such that Ilfll = 1 and f ( x ) = Ilxll. Using L e m m a 3.1, we get for every h 6 K e r ( f ) n Ker(g)"
q0(x + h ) - qg(x)/> IIx + h l l - Ilxll ~ Cllxll~-Pllhll p. If we denote C(x) -
CIIxll l-p,
qg(x + h) - qg(x) ~
we thus have:
C(x)llhll p,
We now use the fact that q9 has a Taylor expansion of order p at x: there exists a polynomial P of degree ~< [p] and a function R such that qg(x + h) - qg(x) = P(h) + R(h), P(O) = 0 and limh~0 R(h)/llhll p = O. We fix e > 0 such that R(h)l ~< (C(x)/2)llhll p whenever h E X, IIh II -- e. Therefore, if h E K e r ( f ) n Ker(g) and IIh II = ~, then:
C(x) P(h) ~> 2 IlhllP" 2 The polynomial Q defined by Q(h) = C(x)cp P(eh) is a separating polynomial of degree ~< [p] on K e r ( f ) N Ker(g). By L e m m a 3.2, there is a separating polynomial of degree ~< [p] on X. [2
4. Non-smooth calculus in Banach spaces The aim of this section is to show how smooth minimization principles allow to define a subdifferential calculus for lower semicontinuous functions in smooth Banach spaces. We start by defining a generalized gradient. Note that this is well-known for convex functions qg:X --+ R U {+cx~} where the definition for any x ~ D ( f ) reads as follows: the subdifferential of q9 at x is the set:
D-q~(x) -- {p E X*; ~o - p has a local minimum at x}.
Perturbed minimization principles and applications
415
However, non-convex functions often appear in many problems from non-linear partial differential equations (as we shall illustrate in the last section) and a non-smooth calculus for semi-continuous functions is needed. Several notions of generalized gradients can be (and has been) defined in a natural way. In this paper, we define notions of subdifferentials and superdifferentials that are directly motivated by the smooth minimization principles. As shown in Section 6, these will play a key role in the resolution of Hamilton-Jacobi equations.
4.1. First order sub- and super-differentials for continuous functions Let X be a Banach space and q9 : X --~ R U { + ~ } be an arbitrary function. If x E D(qg), we define:
D - q g ( x ) - - {g'(x); g ' X --+ R i s C 1 and q9 - g has a local minimum at x}. It is not difficult to check that whenever ~p is convex, this definition coincides with the definition in the sense of convex analysis given above. We say that 99 is subdifferentiable at x if D-qg(x) ~ 0 (for x ~ D(qg), we define D-qg(x) = 0). We can define in a similar way the superdifferential of q9 at x:
D+~p(x) - {g'(x); g ' X --+ R i s C c~ a n d q g - g has a local m a x i m u m at x}. It is elementary to check that the sets D-qg(x) and D+qg(x) are convex and norm-closed, and that, if D-~o(x) and D +qg(x) are both non-empty, then q9 is Fr6chet-differentiable at x, and in this case D-~o(x) = D +qg(x) = {qg'(x) }. We now relate subdifferentials and superdifferentials with differentiability. The following result is proved in [24]. PROPOSITION 4.1 (Characterization of subdifferentials). Assume that X is a Banach space with a C l, Lipschitz continuous bump function b. Let ~p : X --+ R be an arbitrary function, x E X and p ~ X*. The following assertions are equivalent: (1) p 6 D-qg(x); (2) liminfllhll~0((qg(x + h) - qg(x) - (p, h))/[lhll) ~ O. PROOF (Sketch). (1) implies (2) is easy. Conversely, modulo replacing r by the function h -+ sup{cp(x + h) - ~ p ( x ) - (p, h ) , - 1 } we can assume, without loss of generality, that x = 0, p = 0, ~p(x) = 0 and ~0 is bounded below on X. Under these assumptions, condition (2) becomes: liminfjihll_~o(~o(h)/llhll ) >~ O. Denote by p, pl, P2 : R + --> R the functions defined by p(t) - inf{~0(h); Ilhll ~< t}, p~ (t) = f2t p(s) ds and p2(t) - f2t Pl (s) ds. The function P2 is non increasing, C l - s m o o t h and satisfies limt-+o p 2 ( t ) / t - - 0 . If we define g : X --> R by g(x) = p2(d(x)), where d is the function from Proposition 3.2, we get that g is C 1-smooth and that ~0 - g has a minimum (equal to 0) at 0. D From this, we deduce:
416
R. Deville and N. Ghoussoub
PROPOSITION 4.2 (Criterium for differentiability). I f X is a Banach space which admits a C 1 and Lipschitz continuous bump function b and if qg"X --+ R is an arbitrary function. Then q9 is differentiable at a point x if and only if D-qg(x) ~ 0 and D+qg(x) ~ 0. In this case, D-qg(x) = D +~o(x) = {~p'(x) }. It follows from the above criterium that in a space satisfying (HI), a convex continuous function is differentiable at some point x if and only if D +~0(x) ~ 0.
We shall see that it is possible to define a calculus involving sub- and super-differentials. A first difficulty is that even in finite dimensions, the sets D-qg(x) and D+~o(x) can be empty at many points. In fact, the sets D - = {x e X; ~p is subdifferentiable at x} and D + -- {x E X; q9 is superdifferentiable at x } need not be residual in X as shown by the following simple example: REMARK 4.1. If ~0:X --+ R is a continuous nowhere differentiable function on X (such functions do exist even when X -- R), then D - n D + = 0. In infinite dimensions, a Lipschitz function can be nowhere subdifferentiable and nowhere superdifferentiable: take for instance the function ~0:~ 1(N) --+ R defined by
n even
n odd
However, the smooth perturbed minimization principle yields that under some geometrical assumptions on the space X, lower semi-continuous (respectively upper semicontinuous) functions on X are subdifferentiable (respectively superdifferentiable) on a dense set. A key geometrical assumption on X is of course the existence of a smooth bump function from X. We shall start with the following result, originally due to Borwein and Zhu under the assumption of the existence of an equivalent Frechet-differentiable norm. In its present form, it is due to Bachir [5]. THEOREM 4.1 (Minimization of the sum of two functions). Let X be a Banach space satisfying (HI). Let u l, u2 :X --+ R U {+cx~} be two lower semi continuous and bounded below functions such that limn~ 0 inf{ul (Xl) + u2(x2); Ilxl - x211 < n} < + ~ . Then, f o r each e > O, there exist xl, x2 E X, Pl E D - u l (Xl) and P2 e D - u 2 ( x 2 ) such that: (1) Ilxl -x211.(1 + IlPl II + liP211) < e, (2) IIel + P2 II < e, and (3) Ul(Xl) + u2(x2) < limo~o inf{ul(y) + u2(z); IlY - zll < rl} + ~. PROOF (Sketch). Let d be the Lipschitz continuous function given by Proposition 3.2. For each a > 0, define Wa :X • X --+ R by Wa(X, y) = Ul (X) + ue(y) + ad2(x - y). Wa is a lower semi continuous and bounded below function on the Banach space X x X and B(x, y) = b ( x ) b ( y ) is a Lipschitz continuous, cm-smooth bump function on X • X.
Perturbed minimization principles and applications
417
A c c o r d i n g to the s m o o t h m i n i m i z a t i o n principle, there exists a Lipschitz continuous, C 1s m o o t h function g" X x X --+ R such that: (a) wa 4- g has a strong m i n i m u m at s o m e point (Xa, ya) 9 X x X , (b) Ilgll~ - - s u p { l g ( x , Y)I; (x, y) 9 X x X} < l / a , (c) sup{ll ~Og (X , y)][ ~ " (X , y ) 9 X x X} < 1 / a and 0g (X, y)[] ~ ; (x, y) 9 X x X} < 1 / a (d) sup{ II~S
[]
Og D e n o t e Xl -- Xa, X2 -- Ya, Pl = - a ( d 2 ( x - Y))' -~(Xa, Ya) and p2 - a ( d 2 ( x - y ) ) ' o__~g ay (Xa, Ya). For a large enough, properties (1), (2) and (3) are satisfied. COROLLARY 4.1. Let X be a B a n a c h space satisfying ( H I ) a n d let ~p : X --+ R be a lower s e m i - c o n t i n u o u s a n d b o u n d e d below f u n c t i o n such that D(qg) ~ 0. Fix e, )~ > O, then, f o r every xo 9 X satisfying qg(xo) < infx q9 + eL, there exist Xl 9 X a n d Pl 9 D - q g ( X l ) such that IlPl Ilg* < ~ a n d []Xl - xoll <
e.
PROOF. C h o o s e )~o 9 (0,)~) such that e)~o - (qg(xo) - infx qg) > 0. Define h ( x ) - ~.o(llx xoll - e) and let 0 < a < e)~o - (qg(xo) - infx qg). C h o o s e eo 9 (0, e) in such a way that, if h ( x ) < qg(xo) - e)~o + a - infx qg, then Ilx - xoll < eo. A p p l y i n g T h e o r e m 4.1 with U l - q9 and u2 = h, there exists Xl, x2 9 X and there exist p l 9 D - q g ( X l ) , P2 9 D - h ( x 2 ) such that: (a) Ilxl - x2 II < e - eo, (b) IIP~ + P2 II < Z - Zo, and (c) qg(Xl) + h(x2) < l i m o ~ o inf{qg(x) + h ( y ) ; IIx - yll < r/} + a. Since h is Lipschitz c o n t i n u o u s with Lipschitz constant )~o, we have liP211 ~< )~o. B y (b), we obtain that IlPl II < ~. Since q9 is lower s e m i - c o n t i n u o u s a n d h is Lipschitz continuous, we have
lim inf{qg(x) 4- h ( y ) ; 0----~0
IIx - yll < ~} -
inf {qg(x) 4- h ( x ) ; x 9 X}
<<, ~o(xo) + h(xo). Therefore, h(x2) < qg(xo) 4- h(xo) 4- a - ~O(Xl) <<,99(x0) - e)~o 4- a - infx qg. Thus, ]Ix2 xo II < eo. Together with (a), this implies that IlXl - xo 11 < e. []
COROLLARY 4.2. A s s u m e that X satisfies (HI), a n d that qg" X --+ R is lower semi contin-
uous. Then q9 is sub-differentiable on a dense set.
PROOF. Fix x0 9 X and e > 0. A p p l y i n g T h e o r e m 4.1 with U l = q9 and u2 the function defined by uz(x0) = 0 and u z ( x ) = +cx~ if x r x0, we obtain points X I , X 2 9 X such that IlXl - x21l < e and vectors p l 9 D - q g ( X l ) and p2 9 D - u z ( x 2 ) . Necessarily, x2 = xo. So q9 is subdifferentiable at a point X l such that IIx~ - xo II < e. E3
418
R. Deville and N. Ghoussoub
4.2. The sub-differential o f the sum Formulae for the subdifferential of the sum of two functions defined on a Banach space have been investigated by various authors (Ioffe [55], Fabian [34,35], Deville and E1 Haddad [27]). We present here an extension, observed by Bachir [5], of a result obtained by Borwein and Zhu [9] in the context of Banach spaces admitting an equivalent Fr6chetdifferentiable norm. We need the following: DEFINITION 4.1. Let u l, u 2 : X ~ R W {+cxz} be lower semicontinuous functions. We say that the pair (u l, u2) is locally uniformly lower semicontinuous if, for every x 6 X and every uniformly continuous function qg:X • X ~ R, there exists r > 0 such that, inf ul (y) + u2(y) -+- ~o(y, y) yEB
= liminf{ul(y)+u2(z)+qg(y,z); 77--+0
I l y - z l l < rl, y , z ~ B}.
where B = B x (x , r). This condition is clearly stable if we perturb u l and U2 by uniformly continuous functions, and is always satisfied when dim(X) < +cx~, or when one of the functions is locally uniformly continuous. THEOREM 4.2. Let X be a Banach space satisfying (H1). Let ul, U2 : X ~ R U {+cx~} be lower semicontinuous functions. Assume that the pair (Ul, u2) is locally uniformly lower semicontinuous. Suppose that xo E X and p E D - (u 1 + UZ)(X0) are given. Then, f o r each e > O, there exist xl, x2 E X, Pl E D - u l (Xl) and P2 E D - u z ( x 2 ) such that: (1) Ilxl - xoll < e, Ilx2 - xoll < e and Ilxl - xzll.(llpl II -+- [Ipzll) < e, (2) lUl (Xl) - Ul (xo)l < e and lu2(x2) - u2(xo)l < e, and (3) IlPl + p2 - Pll < e. T h e o r e m 4.3 is a non trivial consequence of Theorem 4.1 on the minimization of the sum of two functions. A Rademacher-Preiss type theorem for uniformly continuous functions in spaces which admit a smooth bump function can be deduced from this formula. Let us recall that according to Rademacher theorem, every Lipschitz continuous function in R n is differentiable almost everywhere. Preiss [62] has extended this result to an infinite dimensional setting. He proved in particular that if X is an Asplund space (in particular if X satisfies (H1)), then every locally Lipschitz continuous real valued function defined on X is differentiable on a dense subset of X. The following result from [27] can be seen as a weak form of Preiss theorem for uniformly continuous functions. COROLLARY 4.3. Let X be a Banach space satisfying (HI). Let u be a uniformly continuous function defined on X. Then f o r every x E X and every e > O, there exist x l, x2 E X, there exist p - ~ D - u ( x l ) and p+ ~ D+u(x2) such that: (1) Ilxl - x II < e and [Ix2 - x II < e, (2) lU(Xl) - u(x)l < e and ]u(x2) - u(x)] < e, (3) liP- - P+II < e.
Perturbed minimization principles and applications
419
In order to prove Corollary 4.3, it is enough to apply Theorem 4.2 with U l ~ - U and u2 = - u , and to observe that D - ( - u ) ( x 2 ) -- D+u(x2). Let us here stress the fact that Preiss' result is considerably harder to prove. PROBLEM 4. If the function u of Corollary 4.3 is nowhere differentiable, then the points Xl and x2 are necessarily different. It is unknown whether one can take p - -- p+ when u is an arbitrary uniformly continuous function (the answer to this question is positive if d i m X = 1). PROBLEM 5. Counterexamples show that the assumption "(u l, u 2 ) is locally uniformly lower semicontinuous" cannot be dropped in Theorem 4.1 (see for instance [29]). These counterexamples involve extended valued functions. It is an open problem to know whether a formula for the subdifferential of the sum of two lower semi continuous, real valued functions in infinite dimensions always holds true. REMARK 4.2. Theorem 4.2 will be applied in the next section in the proof of uniqueness of the viscosity solution of a particular Hamilton-Jacobi equation. The conclusion "llxl x211.(llpl II + lip211) < e" in Theorem 4.2 is usually needed in the case of Hamilton-Jacobi equations coming from optimal control or differential games. Most results can be established for weaker types of differentiability. Indeed, let u : X R tO {+cx~} be an arbitrary function. If x 6 X, we may define for example the G-viscosity subdifferential of u at x: DGU(X) -- {gt(x); g ' X --+ R is G~teaux-diff. and u - g has a local min. at x}.
The G-viscosity superdifferential of u at x is defined similarly. The following result is the analogue of Theorem 4.2. 4.3. Let X be a Banach space on which there exists a Lipschitz continuous, GCtteaux-differentiable bump function b on X such that b t is norm to weak* continuous. Let u l, u2 be two real valued functions on X such that u l is lower semi continuous and u2 is uniformly continuous. Suppose that xo E X and p E DG(Ul + uz)(x0) are given. Then, f o r PROPOSITION
each e > 0 and each weak* neighbourhood V o f p, there exist xl, x2 E X, Pl E DGUl(Xl) and P2 E DGU2(X2 ) such that: (1) [[Xl - x0 [[ < e and [[x2 -- X0 [[ < 6, (2) [ul(xl) - ul(xo)[ < e and [u2(x2) - u2(xo)[ < e, (3) P l + p 2 E V .
4.3. Second order differentials We now investigate analogous results for second order sub-differentials defined as follows: Let X be a Banach space and qg:X --+ R U { + ~ } . The Second order subdifferential and
420
R. Deville and N. Ghoussoub
the Second order superdifferential of q9 at x, are the sets: De'-~p(x) = { (g'(x), g" (x)); g ' X ---> R is C 2 and ~o - g has a local minimum at x } and De'+99(x ) = { (g'(x), g ' ( x ) ) ; g ' X --> R is C 2 and ~p - g has a local maximum a t x } . The second order smooth minimization principle tells us that whenever X is a Banach space which admits a C2-smooth bump function with Lipschitz derivative, then for every s function ~0"X ~ R, D2'-~p(x) r 0 for every x in a dense subset of X. On the other hand, it is possible that both D2'-~o(x) and D2'+~0(x) are non empty, and yet q9 is not twice differentiable at x" take ~o'R --+ R defined by: ~0(x) " - x 2 s i n ( I / x ) if x ~ 0 and f ( 0 ) - 0. Although a second order minimization principle is available in infinite dimensions, not much more is known in this context and our goal in this section is to present some results that are valid in finite dimensions, hoping to motivate the problem of their extensions to an infinite dimensional setting. For example, we present a formula for the second order subdifferential of the sum of two lower semicontinuous functions that is available only in finite dimensions (see [28]). The main ingredient of the proof is the following theorem due to Jensen who established uniqueness of viscosity solutions of Hamilton-Jacobi equations of order two [56]. THEOREM 4.3 (Jensen) (Second order information in a minimization problem). Let u" R n -+ R be a function such that x --+ u ( x ) - ~llxll 2 is concave (such a f u n c t i o n u is called semi-concave). A s s u m e that u has a strong m i n i m u m at xo. Then, f o r every e > O, there exists x ~ R n with IIx - x011 < e and (p, Q) ~ R n • S(n) with Ilpll < E a n d 0 <~ Q <. 2)~I such that u ( x + h) -- u ( x ) 4- (p, h) + Q(h, h) 4- o(llhl12).
Note that Alexandroff's theorem (See Problem 3) applies to a semi-concave function so that it has order expansions almost everywhere, that is, there exist (a dense set of) points x E R n such that there exists (p, Q) 6 H • S(R n) satisfying f ( x + t h ) - f ( x ) + t . ( p , h ) + t 2 . Q(h, h) + o(t 2) for every h E R n . Jensen's theorem gives more information on the value of p and Q. The idea of the proof is the following: Fix r > 0 and consider u the greatest convex function on Br dominated by u (u is called the convex lower envelope of u on Br). Jensen proved that if u is semiconcave, then the Lebesgue measure of the set A - {x ~ Br ; u ( x ) = u (x)} is positive (see [18]). Therefore, by Alexandroff's theorem, there exists x 6 A and (p, Q) ~ R n x S(n) such that u ( x + h) - u ( x ) + (p, h) + Q(h, h) + o(llhl12). Because u ( y ) >~ u ( x ) for all y and u ( x ) = u (x), we have Q ~> 0. And because ofthe semi-concavity
of u, we also have Q ~< 2)~I. The conditions IIx - x0ll < e, Ilpll < e can be obtained by choosing r > 0 small enough.
Perturbed minimization principles and applications
421
PROBLEM 6. Is it possible to extend Jensen result mentioned above to an infinite dimensional setting? To our knowledge, the work [ 18] gives the best information in this direction. As in the first order sub-differentials, we have the following two results which are only established in a finite dimensional setting. THEOREM 4.4. Let u l, u2 be two real valued, bounded below, lower semicontinuous functions on R n. Then, f o r every e > O, there exist x 1, x2 E R n , (Pl, Q l) E D 2'- (u 1) (Xl) and (p2, Q2) E D 2'- (uz)(x2) such that: Ilxl - x211 < ~,
Ilpl q- p2ll < ~
and
II Ol Jr- Q21I < ~.
THEOREM 4.5 (Deville-E1 Haddad). Let u l, u2 be two real valued lower semicontinuous functions on R n. Suppose that xo and (p, Q) E D Z ' - ( U l + uz)(x0) are given. Then, f o r every ~ > O, there exist Xl,X2 E R n, (Pl, Q1) E D2'-Ul(Xl) and (P2, Q2) E o Z ' - u z ( x 2 ) such that: (i) ]lxl - xo II < e and Ilx2 - xo l[ < e, (ii) lul (xl) - ul (xo)] < e and ]u2(x2) - u2(xo)l < e, (iii) IlPl + p2 - Pll < e and II Q1 + Q2 - QII < e. Unfortunately, this result does not extend to infinite dimensional Banach spaces. For instance, let X = g2(N). For x = (xn) E g2(N), denote x + = sup{x, 0} and x - = - inf{x, 0}. If u l ( x ) = [Ix+ll 2 and uz(x) = I[x-II 2, then (Ul + uz)(x) = I[x[]2. So ( 0 , 2 I ) E DZ'+(Ul + u2)(0). On the other hand, for every Xl,X2 E g2(N) and every (pl, Q1) E D2'+(Ul)(Xl), ( P 2 , Q 2 ) E D2,+(uz)(x2), w e have [IQ1 + Q 2 - 2111 >~ 2. We just give a sketch of the proof of T h e o r e m 4.5 so as to motivate the need for the extension of Alexandroff's and Jensen's theorems to infinite dimensions. For more details, we refer to [28]. As in the first order case, one first reduce to the case (p, Q) = (0, 0) and u 1 -Jr- U2 has a m i n i m u m at xo. We then minimize the function equal to Ul (x) + u2(y) -+- Mllx - yl[2 in a n e i g h b o u r h o o d of (xo, xo) and equal to +cx~ outside of this neighbourhood. This function attains its m i n i m u m at some point (x l, x2). We get then all the claims but the condition II Q1 + Q2 - Q II < ~. In order to overcome this difficulty, the first step is to regularize by using Inf convolutions: one introduces the functions Ui,Z (X) = infzERn {Ul (Z) + 89IIz - x II2 } for i = 1, 2. The functions u l,z and u2,z are difference of convex functions and are approximations (as)~ --+ cx~) of u 1 and u2. The idea is then to prove T h e o r e m 4.5 for u l,z and u2,z and then to pass to the limit. The key ingredient for proving T h e o r e m 4.5 for u 1,z and u2,z is the result of Jensen mentioned above (see [ 18] for details). The following corollary of T h e o r e m 4.5 can be seen as a weak version of Alexandrov's theorem for continuous functions. C O R O L L A R Y 4.4. Let u be a continuous function defined on R n. Then f o r every x E R n and f o r every e > 0 there exist x l , x 2 E R n, ( p - , Q - ) E D Z ' - u ( x l ) and (p+, Q+) E oZ'+u(x2) such that: (i) IlXl - x II < e and ][x2 - x [I < e, (ii) ]u(xl) - u(x)[ < e and [u(x2) - u(x)l < e, (iii) liP- - P+ II < ~ and II Q - - Q+ II < e.
422
R. Deville and N. Ghoussoub
4.4. Stability of subdifferentials and superdifferentials The following stability result is from [22] where one can find a proof and variants of this result, as well as references on other stability results. We only state results for superdifferentials, since the case of subdifferentials is similar. THEOREM 4.6. Let X be a Banach space satisfying (HI). Let q) and (q)n, n ~ N) be upper semi continuous functions from X into R. Assume that (~On) converges locally uniformly to ~p. If x E D(qg) and p E D+qg(x), then there exist (Xn) C X and Pn E D+qgn(Xn) such that (Xn) converges to x, (~O(Xn)) converges to qg(x) and (Pn) converges to p. We now present a stability result of viscosity superdifferentials for the upper semicontinuous envelope of the supremum of a family of upper semi-continuous functions. This result can be used to show that, under very mild assumptions, a supremum of viscosity subsolutions is a viscosity subsolution. This last result is a key ingredient to prove the existence of a viscosity solution for Hamilton-Jacobi equations via the Perron method (see Section 6.4). THEOREM 4.7. Let X be a Banach space satisfying (HI). Let $ be a locally uniformly bounded family of upper semi-continuous functions from X into R. Let u* be the upper semi continuous envelope of the function u(x) -- inf{g(x); g E S}. If x E X and p E D+u*(x), then there exist (Xn) C X, (q)n) a sequence of functions in S and Pn E D+q)n(Xn) such that (Xn) converges to x, (q)(Xn)) converges to u*(x) and (Pn) converges to p. Under assumption (H1G) (respectively (H2)), similar stability results hold for the G~teaux superdifferential (respectively second order superdifferential).
4.5. Mean value theorems We now turn to mean value theorems. There are many! The first and the last theorems presented below are valid for nonsmooth functions defined in (HI) spaces. The first is from [22], the second is from [2], while the third can be found in E1Haddad's thesis. THEOREM 4.8. Let X be a Banach space satisfying (HI), b / a n open convex subset of X and q) : bl --+ R be lower semicontinuous. Assume that there exists a constant K > 0 such that for all x E bl and for all p E D-qg(x), IlPll <~ K. Then ~o is Lipschitz continuous, with Lipschitz constant K. The following result holds in arbitrary Banach space. Observe that below, we only refor one p E D-q)(x) (instead of all p E D-q)(x)).
quire IlPll ~< M
THEOREM 4.9 (Mean value theorem for subdifferentiable functions). Let X be a Banach space, bl an open convex subset of X and q) : bl --+ R a continuous G~teaux subdifferentiable function. Suppose that there exists M >/0 such that for every x E lg, there exists p E D-qg(x) with IlPll <<,M. Then q) is Lipschitz continuous, with Lipschitz constant K.
Perturbed minimization principles and applications
423
The following mean value inequality is a variant of the mean value inequality of Clarke and Ledyaev (see [ 13,14]). It is obtained formally from the mean value inequality of Clarke and Ledyaev by considering, in the Banach space E = X • R, the point (to, x0) and the (unbounded) convex C = {0} x X. It is useful to prove uniqueness results for viscosity solutions of parabolic Hamilton-Jacobi equations. COROLLARY 4.5 (A mean value inequality). Let X be a Banach space satisfying (HI) and let ~p: [0, T) • X --+ R be lower semicontinuous and bounded below. Assume that f o r all x E X, ~p(O,x) ~ 0 and fix (to, xo) E (0, T) • X. Then f o r every e > O, there exist (t, x) E (0, T) • X and (a p) E D-~o(t x) such that a < ~(to,xo) + e and Ilpll < e. to '
'
5. Application to critical point theory Many problems of Mathematics and Physics amount to the search of a critical point of a differentiable map q9 on a Banach space X: that is points x E X such that qg'(x) = 0. The simplest way of finding critical points is to consider minimization problems. Our first result gives the existence and the localisation of an "almost critical point" and is an immediate application of Ekeland's theorem.
5.1. Palais-Smale minimizing sequences PROPOSITION 5.1. Let X be a Banach space and ~p : X --+ R be a differentiable mapping such that ~p is bounded below. Let F be a closed subset o f X such that infx 99 -- infF q9 = c. Then there exists a sequence (Xn) C X such that limqg(Xn) = c, lim II~0'(xn)ll- 0 and lim dist(xn, F) = 0.
The proof follows from Ekeland's theorem applied to any minimizing sequence (which can be chosen in F) and the condition qg(x) - qg(xn) ~> - 1 / n l l x - xnll which implies that II~o'(Xn)llx* <~ 1/n. In order to get a critical point, we need to insure that this sequence of "almost critical points" located around F, is convergent, hence we are led to introduce the Palais-Smale around F, at altitude c ((PS)F,c), which simply means that any sequence satisfying the claim of the above proposition, is relatively compact. Whenever F -- X, we shall write (PS)c instead of (PS)x,c. With this condition, we clearly obtain an x E F such that qg(x) = c and qg'(x) = 0. If we assume some smoothness on the domain X, then the first order smooth perturbed minimization principle allows to extend the above result to functions that are merely lower semi-continuous: PROPOSITION 5.2. Let X be a Banach space satisfying (HI) and let ~p : X --+ R be a brunded below, lower semi continuous function. Let F be a closed subset o f X such that infx q9 = infF qg. Then there exist a sequence (Xn) C X and Pn E D-~p(xn) such that: limqg(Xn) = inf~p, x
lim Ilp~ Ils* = 0
and
lim dist(xn, F) = 0.
424
R. Deville and N. Ghoussoub
The next result gives second order information. The proof is similar to the above, but uses the second order smooth minimization principle. PROPOSITION 5.3. Let X be a Banach space satisfying (H2) and let r : X --+ R be a bounded below function which is twice differentiable at every point. Then f o r every e > 0 and any ye ~ X such that qg(ye) <~infq9 + e 2, there exists xe ~ X such that: (1) (2) (3) (4)
Ilx~ - Y~II ~< e, qg(xe) ~< infq9 + e, II~0'(x~)IIx, ~< E, and ~p"(x~)(h,h) ~ -ellhll2 f o r a l l h ~ S.
PROBLEM 7. In view of Proposition 5.1 which is valid in any Banach space, the following question is quite natural: does Proposition 5.3 remain true without the assumption "X satisfies (H2)"?
5.2. The mountain pass theorem The assumption "q9 is bounded below" can be replaced by an assumption about the existence of a mountain range separating two points. We then obtain the mountain pass theorem of Ambrosetti and Rabinowitz. The following result is a variant of that theorem with an additional information about the location of the sequence of "almost critical points". THEOREM 5.1 (Ghoussoub-Preiss). Let X be a Banach space and let qg : X --+ R be a differentiable mapping. For a, b ~ X, consider F to be the set o f all continuous functions y : [0, 1] --+ X such that y (0) = a and y (1) = b. Denote c -- inf max ~0(y (t)). yeF
t e [ O , 1]
Let F be a closed subset of X such that a, b ~ F, F C {q9 ~ c} and F separates a and b, i.e.,for every y ~ F, there exists t ~ ]0, 1[ such that y ( t ) ~ F. Then there exists a sequence (Xn) C X such that limqg(Xn) = c, lim llqg'(Xn) I[ = 0 and limdist(xn, F) = 0. I f moreover q9 satisfies the (PS)F,c, then there exists x ~ F such that qg(x)= c and ~o' (x) = O.
PROOF (Sketch). Fix e < min{ 1, dist(a, F), dist(b, F) }, pick a continuous path g : [0, 1] --+ X such that max{qg(y(t)); t 6 [0, 1]} < c + e Z / 4 a n d d e f i n e L = {t 6 [0, 1]; dist(g(t), F) ~> e}. L is a closed subset of [0, 1] and the set F0 = {p 6 C([0, 1], X); PIE = )tiE} is a closed subset of the Banach space C ([0, 1], X) (endowed with its usual supremum norm) that is included in F . Finally, define I :L --+ R by I (p) -- max {(r + s
t E [0, 1]\L },
where g : X --+ R is given by the formula g(x) = max{0
; 62 -
e.dist(x, F) }.
Perturbed minimization principles and applications
425
The functional I is continuous, bounded below and defined on the closed subset of a Banach space. According to Ekeland's minimization principle, there exists )to 6 F0 such that: (i) I ( y o ) <<.I(y); (ii) IlYo - zll ~< e/2; (iii) I ( p ) >1 I ( y o ) - ~ l i p - zoll for all p E Fo. Denote M = {t ~ [0, 1]; (~p + g)(yo(t)) = 1(9/o)}. There exists then t E M such that: (a) c <<,~p(yo(t)) <<,c 4- 5s2/4; (b) dist(y0(t), F) ~< 3e/2; (c) II~o'(zo(t))llx, < 5e/2. Since this can be done for every e > 0, we obtain Theorem 5.1. See [50] for details. [] The localisation in the above mountain pass theorem can be used for studying the structure of the set of critical points of q9 (see [42-44,38-40] and [50]). Let us show here how it can be used to construct nice paths. COROLLARY 5.1 (Taubes). Let X, ~p, a, b E X, F a n d c be as in Theorem 5.1. A s s u m e that qg(a) < c, ~p(b) < c, a n d that q9 satisfies (PS)c. Then, f o r each s > O, there exists a continuous p a t h y :[0, 1] --+ X such that y (0) = a, y (1) = b, a n d 6 > 0 such that:
PROOF. Since q9 is uniformly continuous on compact subsets of X, it is enough to prove that for each e > 0, there exists a continuous path y :[0, 1] ~ X such that y (0) = a, y(1) = b and ~0(y(t)) = c :=~ II~o'(Z(t))llx* < e. If not, then there exists e > 0 such that the set F -- F, -- {x ~ X; Ilqg'(x)lls, ~ ~ and ~o(x) -- c} is a closed subset of X that separates a and b. By the preceding theorem, there exists x 6 F such that ~0~(x) = 0, which is a contradiction. M REMARK 5.1. There exist also non-smooth versions of the mountain pass theorem. A first result in this direction has been obtained by Chang [12]. More recently, Corvellec, de Giovanni and Marzzochi [17] and Fang [39] introduced the notion of weak slope which allows them to develop critical point theory for lower semi continuous functions.
As in minimization problems, it is possible to give second order information in the mountain pass theorem. This is done in [41,42] under some conditions of uniform continuity of the first and second order derivatives of 99. These conditions do not seem to be severe in applications. THEOREM 5.2 (Fang-Ghoussoub). Let X be a B a n a c h space a n d let qg : X --+ R be a C 2functional. Let a, b ~ X, F, c a n d F be as in Theorem 5.1. Suppose that q9t a n d r are H61der continuous on a n e i g h b o u r h o o d o f {r = c}. Then there exists a sequence (Xn) C X such that:
R. Deville and N. Ghoussoub
426
(1) limqg(Xn) = c; (2) limllqg'(xn)]l = 0 ; (3) limdist(xn, F) = O; 1 2 for every h in a subspace En of codimension at most 1. (4) ~o"(xn)(h, h) >1-~llhll
6. Applications to Hamilton-Jacobi equations in Banach spaces General first order Hamilton-Jacobi equations are of the form
H (x, u(x), Du(x)) = 0 in the stationary case, and of the form
H (t, x, u t, x), OuSt, x)) = o in the evolution case. They arise in optimal control theory and in differential games. We refer to [ 19] and [20] for a detailed discussion of these equations, and, for instance, to [6] for an introduction to the theory of Hamilton-Jacobi equations. The problem is to deal with existence and uniqueness of a global "solution" of these equations, under suitable hypothesis on the Hamiltonian H (monotonicity of H in the variable u, continuity properties of H, and in some cases, convexity properties of H in the variable Du(x)). In this section, we shall concentrate on a couple of examples that will illustrate the power of the smooth minimization principles described previously.
6.1. The maximum principle for stationary first order Hamilton-Jacobi equations Let I2 be a bounded open subset of R n and H" S2 x R n ~ R be an arbitrary continuous function ( H is called an Hamiltonian). Consider the following Hamilton-Jacobi equation: u(x) + H(x, Du(x)) = 0 u (x) -- 0
for all x 6 S2, for all x 6 OS-2.
(H J1)
We first examine the maximum principle for classical solutions. Let u ~ C ( ~ ) N C 1 (if2) be a (classical) subsolution: i.e., u(x) + H(x, Du(x)) <~0
u(x) <<,0
for all x ~ S'2, for all x ~ 0 $2,
and let v 6 C(I2) N C1 (12) be a supersolution" i.e., v(x) + H(x, Dr(x)) ~ 0
v(x) >10
for all x 6 Y2, for all x 6 O12.
Perturbed minimization principles and applications
427
Assume supse(u - v) > 0. Then there exists x0 E I2 such that u(xo) - v(xo) = sups2 (u - v) > 0. So Du(xo) = Dv(xo). Since u(xo) + H(xo, Du(xo)) <<.0 and v(xo) + H(xo, Dv(xo)) >~ O, we get by subtracting u(xo) - v(xo) <<.O. This contradiction means that for all x E S-2, u (x) ~< v (x). Assuming that u and v are bounded, a similar proof still works on unbounded domains or in infinite dimensions, but since we no longer have compactness, sups ~ (u - v) is not necessarily attained, and we therefore need to apply perturbed minimization principles. It is therefore necessary to impose more restrictive hypothesis on the Hamiltonian H (for instance H is uniformly continuous) in order to estimate H(xo, Du(xo)) - H(xo, Dv(xo)). Indeed, the following counterexample shows that some hypothesis on H is needed. The linear equation: u(x) - x(1 + IxI)u'(x) = 0 on R has 0 and x / ( 1 + Ixl) as two classical bounded solutions on R, that is the maximum principle is violated in this case. In general, classical solutions do not exist, but viscosity s o l u t i o n s - as defined b e l o w do exist under relatively general assumptions. We shall then see that, using the calculus involving sub- and superdifferentials, the above proof of the maximum principle can be adapted to viscosity subsolutions and supersolutions. Let us recall the definition of viscosity solutions for Hamilton-Jacobi equations of the form given above. DEFINITION 6.1. A function u :X --+ R is a viscosity subsolution of (HJ1) if u is upper semi continuous and, for every x 6 X and every p E D+u(x):
u(x) 4- H ( x , p) <<.O. The function u is a viscosity supersolution of (HJ1) if u is lower semi continuous and, for every x 6 X and every p E D - u (x):
u(x) § H ( x , p) ~ O. Finally, u is a viscosity solution of (HJ 1) if u is both a viscosity subsolution and a viscosity supersolution of (HJ 1). Let X be a Banach space and H :X • X* --+ R be a continuous Hamiltonian satisfying the following condition: 1 p)
-
H(x2, P)l ~ 0
as IIx~ - x211(llpll + 1) --+ 0,
IH(x, P l )
-
H ( x , P2)I--+ 0
as IlPl - P2 II ~ 0.
IH(xl,
(*)
If H is uniformly continuous, then H satisfies condition (,) which in turn implies the continuity of H. We now prove the uniqueness of a bounded continuous viscosity solution u : X -+ R of equation (HJ 1): THEOREM 6.1. Let u, v be two real valued bounded functions defined on X. Assume that the Banach space X satisfies (H1) and that the Hamiltonian H satisfies (,). If u is a 1 Thiscondition is satisfiedby the Hamilton-Jacobi equations arising in deterministic optimal control problems and in differential games.
428
R. Deville and N. Ghoussoub
viscosity subsolution of (HJ1) and if v is a viscosity supersolution of (HJ1), then u <<.v on X. PROOF. Fix e > 0. The functions v a n d - u are lower semicontinuous continuous and b o u n d e d below. According to T h e o r e m 4.1 on the minimization of the sum of two functions, for each e > 0, there exist xl, x2 E X and pl E D+u(xl), P2 E D-v(x2) satisfying: (1) (1 + Ilplll + IIp211), ]lxl - xzll ~ e, (2) V(X2) -- U(Xm) < l i m o s 0 inf{v(y) -- u(z); IlY -- zll < 0} + e, and (3) liP2 - Pl II < e. Condition (2) implies that: v(x2) - U(Xl) < inf{v(x) - u(x); x 6 X} + e. Since u is a viscosity subsolution of u + H(x, Du) = 0, we have U(Xl) + H(xm, Pm) ~< 0. Since v is a viscosity supersolution of v + H(x, Dr) = 0, we have v(x2) + H (x2, P2) ~> 0. Consequently inf(v
-
u)
>
v(x2)
-
U(Xl)
-
6
x
H(xj,
Pl)
--
H(x2,
P2)
-
e
>/ H(xl, P l ) - H(x2, Pl) -+- H(x2, P l ) - H(x2, P2) - e.
(1)
Since Ilxl - x 2 l l ( l l p l II + 1) ~ e, we obtain that H(xl, Pm) - H(x2, P l ) ~ 0 as e --+ 0. Since IIp~ - p2 II ~< e, we get that H (x2, Pl) - H (x2, P2) ~ 0 as e --+ 0. Plugging this information in (1), we get: inf(v - u) ~> 0.
[]
x
REMARK 6.1. It is important to have comparison theorems for semi-continuous functions rather than for uniformly continuous solutions for several reasons: This need appears in the existence results via Perron's method where discontinuous viscosity subsolutions are used in the proof as well as in optimal control with boundary conditions where the value function can be discontinuous.
6.2. The maximum principle for parabolic Hamilton-Jacobi equations Our aim here is to prove uniqueness of a b o u n d e d uniformly continuous viscosity solution u : R + • X --+ R of the following evolution equation:
ut + H(x, Ux) -- O, u(O, x) - uo(x), where u o : X ~ tinuous.
(HJ2)
R is the initial condition which is assumed bounded and uniformly con-
429
Perturbed minimization principles and applications
DEFINITION 6.2. Say that a function u : R + x X --+ R is a viscosity subsolution of (HJ2) if u is upper semi continuous and, for every (t, x) 6 X and every (a, p) E D + u ( t , x):
u(O, x) <<. uo(x). Likewise, we can define the viscosity supersolutions for (HJ2). THEOREM 6.2. Let u, v be two real valued uniformly continuous functions defined on R + x X. Assume that the Banach space X satisfies (H1) and that the Hamiltonian H satisfies (.). I f u is a viscosity subsolution o f ( H J 2 ) and if v is a viscosity supersolution o f (HJ2), then u <<.v on X. PROOF. Fix T E (0, §
(v-u)
inf
and, aiming for a contradiction, let us assume that <0.
[0,T)xX
Fix s > 0. The function v - u is uniformly continuous and non-negative on {0} • X, hence b o u n d e d below on [0, T) x X. Thus, there exists (to, x0) E (0, T) x X such that
(v - u)(to, xo) <
inf
(v - u) § s T
and
(v - u)(to, xo) < O.
[0,T)•
According to the mean value inequality, there exists (t, x) E (0, T) x X and (a, p) E D - (v - u) (t, x) such that a<
(v - u)(to, xo) to
+s
and
[IP[l<s.
By the formula for the subdifferential of the sum applied to U l = 1) and U 2 "-- m U , there exist ( t l , x l ) , (t2, x2) E (0, T) x X, there exist (al, P l ) 6 D - v ( t l , x l ) and (a2, P2) E D+u(t2, X2) satisfying: (1) [Ixl - x0[I < s, Ilx2 - xoll < s, Itl - to[ < s and It2 - tol < s, (2) IIpl - p2 _ pll < e and lal - a2 _ al < e, and (3) (llpl II + lall + Ilpzll + [azl).(llxl - xzl[ + Itl - tzl) < s. The function u is a viscosity subsolution of u t + H (x, Ux) : 0, so: a2 § H1 (x2, P2) ~< 0. The function v is a viscosity supersolution of vt § H (x, Vx) = 0, so:
al + H ( x l , p l ) ~ O. Consequently
inf[o,T)•
>
(v - u)(to, xo)
mS>
T > a-2s>al-az-3s/>H(xz,
(v - u)(to, xo)
m s
to pz)-H(xl,pl)-3s.
R. Deville and N. Ghoussoub
430
Observe that (3) implies that (llplll 4- [Ip2ll).llxl - x211 < e. In the same way as in the stationary case, using the fact that H satisfies ( . ) , and sending e to zero, we get: inf[0,r)xX(V - u) ~> 0,
which is a contradiction.
6.3. Uniqueness in second order Hamilton-Jacobi equations We conclude by explaining how to use T h e o r e m 4.5 to prove uniqueness of viscosity solution for second order Hamilton-Jacobi equations. Our model is the following stationary equation: u +
(HJ3)
where H is a uniformly continuous Hamiltonian. DEFINITION 6.3. A function u : R n --+ R is a viscosity subsolution of (HJ3) if u is upper semi continuous and, for every x 6 R n and every [p, Q) ~ D2'+u(x):
u(x) 4- H ( x , p, Q) <. O. The function u is a viscosity supersolution of (HJ3) if u is lower semi continuous and, for every x 6 R n and every (p, Q) E DZ'-u(x):
u(x) 4- H ( x , p, Q) >~O. Finally, u is a viscosity solution of (HJ3) if u is both a viscosity subsolution and a viscosity supersolution of (HJ3). THEOREM 6.3. Let u, v be two bounded real valued functions defined on R n. If u is a viscosity subsolution o f ( H J 3 ) and v is a viscosity supersolution of(HJ3), then u <. v. PROOF. Fix e > 0, the function u - v is upper semicontinuous and b o u n d e d above, so by the smooth minimization principle, there exist xo 6 R n and (p, Q) 6 D 2'+ (u - v)(xo) such that I]P]I < e, II QII < e and (u - v)(xo) > sup(u - v) - e. Applying T h e o r e m 4.5 with ul = v and u2 = - u , there exist Xl, x2 in R n, (pl, Q1) E D2'-V(Xl) and (p2, Q2) E D2,+u(x2), such that: (1) [[x l - xo [[ < e and [Ix2 [[ < 6. (2) IV(Xl) - v(x0)ll < E and lU(X2) u(x0)ll < e. (3) liP2 - Pl - pll < e and II Q2 - - Q 1 - QII < e. The function u is a viscosity subsolution of u 4- H (x, Du, D2u) = 0, so --
XO
--
u(x2) 4- H(x2, P2, Q2) ~< 0.
Perturbed minimization principles and applications
431
The function v is a viscosity supersolution of v + H (x, D r , D2v) - - 0 , so v ( x l ) + H ( x l , p l , Q1) >/0. Consequently sup(u - v) <<, (u - v)(xo) + e < u(x2) - l)(Xl) + 3e < H ( X l , Pl, Q1) - H ( x 2 , p2, Q2) + 3e. Moreover, Ilxl-x2ll ~ Ilxi-xoll-+-llxo-x21[ < 2e, I I p 2 - p l II ~ I I p 2 - p l - p l l - + - I l p l l < 2e and IIQ2 - Q1 II ~< IIQ2 - Q1 - Q II + IIQ II < 2~. Using the uniform continuity of H and sending ~ to zero, we get that sup(u - v) ~< 0. E2 REMARK 6.2. Using these techniques, the reader will now be able to deal with more general Hamiltonians and to prove uniqueness results for second order parabolic HamiltonJacobi equations. We have omitted several topics of interest as we did not mention boundary conditions and how they can be interpreted in the viscosity sense. We did not mention either the uniqueness results for viscosity solutions of Hamilton-Jacobi equations in some classes of unbounded functions. For more on these topics, see [6,18], and references therein.
6.4. The existence o f viscosity solutions f o r H a m i l t o n - J a c o b i equations A classical method to show the existence of an harmonic function w on an open subset 12 of R n satisfying a given boundary condition (say w = f on Ol2) is due to Perron: the idea is to first establish the existence of a subharmonic function u0 and of a superharmonic function v0 such that u0 ~< v0 on S-2 and u0 = v0 - f on OY2, and then to show that the supremum of all subharmonic functions w satisfying u0 ~< w ~< v0 is actually harmonic and satisfies the desired boundary condition. Our aim here is to show, following Ishii, that this method can be adapted to yield the existence of viscosity solutions for Hamilton-Jacobi equations under very general hypotheses. In order to do this, we introduce some notation. If I2 is an open subset of a Banach space X and if u is a function defined on S-2, the upper semicontinuous envelope u* of u is u* = inf{v; v is continuous on 12 and v >~ u on $2 } and the lower semicontinuous envelope u. is defined similarly. We can now state: THEOREM 6.4. Let X be a Banach space satisfying (H1) and let 12 be an open subset o f X. Let H : 1-2 • R x X* --+ R be continuous. Let uo, vo be respectively a viscosity subsolution and a viscosity supersolution o f H (x, u, D u ) -- 0 on S-2 such that uo <<.vo on 1-2. Then there exists a function u : S-2 --+ R such that uo <<.u ~ vo on S-2, u* is a viscosity subsolution o f H (x , u, D u ) = 0 on S-2 and u , is a viscosity supersolution o f H (x , u, D u ) = 0 on S-2.
R. Deville and N. Ghoussoub
432
PROOF. Let S be a family of functions which are locally uniformly bounded above and which are viscosity subsolution of H ( x , u, Du) -- 0 on S2. Let u = sup{w; w 6 S}.
CLAIM. u* is a viscosity subsolution of H (x, u, Du) -- 0 on I2. In particular, when S = {ul, u2}, we obtain that the supremum of two viscosity subsolutions of H ( x , u, Du) -- 0 on S-2 is a viscosity subsolution of H (x, u, Du) -- 0 on S-2. Indeed, let x E s and p ~ D+u*(x). According to the stability of superdifferentials (Theorem 4.7), there exist (Xn) C S-2, (Un)a sequence of functions in S and Pn ~ D+un(xn) such that: (1) (Xn) converges to x, (2) (un(xn)) converges to u*(x), and (3) (Pn) converges to p. Since Un is a viscosity subsolution of H ( x , u, Du) -- 0 on I2, we have H(xn, Un(Xn), Pn) -0. By continuity of H , we obtain, as n tends to infinity, H ( x , u*(x), p) = 0. Therefore, u* is a viscosity subsolution of H (x, u, Du) = 0 on S2. Define now
S = { w ' X --+ R; u0 ~ w ~< v0 and w is a viscosity subsolution of H (x, u, Du) -- 0 on S2 } and let u - - s u p { w ; w E S}. According to the above claim, u* is a viscosity subsolution of H (x, u, Du) -- 0 on S2. Let us now prove that u, is a viscosity supersolution of H (x, u, Du) = 0 on s Indeed, otherwise, there exist x0 6 S2 and a C l-smooth function g ' X --+ R such that: (i) u,(xo) - g(xo) -- 0 and u , ( x ) - g(x) >~0 for all x 6 X'2, (ii) H(xo, u,(xo), g'(xo)) < O. Observe that necessarily g(xo) < v(xo). Indeed, if not, then g(x) <~ u , ( x ) <<. v(xo) and g(xo) = u,(xo) - vo(xo) which imply that v0 - g attains its infimum at x0. Since v0 is a viscosity supersolution of H ( x , u, Du) = 0 on S2, H(xo, vo(xo), g~(xo)) ~ O, which contradicts (ii). Denote B(xo, 3) the open ball centered x0 and with radius 6 and B(xo, 6) its closure. By continuity of H and g~, there exist 6 > 0 such that B(xo, 23) C S2, and a bump function B ' X --+ R with support in B(xo, 6), such that b(xo) > 0 and which satisfies
H (x, g(x) + b(x), g'(x) + b'(x)) < 0
for all x E B(xo, 26),
and
g(x) + b(x) <~ vo(x)
for all x E S-2.
Indeed, this is possible if 6 > 0, and if [[bl[~ and [Ib'[[~ are small enough, because g(xo) < v(xo) and g(x) <~ v(x) for all x E $2. Define now
w(x)=
max{g(x) + b(x), u(x)}
u(x)
if x E B(x0, 26), if x ~ S-2\ B (x0, 26).
Perturbed minimization principles and applications
433
By the above claim, w is a viscosity subsolution of H (x, u, Du) -- 0 on I21 = B(x0, 26). On the other hand, if x 6 S'22 -'- S - 2 \ B ( x o , (~), then w ( x ) = u(x). Therefore, w is also a viscosity subsolution of H ( x , u, D u ) - 0 on S22. Since S21 and ~f'22 are open, w is a viscosity subsolution of H ( x , u, Du) = 0 on S-2 = S21 U S22. Since u0 ~< w ~< v0, we obtain that w ~< u on s But u (x) >~ w (x) ~> g (x) + b (x) on B(x0, 6) which implies u , ( x ) ~ g(x) + b(x) on B(xo, 6), and this contradicts the fact that u,(xo) = g(xo). [] REMARK 6.3. A slight modification of the above argument yields an existence result for second order H a m i l t o n - J a c o b i equation on a Banach space satisfying (H2). COROLLARY 6.1. Let X be a Banach space satisfying ( H I ) and let I-2 be an open subset o f X. Let H :I-2 x X* --+ R be a continuous Hamiltonian satisfying (,). Assume moreover that the function x --+ H (x, O) is bounded. Then there exists a unique bounded function u : X --+ R such that u is a viscosity solution o f u + H (x, Du) = 0 on I-2. PROOF. By hypothesis, there exists constants m, M such that, for all x E I2, m ~< H ( x , O) <~ M. Thus the constant functions equal to - m and - M are respectively viscosity supersolution and viscosity subsolution of H ( x , u, Du) = 0 on I-2. According to T h e o r e m 6.4, there exists a function u:I-2 --+ R such that - M ~< u ~< - m on s u* is a viscosity subsolution of H ( x , u, Du) - - 0 on I2 and u , is a viscosity supersolution of H ( x , u, Du) - - 0 on I2. Since u is bounded, u , and u* are bounded. According to the m a x i m u m principle (Theorem 6.1), u* ~< u , . Since obviously u , ~< u ~< u*, we get that u = u* - - u , is a continuous viscosity solution of (HJ1). Again the m a x i m u m principle shows that u is the unique b o u n d e d viscosity solution of (HJ 1). D
References [1] J.P. Aubin and H. Frankowska, Set Valued Analysis, Birkh~iuser, Basel (1990). [2] D. Azagra and R. Deville, Subdifferential Rolle's and mean value inequality theorems, Bull. Austr. Math. Soc. 56 (1997), 319-329. [3] D. Azagra and R. Deville, Starlike bodies in Banach spaces and a new characterization of separable Asplund spaces, Preprint. [4] M. Bachir, On generic differentiability and the Banach-Stone theorem, Preprint. [5] M. Bachir, Subdifferential calculus and smooth variational principles, Preprint. [6] G. Barles, Solutions de viscositd des dquations de Hamilton-Jacobi, Mathematiques & Applications (Paris), Springer-Verlag (1994). [7] R. Bonic and J. Frampton, Smooth functions on Banach manifolds, J. Math. Mech. 15 (1966), 877-898. [8] J.M. Borwein and D. Preiss, A smooth variational principle with applications to subdifferentiability and to differentiability of convex functions, Trans. Amer. Math. Soc. 303 (1987), 517-527. [9] J.M. Borwein and Q. Zhu, Viscosity solutions and viscosity subderivatives in smooth Banach spaces with applications to metric regularity, SIAM J. Control Optim. 34 (1996), 1568-1591. [ 10] K-C. Chang, Variational methods for non-differentiable functionals and their applications to partial differential equations, J. Math. Anal. Appl. 80 (1981), 102-129. [ 11 ] M. Choulli, R. Deville and A. Rhandi, A general mountain pass principle for non differentiable functionals, Rev. Mat. Apl. 13 (1992), 45-58. [12] F. Clarke, Generalized gradients and applications, Trans. Amer. Math. Soc. 205 (1975), 247-262. [13] F. Clarke and Y. Ledyaev, Mean value inequalities, Proc. Amer. Math. Soc. 122 (1994), 1075-1083.
434
R. Deville and N. Ghoussoub
[14] E Clarke and Y. Ledyaev, Mean value inequalities in Hilbert spaces, Trans. Amer. Math. Soc. 344 (1994), 307-324. [15] E Clarke, Y. Ledyaev and R. Stem, Invariance, monotonicity and applications, Non Linear Analysis, Differential Equations and Control, NATO Science Series, Series C 528, Kluwer Academic, Dordrecht (1999), 207-305. [16] E Clarke, Y. Ledyaev, R. Stern and P. Wolenski, Nonsmooth Analysis and Control Theory, Grad. Texts in Math. 178, Springer-Verlag, New York (1998). [ 17] J-N. Corvellec, M. Degiovanni and M. Morzocchi, Deformation properties for continuous functionals and critical point theory, Topol. Methods Nonlinear Anal. 1 (1993), 151-171. [ 18] M.G. Crandall, H. Ishii and EL. Lions, User's guide to viscosity solutions ofsecond order partial differential equations. Bull. Amer. Math. Soc., New Ser. 27 (1) (1992), 1-67. [19] M.G. Crandall and EL. Lions, Viscosity solutions of Hamilton-Jacobi equations, Trans. Amer. Math. Soc. 277 (1983), 1-42. [20] M.G. Crandall and EL. Lions, Hamilton-Jacobi equations in infinite dimensions, Part I: Uniqueness of viscosity solutions, J. Funct. Anal. 62 (1985), 379-396; Part II: Existence of viscosity solutions, J. Funct. Anal. 65 (1986), 368--405; Part III: J. Funct. Anal. 68 (1986), 214-247; Part IV." Unbounded linear terms, J. Funct. Anal. 90 (1990), 237-283; Part V." B-continuous solutions, J. Funct. Anal. 97 (1991), 417-465; Part VI: Nonlinear A and Tataru's method refined; Part VII: The HJB equation is not always satisfied. [21] D.G. de Figueiredo, The Ekeland Variational Principle with Applications and Detours, Tata Institute of Fundamental Research, Bombay (1989). [22] R. Deville, A mean value theorem for non differentiable mappings, Serdica Math. J. 21 (1995), 59-66. [23] R. Deville, Stability of subdifferentials of nonconvex functions in Banach spaces, Set Valued Analysis, Vol. 2, H. Attouch and M. Thera, eds (1994), 141-157. [24] R. Deville, G. Godefroy and V. Zizler, A smooth variational principle with applications to Hamilton-Jacobi equations in infinite dimensions, J. Funct. Anal. 111 (1993), 197-212. [25] R. Deville, G. Godefroy and V. Zizler, Smoothness and Renormings in Banach Spaces, Pitman Monographs in Mathematics 64, Longman Scientifc and Technical (1993). [26] R. Deville and C. Finet, Vector valued variational principles, Preprint. [27] R. Deville and E.M. E1 Haddad, The subdifferential of the sum of two functions in Banach spaces, I. First order case, J. Convex Anal. 3 (2) (1996), 295-308. [28] R. Deville and E.M. E1Haddad, The subdifferential of the sum of two functions in Banach spaces, L Second order case, Bull. Austr. Math. Soc. 51 (1995), 235-248. [29] R. Deville and M. Ivanov, Smooth variational principle with constraints, Math. Nachr. 69 (1997), 418-426. [30] R. Deville, R. Gonzalo and J. Jaramillo, Renormings of LP (Lq), Math. Proc. 126 (1999), 155-170. [31] R. Deville and J. Revalski, Porosity of ill-posed problems, Proc. Amer. Math. Soc., to appear. [32] I. Ekeland, Nonconvex minimization problems, Bull. Amer. Math. Soc. 1 (1979), 443-474. [33] I. Ekeland, Convexity Methods in Hamiltonian Mechanics, Springer-Verlag, Berlin (1990). [34] M. Fabian, Subdifferentials, local e-supports and Asplund spaces, J. London Math. Soc. 34 (1986), 568576. [35] M. Fabian, Subdifferentiability and trustworthiness in the light of a new variational principle of Borwein and Preiss, Acta Univ. Carolinae 30 (1989), 51-56. [36] M. Fabian, E H~ijek and J. Vanderwerff, On smooth variational principles in Banach spaces, J. Math. Anal. Appl. 197 (1996), 153-172. [37] M. Fabian, D. Preiss, J.H.M. Whitfield and V. Zizler, Separating polynomials on Banach spaces, Quart. J. Math. Oxford 40 (1989), 409-422. [38] G. Fang, The structure of the critical set in the general mountain pass principle, Ann. Fac. Sci. Toulouse, S6r. 6 III (1994), 345-362. [39] G. Fang, On the existence and the classification of critical points for non-smooth functionals, Canadian J. Math. 47 (1995), 684-717. [40] G. Fang, Min-max methods in critical point theory-selected topics, PhD dissertation, The University of British Columbia (1993). [41 ] G. Fang and N. Ghoussoub, Morse type information on Palais-Smale sequences obtained by min-max principles, Comm. Pure Appl. Math. 47 (1994), 1595-1653.
Perturbed minimization principles and applications
435
[42] G. Fang and N. Ghoussoub, Second order information of Palais-Smale sequences in the mountain pass theorem, Manuscripta Math. 75 (1992), 81-95. [43] N. Ghoussoub, Duality and Perturbation Methods in Critical Point Theory, Cambridge Tracts in Math. and Mathematical Physics 107, Cambridge University Press (1993). [441 N. Ghoussoub, Location, multiplicity and Morse indices of min-max critical points, J. Reine Angew. Math. 417 (1991), 27-76. [451 N. Ghoussoub, J. Lindenstrauss and B. Maurey, Analytic martingales and plurisubharmonic barriers in complex Banach spaces, Contemporary Math. 85 (1989), 111-130. [461 N. Ghoussoub and B. Maurey, H~-embeddings in Hilbert space and optimization on G 6-sets, Mem. Amer. Math. Soc. 349 (1986), 1-101. [47] N. Ghoussoub and B. Maurey, Plurisubharmonic martingales and barriers in complex quasi-Banach spaces, Ann. Inst. Fourier 39 (4) (1989), 1007-1060. [481 N. Ghoussoub and B. Maurey, Balayage, dentability and optimization on non-compact manifolds, in preparation. [49] N. Ghoussoub, B. Maurey and W. Schachermayer, Pluriharmonically dentable complex Banach spaces, J. Reine. Angew. Math. 402 (1989), 76-127. [50] N. Ghoussoub and D. Preiss, A general mountain pass principle for locating and classifying critical points, Ann. Inst. H. Poincar6 6 (1989), 321-330. [511 G. Godefroy, Some remarks on sub-differential calculus, Revista Matem~itica Complutense 11 (1998), 269279. [521 R.G. Haydon, A counterexample to several questions about scattered compact spaces, Bull. London Math. Soc. 22 (1990), 261-268. [531 R.G. Haydon, Normes infiniment differentiables sur certains espaces de Banach, C. R. Acad. Sci. Paris S6rie 1 315 (1992), 1175-1178. [541 R.G. Haydon, Trees and renorming theory, to appear. [551 A.D. Ioffe, On subdifferentiability spaces, Ann. New York Acad. Sci. 410 (1983), 107-119. [561 R. Jensen, The maximum principle for viscosity solutions of fully nonlinear second order partial differential equations, Arch. Rat. Mech. Anal. 101 (1988), 1-27. [57] N.J. Kalton, N.T. Peck and J.W. Roberts, An F-space Sampler, London Math. Society Lecture Notes 89, Cambridge University Press (1985). [58] P. Kenderov and R. Lucchetti, Generic well posedness of supinf problems, Bull. Austr. Math. Soc. 54 (1996), 5-25. [59] M. Leduc, Densitd de certaines familles d'hyperplans tangents, C. R. Acad. Sci. Paris 270 (1970), 326-328. [60] A. Maaden, Thdorkme de la goutte lisse, Rocky Mountain J. Math. 25 (1995), 1093-1101. [61] R.R. Phelps, Convex Functions, Monotone Operators and Differentiability, 2nd edn, Lecture Notes in Math. 1364, Springer-Verlag, New York (1998). [62] D. Preiss, Differentiability of Lipschitz functions on Banach spaces, J. Funct. Anal. 91 (1990), 312-345. [63] L. Zajf6ek, Porosity and or-porosity, Real Anal. Exchange 13 (1987-1988), 314-350.
This Page Intentionally Left Blank
CHAPTER 11
Operator Ideals Joe Diestel Department of Mathematics, Kent State University, Kent, OH, USA E-mail: diestel@ mcs. kent. edu
Hans Jarchow Department of Mathematics, University of Ziirich, Ziirich, Switzerland E-mail: [email protected]
Albrecht Pietsch Department of Mathematics, Jena University, Jena, Germany E-mail: [email protected]
Contents 1. Basic concepts and closed ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Banach and quasi-Banach ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Approximation numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Ultraproducts, maximality and trace duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. S u m m i n g operators and some of their relatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. L p-factorable operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7. Grothendieck's theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. Concrete operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9. Rademacher type and cotype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10. U M D operators, Haar type, and uniform convexity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11. Extension of classical theorems to vector-valued functions . . . . . . . . . . . . . . . . . . . . . . . . . 12. Operator ideals and tensor products, the approximation property . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
H A N D B O O K OF T H E G E O M E T R Y OF B A N A C H SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 437
439 448 452 455 458 464 466 468 471 476 480 484 490
This Page Intentionally Left Blank
Operator ideals
439
1. Basic concepts and closed ideals Standard references: [ 135, Part 1], [42]. 1.1. The theory of operator ideals was born in 1941 with the following observation of Calkin [23]: The ring B of bounded everywhere defined operators in Hilbert space contains nontrivial two-sided ideals. This fact, which has escaped all but oblique notice in the development of the theory of operators, is of course fundamental from the point of view of algebra and . . . . As examples of two-sided ideals in 13 we may mention here the class of all operators A such that R(A), the range of A, has a finite dimension number, the class of all operators of Hilbert-Schmidt type, and the class of all totally continuous operators. Let H be an infinite dimensional separable Hilbert space, and write ~ ( H ) instead of B to denote the ring of (bounded linear) operators from H to itself. Then the main results of Calkin's paper can be stated as follows: 9 ~ ( H ) , the set of finite rank operators, is the smallest proper ideal in ,~(H). 9 ~3(H), the closed set of completely (totally) continuous operators, is the largest proper ideal in ~ ( H ) . 9 Ideals between ~ ( H ) and ~23(H) exist in abundance; we take particular note of the Schatten-von Neumann classes ~ p (H), the most prominent of which is 6 2 ( H ) , the class of Hilbert-Schmidt operators; see [ 166,169]. Identifying H with 12, operators T E ~52(/2) are characterized by the property that the representing matrix (rhk) has finite double norm; see [66,171]:
o'2 (T) "--
IZ'hk 12
.
h--1 k=l Obviously, ~52 (/2) is a Hilbert space with this norm, which differs from the operator norm. Schur's inequality [ 173] says that
IXnl2
~< o'2(T),
where (Xn) denotes the eigenvalue sequence of T 6 ~52(H). In 1949, Weyl [190] proved a natural generalization of this inequality to operators T ~ ~ p ( H ) with 0 < p < ~ . Since their inception, Schatten-von Neumann operators have become an indispensable tool of spectral theory in Hilbert spaces; see [55,157,168] and [181]. 1.2. Schatten and von Neumann actually dealt with more general classes of operators, so-called cross-spaces of linear transformations. They phrased their results in the language of direct (now tensor) products of Banach spaces [166,167] and [169]. Historically, this work was one of the starting points of Grothendieck's fundamental investigations. His
J. Diestel et al.
440
famous Rdsumg [60] became the source of modern Banach space theory. However, after a first attempt made by Amemiya and Shiga [2], an incubation period of more than a decade was needed to understand and appreciate Grothendieck's ideas. In the late sixties, Lindenstrauss and Petczyfiski published a pioneering paper [104], which stresses the following point of view: Though the theory of tensor products has its intrinsic beauty we feel that the results of Grothendieck's paper and their corollaries can be more clearly presented without the use of tensor products. Already the revised theory of nuclear locally convex spaces [129] and the study of absolutely summing operators had shown that there is a very handy alternative to tensor products: operator ideals. 1.3. Before we proceed, let us fix some notation: Throughout, t_ -- {X, X0, Y, Y0, Z . . . . } stands for the class of all (real or complex) Banach spaces, and IN will denote the scalar field, be it IR or C. The norm of a Banach space X is denoted by I1" II, but sometimes we use the symbol I1" I X II for better distinction. We denote the set of all (bounded linear) operators from X (domain) into Y (codomain) by t3(X, Y), and write t2 " - ~ t2(X, Y), where the union ranges over X, Y E t_. As usual, I[TI[ or I [ T ' X --+ YII stands for the operator norm of T E t3(X, Y). The identity map of a Banach space X will be denoted by Ix, or just by I. Furthermore, I will also refer to canonical embeddings like I ' l l --+ loc. Given any functional x~ in the dual space X* and any element y0 in Y, we let x 0 | y0" x ~ (x, x 0) y0. We stress that C and _ are used in the same way as < and ~<; that is, the symbol C indicates proper inclusions. 1.4. The notion of an operator ideal was invented once it became clear that restriction to the ring t2(X) of operators on a fixed Banach space X is too limiting. Rather, in contrast to the Hilbert space setting, operators between different Banach spaces must be taken into account. The basic concepts were first presented by Pietsch [ 131,133], in the preliminary version of his monograph Operator Ideals. The subject has experienced, and continues to experience, considerable development accompanied and inspired by striking applications in diverse areas of mathematical endeavour. Suppose that, in every 13(X, Y), we are given a subset 9A(X, Y). Then
~1 "-- U 9A(X, Y) X,Y
is said to be an operator ideal, or just an ideal, if the following conditions are satisfied: (Oil)
x* | y E 9,1(X, Y) for x* E X* and y E Y.
(OI2)
S + T E 9,1(X, Y) for S, T E 9A(X, Y).
(013)
B T A E P,I(X0, Y0) for T E 9A(X, Y), A E 13(X0, X) and B E t3(Y, Y0).
Operator ideals
441
Since (OI3) implies that )~T E 9A(X, Y) for T E 9A(X, Y) and ~ E K , every component 9A(X, Y) is a linear subset of s Y). For brevity, we shall write 9A(X) instead of ~(X,X). 1.5. An illustrative example might convince the reader of the power inherent in ideal theory. Given any complex Banach space X, the Fredholm resolvent set p ( T ) of an operator T E ~ ( X ) consists of all )~ E C for which I - )~T has an inverse in ~ ( X ) . In the setting of integral equations the Fredholm resolvent is defined by F(~) :-- T ( I - s -1 . Then ( I - ~T) -1 = I + LF()~), and we have the advantage that F(L) E 9A(X) whenever T belongs to an ideal 9A(X). 1.6. We emphasize that the strength of ideal theory lies in its concrete examples. Many of them will be discussed in the course of this article. An operator T E ~ ( X , Y) has finite rank if rank(T) "-- dim{Tx" x E X} is finite. Such an operator admits a finite representation
T--
xk |
with x 1* . . . . . x n* E
X*
and Yl . . . . . Yn E Y.
(1.6a)
k--I
These operators form the smallest ideal, denoted by ~. The symbol | used above and the formal identification ~ ( X , Y) = X* N Y indicate the close link to the theory of tensor products, which will be presented in Section 12. 1.7. The dual ideal ~..[dual consists of all operators T such that T* belongs to the given ideal 9,1. We say that 9,1 is symmetric if 9,1 -- 9,1dual. The closed hull ~ of an ideal 9,1 has the components 9A(X, Y); the closure refers to the uniform topology on s Y). Obviously, 9,1 is also an ideal; and ideals 9A such that 9,1 -- 9,1 are said to be closed. Apart from the very last paragraph, the rest of this section is devoted to closed ideals. m
1.8. For obvious reasons, we refer to operators in ~ as approximable. Using the principle of local reflexivity, Hutton [70] proved that ~ is symmetric. That even the elementary concepts above have some power, can be illustrated by sketching how to reduce a certain infinite dimensional problem to a problem of linear algebra. Suppose we are given an operator T E ~ ( X ) . Choose some A E ~ ( X ) such that Q - T - A has norm less than 1. Then I - Q is invertible and, as was already noted by Schmidt [171, Part II], the equations x - T x - - y and x - ( I - Q ) - l A x ( I - Q ) - l y are equivalent. Since ~ ( X ) is an ideal, we have (I - Q ) - I A E ~ ( X ) . Hence the element x that we are looking for can be found by solving a linear system in finitely many variables. Considerations like these were among the starting points of functional analysis. 1.9.
In the classical work of Hilbert [66, p. 200], a scalar-valued function f on the U3
l/)
closed unit ball of 12 is called vollstetig if Xn --+ x implies f (Xn) --+ f (x). We use Xn --+ x to indicate that Xn tends to x coordinatewise. However, since we are in a bounded set, this
J. D i e s t e l et al.
442
is just weak convergence. Hilbert, following the tradition of Weierstrass and Frobenius, applied his definition to bilinear and quadratic forms. Soon thereafter, Riesz [155, p. 96] introduced the concept of a substitution complktement continue acting in 12. Nowadays, an operator T 6 13(X, Y) is called completely continuous if it carries weakly convergent sequences to norm convergent sequences. These operators form a non-symmetric closed ideal, denoted by ~3. Already Riesz [155, p. 113, footnote] observed that ~J(12) - ~(12). On the other hand, we know from Schur [174] that ~J(ll) = t3(11). 1.10.
In 1916, Riesz [156] extended Hilbert's spectral theory to operators acting on
C[a, b] and, in fact, to Banach spaces: Die in der Arbeit gemachte Einschr~inkung auf stetige Funktionen ist nicht von Belang. Der in den neueren Untersuchungen tiber diverse Funktionalr~iume bewanderte Leser wird die allgemeinere Verwendbarkeit der Methode sofort erkennen. As a first step, he had to find a suitable class of operators, since certain components of the ideal ~ are too large. We do not know his reasoning. Nevertheless, the crucial Riesz lemma says that a Banach space is finite dimensional if and only if all bounded subsets are relatively compact. This basic observation suggested to consider those operators that map bounded sets into relatively compact sets. For many years such operators were called alternatively vollstetig (Schauder) or opdrations totalement continues (Banach). It was Hille [67, p. 14] who proposed to use the label compact. The compact operators form a symmetric closed ideal, which will be denoted by .~. The symmetry expresses a classical result of Schauder [170] who proved the equivalence of T E N and T* 6 .~. We also note that an operator T :X --+ Y is compact if and only if every bounded sequence in X admits a subsequence whose image is norm convergent in Y. The celebrated Riesz-Schauder theory says that the determinant-free results of Fredholm's theory of integral equations carry over to operators T E N(X) and their duals. So it is extremely important to verify the compactness of concrete operators. We illustrate the facility and elegance of ideal theoretic considerations by a typical application. Let S2 be a bounded domain in IRn with a smooth boundary. Denote by W2 (I-2) the Sobolev space consisting of all functions which, together with their second weak derivao
tives, belong to Lp (I2), where 1 < p < ec. Form W2 (I2), the subspace of all functions in W2 (g2) that vanish on the boundary. Then the Laplacian A defines an isomorphism between IV 2p (I-2) and L p (I2). Moreover, the embedding map I" W2 (I2) --+ L p (S2) is compact: A-1
Lp(~2) ,
A
o
~ W2p(S-2)
I
,. Lp(S-2).
Thus the ideal properties of .~ tell us that I A -1 is a compact operator on L p(I2), and the Riesz-Schauder theory applies. Classically, this was done by representing I A -1 as an integral operator defined by the Green's function. 1.11. An operator T 6 t3(X, Y) is called weakly compact if it takes bounded subsets of X to weakly relatively compact subsets of Y. These operators form a closed ideal, which
Operator ideals
443
we denote by 913, with symmetry reflecting a now classical result of Gantmacher [50]. She also knew that weak compactness is characterized by the property that T** maps X** into Y, viewed as a subspace of Y**. Thanks to the Eberlein-Smulian theorem, T" X --+ Y is weakly compact if and only if every bounded sequence in X admits a subsequence whose image is weakly convergent in Y. Weakly compact operators were introduced by Kakutani [82] and Yosida [193] when they proved a Banach space version of the mean ergodic theorem: if the norms ]lTn ]l of T ~ 9I]'(X) are uniformly bounded, then the averages x+Tx+"+vn-'x converge in norm to a limit Px for all x 6 X, and x ~ Px defines a n projection from X onto T's set of fixed points. 1.12.
The infinite summation operator ~ E ~(/1, lo~) defined by
r " (~k) k=l
is not weakly compact; indeed, it is in a sense the prototype of all operators that fail to be weakly compact. An operator To" X0 --+ Y0 not contained in an ideal 9,1 is called CgA-universal if it factors a
through all operators T ' X --+ Y in the complement of 9,1. More precisely, To'Xo X
T
Y B ~ Y0. In this sense, Z" is C~3-universal, while the embedding map I ' l l --+ lo~ turns out to be C.~-universal; see [104] and [79]. More examples will be given in 1.18 and 11.6. On the other hand we know from [52] that a C93'-universal operator does not exist. 1.13. With every operator ideal we associate the collection of all Banach spaces X 6 k such that Ix, the identity map of X, belongs to 9A. The class A so obtained has the following properties"
(s~) (s2) (s3) (s4)
A contains all finite dimensional Banach spaces. If X belongs to A, then so does any isomorphic copy. A is stable when passing to complemented subspaces. A is stable under the formation of finite direct sums.
Our general strategy will be to use Gothic bold uppercase letters to denote operator ideals, while the associated classes of spaces are denoted by the corresponding sans serifs. For example: ~ --+ F, .~ --+ K, 93 --+ V and 913 --+ W. Banach spaces in V are said to have the Schur property: norm and weak convergence of sequences coincide. In particular, 11 6 V. Moreover, W is the class of reflexive spaces. Since F - K, different ideals may well define the same class of spaces. This is in particular true for 9A and 9,1. So the language of operator ideals offers manifold opportunities to describe specific properties of Banach spaces, giving this landscape a much more colourful view. Let A be any class of spaces satisfying (S1) to ($4). Then all operators factoring through some member in A form an ideal, which is said to have thefactorization property; see [65]. In this way, 913 can be reconstructed from W: identities of reflexive Banach spaces provide the prototypes for all weakly compact operators [31 ]. Further examples are treated in 1.16, 1.18 and 11.6.
J. Diestel et al.
444
1.14. So far, we have dealt with the oldest generation of closed ideals: ~ , .~, ~3 and ~ ' . For further consideration it will be advantageous to have a few additional abstract concepts in hand. Let 9,1 and f13 be ideals. The (X, Y)-component of the product ~ o 9,1 consists of all A
B
operators admitting a factorization T ' X >M > Y, where M is a suitable Banach space, A 6 9,1 and B 6 f13. We have ~3 o ~ = .~. An ideal 9,1 is said to be idempotent if 9,1 = 9,1 o 9,1. The ideals ~, ~ , .~ and ~ 3 enjoy this property, but ~23 does not [135, p. 60]. Every ideal with the factorization property is idempotent. But, as shown by .~, the reverse implication may fail. The (X, Y)-component of the quotient f13 -1 o 9,1 consists of all operators T 6 ~ ( X , Y) T such that, regardless of how we choose Y0 and B 6 fl3(Y, Y0), the composition X > y
8
Y0 is a member of 9,1. Similarly, the (X, Y)-component of the quotient f13 o 9,1-1 is A
T
the set of all operators T 6 s Y) such that X0 >X > Y belongs to f13 whenever A 6 9A(X0, X). It is no problem to verify that ~23 o 9.1, f13-1 o 9.1 and f13 o ~.[-- 1 are ideals. 1.15. Grothendieck [58] was the first to define a subclass of Banach spaces using operator ideals. A space X is said to have the Dunford-Pettis property if every weakly compact operator T" X ~ Y is completely continuous, for any choice of Y. The collection of these spaces will be denoted by DP. Dunford and Pettis [41 ] proved that L1 (M, #) 6 DP and, more than 10 years later, Grothendieck added the relation C ( K ) E DP. He even showed that 93(C(K), Y) -- ~ 3 ( C ( K ) , Y) for all Y. Therefore the Riesz-Schauder theory extends to completely continuous operators on C (K), since their squares are compact. A Banach space X has the Dunford-Pettis property if and only if Ix belongs to 53 ~3 "-- ~ 3 - 1 o f13. So it is natural to use the same terminology when referring to operators T" X --+ Y in this quotient. A necessary and sufficient condition is that the scalar sequence ((Txn, y~)) tends to zero whenever (Xn) is a weak null sequence in X and (Yn*) is a weak null sequence in Y*. As an easy corollary we get X)9~3dual ___53r hence DP dual c_ DP. If X - - Y --/2, then T E ~D~]3(/2) means that the bilinear form (x, y*) ~ (Tx, y*) on 12 • 12 is vollstetig in Hilbert's original sense. 1.16. An operator T : X --+ Y is said to have the Banach-Saksproperty if every bounded sequence in X admits a subsequence whose image has norm convergent arithmetic means. The class of these operators, denoted by f13~, is a non-symmetric closed ideal with the factorization property; see [3,47,7] and [65]. Closedness is proved with the help of a Ramsey type theorem. The terminology refers to a classical result of Banach and Saks [4] who showed that L p E [35 if 1 < p < oo. We have f13 ~ C ~3. 1.17. Let us proceed with some more notation: An operator J E ~(X, Y) is called an injection if there exists a constant c > 0 such that ]lJx[] >~ cIIx[[ for all x 6 X. In the case when [[Jx[] = IIx[[, we speak of a metric injection. For every (closed linear) subspace M of X, the canonical embedding from M into X is denoted by JM x. An important example of a metric injection is the canonical map K x : X ~ X**, which assigns to every element x in X the functional x* w-~ (x, x*) in X**.
Operator ideals
445
An operator Q from X onto Y is called a surjection. In the case when the open unit ball of X is mapped onto the open unit ball of Y, we speak of a metric surjection. For every (closed linear) subspace N of Y, the quotient map Q~v from Y onto Y / N has this property. 1.18. One may ask whether there exist larger classes of operators for which the basic results of the Riesz-Schauder theory remain true. This question led Kato [85] in 1958 to introduce the notion of a strictly singular operator. Such operators are defined by the property that T j X cannot be an injection for any infinite dimensional subspace M. In other words, if IITxll ~> cllxll for all x 6 M and some c > 0, then dim(M) < cx~. These operators form a closed ideal, which will be denoted by 6 . Petczyfiski [121 ] found a dual counterpart, the closed ideal ~s of strictly cosingular operators. By definition, T E if(X, Y) if Q~v T cannot be a surjection for any infinite dimensional quotient Y/N. Bourgain and Diestel [ 17] showed that an operator T is strictly cosingular if T* carries weak* convergent sequences to norm convergent sequences. It turns out that .~ C ~ and N C ~s The righthand ideals are much larger. For example, the summation operator 27:11 --+ 1~ is strictly singular, and Kco'co --+ c~* is strictly cosingular. Moreover, ~d,at C ~ and ~5d"at C ~. The concept of strict singularity admits various generalizations. For 1 ~< p < c~, an operator T E ~ ( X , Y) is said to be lp-singular if T J x cannot be an injection for any subspace M isomorphic to l p. This produces a 1-parameter scale of pairwise different closed ideals, denoted by ~ [ p . In the limiting case p = cx~,the space I~ should be replaced by co. Thus we get the ideal ~c0. From Rosenthal's ll-theorem [159] it follows that an operator belongs to ~11 if and only if the image of every bounded sequence contains weak Cauchy subsequences; such an operator is thus said to have the Rosenthalproperty. Since completely continuous operators map weak Cauchy sequences into norm convergent sequences, it is readily seen that ~ !l ~3"-I o .~. Any isometric embedding from ll into 1~ is CESll-universal. An operator T E s Y) is c0-singular if and only if ~ k ~ J Txk converges in norm for all sequences (xk) such that )--~--11(x~,x*)l < ~ whenever x* E X*; see [151, p. 270]. Those operators are also called unconditionally summing (or converging). The canonical embedding from co into 1~ is C~Sc0-universal. Every/1-singular operator factors through a Banach space not containing any isomorphic copy of 11. That is, ~511 has the factorization property; see [48] and [189]. On the other hand by [53], the ideals ~5 and ~ c 0 even fail to be idempotent. 1.19. In continuation of 1.7 and 1.14, we now describe a few more methods to produce new operator ideals from old. We say that an operator T E ~ ( X , Y) belongs to the regular hull ~.[reg(X, Y) if K r T E 9A(X, Y**). Here Ky denotes the canonical map from Y into Y**. The injective hull 9...[inj (S, Y) consists of all operators T E ~ ( X , Y) that become a member of 9A by extending T J the codomain: X >Y > Y0, where J is an injection into a suitable space Y0. Similarly, the surjective hull 9As"~(X, Y) consists of all operators T E ~ ( X , Y) that become a member of 9,1 by lifting the domain: X0 Q X T > Y, where Q is a surjection defined on a suitable space X0. Due to the extension and lifting properties, in the definitions above we may take Y0 - loc (I[) and X0 -- l l (]I) with an appropriate index set. It is clear that 9Areg, ~inj and 9As"~ are ideals and that ~[reg C ~inj. m
J. Diestel et al.
446
An ideal is called regular if 9,1 = p.~lreg, injective if 9,1 - - ~[inj, and surjective if 9,1 = ~sur. Every dual ideal is regular. Moreover, if 9,1 is injective, then 9,1dual is surjective; similarly, if 9.1 is surjective, then p_~dual is injective. However, in neither case can one conclude the converse. By Enflo's [46] negative solution of the approximation problem, ~ is properly contained in .~, but it is easy to see that ~ inj __ -~sur __ ~.. Obviously, f13sur - ~ . Moreover, Weis [189] proved that ~5 s"r - ~inj_ ~ [ 1 . Injectivity and surjectivity of 9,1 imply that A, the associated class of Banach spaces, is stable w h e n passing to subspaces and quotients, respectively. In case of symmetry, X E A and X* 6 A are equivalent. 1.20.
Properties of ideals considered so far are listed in Table 1.
Table 1 Ideal
Symmetric
Regular
Injective
Surjective
Factorization property
Idempotent
yes yes no yes no no no no no no
yes yes yes yes yes yes yes no yes yes
no yes yes yes no yes yes no yes yes
no yes no yes no yes no yes yes no
no no no yes ??? yes no ??? yes no
yes yes no yes ??? yes no ??? yes no
J~ ~9" ~gt ~3~ ~; ~I1 ~5c0
1.21. Let H denote an infinite dimensional Hilbert space and (. I') its inner product. With e 6 H we associate the functional e* "x --+ (xle). Every operator T E ~ ( H ) admits a Schmidt representation oo
T--
men|
n,
(1.21a)
n=l
where (en) and (fn) are suitable orthonormal sequences and t = (rn) is a null sequence. We m a y arrange by simple manipulations that r l / > 1:2 ~> . . . >/0, and then the coefficients rn are uniquely determined. Indeed, (1.21 a) implies that Ten = rn fn and T* fn = rnen where, deviating from our usual notation, T* stands for the Hilbert space adjoint of T. r nen, and rn is the nth eigenvalue of the positive operator ~/T--;T. We Hence T* Ten 2 refer to Sn (T) := rn as the nth s-number (singular number) of T. Italics indicate that lp and co are viewed as Banach spaces in the usual sense. However, in order to stress the analogy with operator ideals, Gothic bold lowercase letters will be used to denote sequence ideals. These are permutation invariant (rearrangement invariant) -
-
Operator ideals
447
ideals a in the ring of bounded scalar sequences. That is, in addition to the ideal property we require that (rk) E a implies (rrr(k)) E a for every permutation Jr of N. Obviously, there is no other proper and closed sequence ideal in lec than co. Thus we have a counterpart of Calkin's theorem, which asserts that t3(H) contains one and only one proper and closed operator ideal: m
~(H)--.~(H)--93(H) ....
.
Consequently, proper ideals a and 9.1(H) are contained in co and . ~ ( H ) , respectively. Given any proper sequence ideal a, we consider the collection of all operators T E .~(H) having a Schmidt representation with t E a. This process yields a proper ideal in t2(H). Conversely, with every proper ideal 9.1(H) in s we may associate the set of all sequences t = (rn) such that the operator defined by (1.21a) belongs to 9.1(H). Clearly, this property does not depend on the special choice of the underlying orthonormal sequences. Thus we get a one-to-one correspondence: a +-~ 9.1(H). In particular, c0 +-~ .~(H). In this way, ideal theory in ~ ( H ) is reduced to the theory of permutation invariant ideals in [oc. This may be considered as a significant simplification, since a non-commutative situation is turned into a commutative one. Unfortunately, sequence ideals exist in abundance. For our purposes, the ideals [p and co are most important. The above procedure associates with [p the Schatten-von Neumann ideal ~ p(H). Given any Schmidt representation of r r (~p(H), put tip(T) := Iltllp II and note that this quantity is well-defined. It takes some doing to show that tip defines a norm when 1 <~ p < ec and a p - n o r m when 0 < p < 1. In both cases, ~ p (H) becomes complete with respect to the corresponding metric. Similarly, the Lorentz sequence ideals [p,q generate the operator ideals ~p,q(H). Several well-known results on [p have natural substitutes for ~Op(H). A m o n g others, a non-commutative analogue of HOlder's inequality holds [69]: if 0 < p, q, r < oc and 1/r = 1/p + 1/q, then
ST E ~Or(H)
and
tir(ST) <<.tip(S) tiq(T)
for S E ~Op(H) and T E ~Oq(H). Furthermore, every R E ~Or(H) can be written in the form R = ST, where S and r are as above and tir(R) = tip(S)tiq (T). Of particular interest are the Hilbert-Schmidt class ~ 2 ( H ) and the trace class ~ l (H). Originally, the latter was defined via I~51(H) = ~ 2 ( H ) o ~ 2 ( H ) . For 1 < p < oc and 1 / p + 1/p* = 1, the standard duality between [p and [p, also has a counterpart: trace duality. However, the presentation of this needs some preparation, and we postpone the discussion to Section 4. In the context of his Grundlagen der Quantenmechanik (1932), von Neumann looked for traces of infinite operators. His early reasoning was quite heuristic. Later, in collaboration with Schatten, he laid the theoretical background, based on previous work in the finite dimensional case [116]. Von Neumann's influence on the early period of the theory of operator ideals can hardly be overestimated; Calkin, Murray and Schatten were all von Neumann students.
J. Diestel et al.
448
2. Banach and quasi-Banach ideals Standard reference" [135, Part 2]. 2.1. By an ideal norm defined on an ideal PA we mean a rule a that assigns to every operator T 9 P,l a non-negative number a ( T ) such that the following conditions are satisfied: (IN1)
a(x* | y) = ]]x*l] ]]yl] for x* 9 X* and y 9 Y.
(IN2)
~(S 4- T) ~< r
(IN3)
o t ( B T a ) <<,IIBI] or(T) ]]all for T 9 9A(X, Y), a 9 ~(X0, X) and B 9 s
4- or(T) for S, T 9 PA(X, Y). Y0).
Since (IN3) implies that a ( ~ T ) -- ])~]ot(T) for T 9 9A(X, Y) and )~ 9 ]K, we have indeed a norm. Moreover, ]ITI] ~< a ( T ) . For greater clarity, we shall sometimes write a ( T " X --+ Y) instead of a ( T ) . In some examples, we will only have a weak form of (IN1), namely" or(T) > 0 if T r O, which is equivalent to ot (IK) > 0. If the later quantity is different from 1, then a is said to be non-normalized. In this case, (IN1) can be ensured by passing to a(IK)-la. Occasionally, more general concepts will be needed. We call a an ideal quasi-norm if the triangle inequality is replaced by a quasi-triangle inequality: ot(S + T) ~< c[a(S) + a ( T ) ] for S, T 9 PA(X, Y). Here the constant c ~> 1 does not depend on the underlying spaces X and Y. We speak of an ideal p-norm with 0 < p ~< 1 whenever ot(S 4- T) p <~ or(S) p 4ot(T)P. In this case, the quasi-triangle inequality holds for c :-- 21/p-1. Conversely, given some ideal quasi-norm a, a classical renorming process yields an equivalent ideal p-norm
~ ( T ) "= inf
~(Ti)P i=1
where p is defined by c
Ti, T1 . . . . . Tn 9 PA(X, Y), n 9 N
9T =
,
i=1
-- 2 l/p-1.
2.2. A Banach ideal is an ideal 9,1 equipped with an ideal norm a such that all components PA(X, Y) are complete with respect to the associated convergence. Quasi-Banach and p-Banach ideals are defined analogously. As it turns out, the ideal quasi-norm of a quasiBanach ideal is unique up to equivalence. Consequently, we may refer to a quasi-Banach ideal 9.1 without specifying or. In addition, the definition of a concrete quasi-Banach ideal automatically leads to a natural ideal quasi-norm, under which it becomes complete. However, there are cases where one and the same ideal can be obtained in different ways, and then we have, indeed, 'die Qual der Wahl'; see 3.5 and 9.10. We also stress that there are ideals which in no way can be made a quasi-Banach ideal. The most prominent example is ~, the ideal of finite rank operators. Of course, every closed ideal is a Banach ideal with respect to the operator norm. Every ideal p-norm ot defined on an ideal 9A induces a corresponding ideal p-norm otdual, Otreg, Olinj and asur on 9,1dual, ~reg, ~2[inj and PAsur, respectively. The basic intention of the above and the following definitions consists in preserving completeness. Given an ideal p-norm ot on 9,1 and an ideal q-norm fl on ~3, then fl-1 o ~ ( T ) " - - i n f { ~ ( B T ) " B e ~ ( Y , Y0), fl(B) <~ 1, Yo 9 L},
Operator ideals
449
provides an ideal p-norm on the quotient ~3-1 o ~.1. Similarly, we may get an ideal q-norm on ~3 o 9A-1 . Less obvious is the case of products. Then the expression
fl ore(T)'--inf{fl(B)ot(A)" T = B A ,
A E 9A(X, M),
B E~(M,Y),
}
M EL
yields an ideal r-norm on ~3 o 9,1, where 1/r - 1 / p + 1/q. This value of r can be improved in many concrete cases. 2.3. We now treat the most prominent non-closed Banach ideal 911, which consists of the 1-nuclear, or simply nuclear, operators. The index '1' is used with an eye on later generalizations. Passing from finite to infinite representations, we look at operators T" X --+ Y that can be written in the form O(3
T --
eo
E*
Xk | Yk
such that
k=l
x *1 , x 2* , . .
.
E
X*
Z IIx~ II IlYk II < ~,
(2.3a)
k=l
and yl, y2 . . . . E Y. It is readily seen that the infinite series of the terms
x[~ | yk converges in It(X, Y). The 1-nuclear ideal norm is given by eo
IIx~ II IlYk II,
vl (T) "-- inf Z k=l
where the infimum ranges over all 1-nuclear representations (2.3a). Clearly, 9 1 1 is the least Banach ideal, and for all nuclear operators T and any ideal norm a, associated with a complete ideal, we have 0t(T) ~< vl (T). It will be helpful to have at hand an alternate description of nuclearity. By normalization, we get eo
r-
|
with IIx~ll- 1, IlY~II- 1 a n d t - ( r k )
E [1.
k=l
This yields the factorization T 9X A > leo Dt> 11 B > Y. Here A : x ~ ((x,xk)), , Dt'(~k) w-~ (rk~k) and B ' ( o k ) ~ Y~C~__10kYk. Obviously, Dt is 1-nuclear and v l ( D t ) ]]Dt ]] = ]ltlll ]]. Hence diagonal operators from leo to ll are the prototypes for all 1-nuclear operators. The concept of nuclearity was introduced in the early fifties for two reasons. Inspired by the theory of distributions, Grothendieck looked for an abstract version of Schwartz's kernel theorem; this explains the origin of the term nuclear. Secondly, Grothendieck [61 ] and Ruston [160] extended Fredholm's determinant theory to operators in Banach spaces. To this end they needed a sufficiently large class of operators for which the concept of a trace is meaningful. For a Hilbert space H, the Schatten-von Neumann class ~ l (H) turned out to be appropriate. Since 6 1 (H) = 911 (H), it was thought that 9tl would be
J. Diestel et al.
450
the right choice for general Banach spaces; indeed the nuclear operators were referred to as opdrateurs gt trace or trace class operators; see [57] and [162, p. 166]. Ironically, Enflo's negative solution of the approximation problem implied that there were trace class operators without trace! We now address this anomaly. 2.4.
The trace of an (n, n)-matrix T = (rh~) is defined by
trace(T) := ~
r~k = ~
k--1
~k,
k--1
where ~,1 . . . . . )~n are the eigenvalues. With a little bit of linear algebra, the concept of a trace can be extended to finite rank operators T acting from a Banach space X to itself. It turns out that the expression H
trace(T) := Z
(x/~, x/~)
k=l
does not depend on the finite representation
T --
x k @ xk
with x 1* . . . . . x n* ~
X*
and Xl . . . . . Xn ~ X.
k----1
In this way, we get a linear functional T ~ trace(T) on every component ~ ( X ) , which is uniquely determined by the condition that trace(x* | x) - (x*, x) for x* 6 X* and x E X. Very important is the commutative law: trace(AT) = trace(TA)
for T 6 ~(X, Y) and A 6 ~(Y, X).
This formula relates traces defined on components ~ ( X ) and ~(Y). It is natural to think of extending the trace functional by continuity. Using Auerbach's lemma, we obtain the inequality Itrace(T) I ~< nllTll whenever rank(T) ~< n. Since the identity map of l~ shows that the factor n cannot be improved, the trace functional is discontinuous with respect to the operator norm in infinite dimensions. The next idea that comes to mind is to replace the operator norm by a larger ideal norm. But even the largest, namely vl, is not good enough, as we shall explain next. Given T ~(X, Y), put
Ilxk II Ily~ll,
I~1 (T) "-- inf k=l
where the infimum only ranges over finite representations (1.6a). We obviously have [trace(T)] ~< v~(T) for T 6 ~(X). Since the definition of Vl(T) permits representations that are not necessarily finite, we can only conclude that I~1 ( T ) ~ I ~ ( T ) . Even more: the
Operator ideals
451
norms vl (T) and v~ (T) may be non-equivalent on ~(X, Y) for certain spaces X and Y. In the language of tensor products, the situation can be described by stating that the completion of X* | Y -- ~(X, Y) with respect to the norm v~ is canonically mapped onto 9tl (X, Y), but this surjection may fail to be one-to-one; see 12.8. 2.5. If the underlying Banach space X has the approximation property, then the above problems disappear, and every operator in r (X) admits a well-defined trace. This is in particular the case when we work in a Hilbert space H. Then r (H) = 1~1 ( O ) , and the term trace class operator is indeed justified. But there is more. The eigenvalue sequence ()~k) of T 6 ~ l (H) is absolutely summable, and we have LidskiFs trace formula [99], [137, Section 4.2]: oo
trace(T) = Z
)vk.
k=l
2.6.
Every operator T E t~(ll) can be identified with its representing matrix (rhk) via oo
for ( ~ ) ~ l l. k=l
Then o<)
IlZll--
~lrhkl
sup
l~
(x)
and
vl(T)-- Z
sup
I~hkl.
h = l 1~
Infinite matrices for which the right-hand expression is finite were already considered in von Koch's theory of infinite determinants [87] and by Dixon [38, p. 191] when he solved linear systems in infinitely many unknowns. Since 11 has the approximation property, every T = (rhk) 6 r (ll) admits a trace, and we get oo
trace(T) -- Z
rk~,
k=l
as it should be. But even in this comfortable setting, pathologies occur. It was known to Grothendieck [59, Chapter I, p. 171] that the existence of an operator T 6 r (11) with the properties T 2 = O and trace(T) = 1 would yield a negative answer to the approximation problem. However, operators with these 'almost contradictory' properties were constructed only much later [46]. Being nilpotent, T cannot have any eigenvalue different from 0, but its trace is 1. Hence, in this case, Lidskii's trace formula fails. In conclusion, a satisfactory trace theory can only be developed using ideals that are strictly smaller than 9tl. Since 9tl is the smallest Banach ideal, we realize that there is no way to avoid quasi-Banach ideals.
J. D i e s t e l et al.
452
2.7. Grothendieck [59, Chapter II, p. 3] was the first to find non-trivial ideals with a trace. For 0 < p ~< 1, he introduced opdrateurs de puissance p.bme sommable, which are defined by the property that (X3
T -- ~
rk x ; | Yk
with IIx~ll- 1, Ilykll- 1 and t - ( r ~ )
e lp.
k--1
They form the smallest p-Banach ideal ~p; the underlying ideal p-norm being given by ~ p ( T ) = inflltllpll, where the infimum ranges over all admissible representations. The factorization T 9X A > loc Ot> 11 B > Y can be adapted from 2.3, but now we have t -(r~) 6 lp and 4~p(Dt)= IItllpll, Every operator T ~ ~ p ( X ) is compact, and its eigenvalue sequence ()~k) belongs to [r, where 1/r = 1/p - 1/2. Actually, for 0 < p < 1 one can do slightly better: ()~) 6 Ir, p; this is then best possible; [90, pp. 105, 124], [137, p. 163]. Consequently, in order to ensure that ()~) 6 11, we need 0 < p ~< 2/3. This suggests that ~2/3 might be a good candidate for a trace ideal, and indeed, it is. But still a disadvantage arises: if we work on a Hilbert space H, too many operators get lost. In fact, since ~ p ( H ) = ~Sp(H) for 0 < p ~< 1, the component ~2/3(H) is quite small. Again it was Grothendieck [62] who discovered larger ideals with a trace. One of them A
B
consists of all operators admitting a factorization T : X >H > Y such that A is 1nuclear and B is bounded. This ideal coincides with the product t32 o 9 t l , where 132 denotes the Banach ideal of all operators factoring through a Hilbert space H; see 6.3. Now we have t2~2 o r (H) = t ~ 1 (H).
3. Approximation numbers Standard references: [25], [135, Part 3], [137, Chapter 2]. 3.1. The non-increasing rearrangement (r*) of (r~) 6 co is defined as follows: If the sequence has infinitely many terms r~ 7~ 0, then these are ordered such that Irrr(1)l ~> Irrr(2)l > ~ ' " . Otherwise we proceed in the same way as long as possible and fill up with zeros 9
Throughout, ql" stands for the unit circle {( 6 C: Iffl = 1}. In the standard fashion, C ( 2 ) is identified with the Banach space of all 27r-periodic continuous functions f on R. Let E n ( f ) := inf{llf - PII: deg(P) < n}, where the infimum ranges over all trigonometric polynomials P of degree less than n. The numbers En ( f ) form a monotone null sequence whose rate of convergence may serve as a measure of smoothness of the function f . Motivated by these concepts from Fourier analysis and approximation theory, Pietsch [ 128] defined the nth approximation number an(T) "-- inf{ liT - All" A e ~(X, Y), rank(A) < n}
for T e 13(X, Y).
This yields a monotone sequence: IlZll -- al (T) ~> a2(T)/> . . . / > 0. Occasionally, we will write an(T" X --+ Y) instead of an(T).
Operator ideals
453
3.2. Precise computation of approximation numbers is rare; however, for some diagonal operators Dt(~k) -- (r~k) this can be achieved. Let 1 <~ q < p <. ec, 1 / r -- 1 / q - 1 / p and t -- (rk) E [r such that Z'I ~ /72 ~ " ' " ~ 0. Then
an(Dt'lp ~ l q ) -
(L),r r~
k--n In the case p -- q and t - (rk) E C0, we have a n ( D t ' l p ~ lp) -- rn. This result includes an observation of Allekhverdiev [1] who showed that, for compact operators in a Hilbert space, the approximation numbers are just the s-numbers. 3.3.
For 0 < p < ec, the operators T with
Otp(T) "--
<(X3
an(T) p n=l
form the quasi-Banach ideal PAp, which is a natural extension of the Schatten-von Neumann class l~Sp(H). There are operators with a n ( T * ) < a n ( T ) . However, if T is compact, then we have equality. Hence the ideal PAp is symmetric. Since Otp satisfies a quasi-triangle inequality with c -- 21/p, there is an equivalent ideal r-norm, 1/r = 1 4- 1 / p . Unfortunately, apart from the construction described in 2.1, no explicit and aethically pleasing expression of such an r-norm is known. We see from Dt E ~q(lec, ll) 0 t E [q and Dt E PAp(lec, ll) ~ t E [r,p that ~q(Icc, ll) ~ f21p(lec,ll) whenever 0 < r < q ~< 1. Since ~q is the smallest q-Banach ideal, the aforementioned renorming result cannot be improved [ 137, pp. 84 and 110]. 3.4.
It is easy to verify that am+n-1 ( S T ) <~ a m ( S ) a n ( T )
for T E ~ ( X , Y) and S E ~(Y, Z).
Hence we get an analogue of H61der's inequality" if 0 < p, q, r < ec and 1 / r -- 1 / p -Jr-1 / q , then S T E PAr(X, Z)
and
Otr(ST) <~ 21/r 19lp(S) olq(T)
for T E ~.[q(X, Y) and S ~ f21p(Y, Z ) . In fact: PAp o P..lq - Pdr. 3.5. Since Lorentz spaces naturally appear in this context, we introduce the quasiBanach ideal 9.1p,q constituted by all operators T such that ( a n ( T ) ) E [p,q. The following important criterion is adopted from approximation and interpolation theory. An operator T E 13(X, Y) belongs to ~lp,q if and only if it admits a dyadic (p, q)-representation O(3
T -- Z k=0
Ak
with rank(Ak) ~< 2 k,
(3.5a)
454
J. Diestel et al.
where the operators Ak e ~(X, Y) satisfy the condition (2k/PllAkll) ~ [q; see [136] and [ 137, p. 84]. An equivalent ideal quasi-norm is obtained by tglp,q (Z) "=
inf]l (2~/PllA~ll) l lq ll,
the infimum being taken over all possible representations (3.5a). In this case it is indeed a o matter of taste which ideal quasi-norm, Olp,q "-- II(an(. )) I [p,q II or Otp,q, should be considered as more natural; see 2.2. 3.6. Fix T E PAl (X) and choose a dyadic (1,1)-representation as described above. From ]trace(Ak)l ~< 2kllAkl] it follows that OQ
trace(T) "-- Z
trace(Ak)
k=0
yields a continuous trace on PAl. Comparing 9--[1 and ~2/3 shows that PAl (12) is larger than ~2/3(12), whereas PAl (l~, ll) is smaller than ~2/3(1~, 11). 3.7.
A remarkable ideal is given by PA0"--
F]
PAP"
0
Operators in 9,10 are characterized by the property that the sequence of their approximation numbers tends rapidly to zero. If 0 < p < cx~ and 1/r -- 1 + 1/p, then PAr C ~r C PAp. Hence 9-10 coincides with ~0
II
cn(T)"-- inf[ z J ~
II cod(M)
< n}
and the Kolmogorov numbers
dn(T)"= inf{I I
QNTII dim(N)
< n},
respectively. The asymptotic behaviour of the resulting monotone sequences provides measures of compactness of the operator T 6 g ( X , Y). Other quantities with similar properties are the Weyl numbers and the entropy numbers. For further information, the reader is referred to the Gluskin, Gordon and Pajor article [54].
Operator ideals
455
4. Ultraproducts, maximality and trace duality Standard references 9[35, Chapter 6], [106], [135, Chapter 8]. 4.1. Roughly speaking, local theory of Banach spaces deals with those properties that can be expressed in terms of finite dimensional spaces 9The most usual way is to consider quantitative relations that are assumed to hold uniformly for any choice of n elements and/or functionals. The constants involved may or may not depend on n. Local ideal theory generalizes this, transferring such techniques to operators. For a Banach space X the collection of all finite dimensional subspaces is denoted by DIM(X), while COD(X) stands for the collection of all finite codimensional (closed) subspaces. The idea is to gain access to the structure of a given operator T :X ~ Y through an understanding of its 'elementary parts'
QYTJ x
9
M gkx
7">
Y
QY> Q / N
with M E DIM(X) and N E COD(Y). Remarkably, such access is available if we use ultraproduct techniques 9 4.2. Given a family (Xi)i E~ of Banach spaces and any ultrafilter U on the index set IT, all equivalence classes (xi) u = {(x?)" U-limi IIx? - xill = 0} generated by bounded families (xi) such that xi E Xi for i E ~ form a Banach space (Xi) u under the norm II(xi)Ull := U-limi Ilxill. If the family (7~)iE~ with 7) 6 13(Xi, Yi) is bounded, then (Ti)U:(xi) u --+ (Tixi) u defines an operator from (Xi) u into (Yi) u such that ll(7~)ull = U-limi IlTi II. These new objects are referred to as ultraproducts. If Xi -- X, Yi = Y and 7) = T, then we speak of ultrapowers, denoted by X u, y u and T u, respectively. Ultraproducts were introduced by Dacunha-Castelle and Krivine [30]; they constitute a powerful tool in Banach space theory. A few years later, Pietsch [134] showed how to adapt this notion to operators; however, the most important applications to ideal theory are due to Heinrich [63,64]. 4.3. Given an ideal 9.1, we denote by ~[super the collection of operators T for which all ultrapowers T u belong to 9,1. Clearly, 9,1super is again an ideal. It turns out that properties such as closedness, injectivity and surjectivity carry over from 9,1 to 9,1super. We say that 9,1 is ultrapower-stable if 9,1 = 9Asuper. Naturally, 9Asuper is always ultrapower-stable. Let ~Sep denote the closed ideal formed by all operators with separable range 9 Then [super super ~r _..~. So .i~ is ultrapower-stable. Further examples, namely ~ ' l , ~c 0 and ~ZBsuper, will be discussed in 9.9 and 10.9. 4.4. We now choose as our index set the collection of all pairs (M, N) with M E DIM(X) and N 6 COD(Y). Fix an ultrafilter U which contains the final sections {(M, N)" M D M0 and N _c No} associated with M0 E DIM(X) and No E COD(Y). Let X M,N := M, YM,N "-- Y / N and TM,N "-- Q YNT J~l" Define
J~r "x ~ (XM,N)u
with XM,N "--
x
i f x E M,
0
ifx~M,
456
J. Diestel et al.
and QC~ . (YM,N)U w+ H- lim K r y~M N" M,N
Here (YM,N) is some bounded family in Y such that Qruy~M,N -- YM,N. Apply the canonical map K r to pass from Y to Y**. Now Alaoglu's theorem implies the existence of the righthand limit with respect to the weak*-topology. We have I[J~ l] -- 1 and IIQ~ I I - 1. The upshot is the following commutative diagram: ( TM,N ) Lt
(XM,N) u
X
> (YM, N) u
T
> Y
Ky
> Y**
Roughly speaking, we may say that every operator is the ultraproduct of its elementary parts. 4.5. The maximal hull ~2,[max of a quasi-Banach ideal 9A consists of all operators T 13(X, Y) for which
olmax(r) :=sup{ot(QrNTJfvl) 9 M E DIM(X), N ~ COD(Y)} is finite. Since olmax(T) ~ or(T) for T E 9A(X, Y), we have 9A ___9,1max. Clearly, ~max is again a quasi-Banach ideal with respect to Ogmax, and maximality means that 9A = (2[max. If 9.1 is maximal, then 9 A - (9.1dual) dual. We stress that a maximal quasi-Banach ideal is uniquely determined if we just know the values of its ideal quasi-norm for all operators between finite dimensional spaces. Lotz showed in his Lecture Notes [ 106] how the theory of maximal Banach ideals can be derived from the Rdsumd. A quasi-Banach ideal 9A is called ultraproduct-stable if 7) 6 P-d(Xi, Yi) and sup/~(T/) < c~ imply (7~)u ~ 9A((Xi)U, (Yi)U) for all ultrafilters H on arbitrary index sets ~. The central result says that a quasi-Banach ideal is maximal if and only of it is regular and ultraproductstable. Among the ideals considered so far, only t2 and PAp are maximal. The concept of ultrapower-stability makes sense for all ideals, while that of ultraproductstability requires an ideal quasi-norm. Ultraproduct-stable quasi-Banach ideals are all the more ultrapower-stable. The closed ideal ~ shows that the converse implication is not true. 4.6. The real significance of maximality lies in its close ties to trace duality, a topic we broach next. In this context it is natural to restrict the considerations to B a n a c h ideals. Duality of sequence spaces is based on the bilinear form (s, t) " - Y~__ 1 cr~r~ defined for s = (crk) and t -- (r~) in I~. In order to guarantee the convergence of the series we assume st El1. Replacing [1 by ~ 1 (H), there exists a perfectly good working counterpart. In fact, letting (S, T) := trace(ST) whenever S, T E s and S T E ~ 1 (H) yields a duality for
Operator ideals linear subsets of s
457
Note the following striking analogies:
c0* = 11 +-~ .~(H)* = ~ 1 (H) , IT = l ~ ~ ~5~ ( H ) * = ~ ( H ) ,
[p -=- [p. +-~ ~;,p(H)* -- ~;,p.(H),
1 < p <(x~.
What about a generalization to Banach spaces? We know already that 9"tl cannot be used as a substitute for ES1 (H), since there exists no trace on ~ 1 (X) unless X has the approximation property. But local ideal theory offers a possibility to get around this problem by appropriate reduction to a finite dimensional setting. First, for finite dimensional spaces E and F, we easily get the right-hand counterparts of the left-hand relations:
([~)* -- [~ ~ ~ ( E , F)* = 9t, (F, E), (l~)* -- In ~ r
(F, E)* = s
F).
Even more is so: in this case every norm ot on s s F) defined by
E) gives rise to an adjoint norm on
ot* (T " E --+ F ) : = sup{ Itrace(ST)l" ot(S " F ~ E) <<.1 }. Let now 9A be a Banach ideal and T 6 ~ ( X , Y). The construction of the maximal hull suggests that we put
o t * ( T ) ' = s u p { o t * ( Q Y x T J ~ ) 9M 6 D I M ( X ) , N 6COD(Y)}. The operators for which this quantity is finite constitute the adjoint Banach ideal 9.1". Of course, 9A* is maximal. Moreover, 9A** = p..[max. It turns out that a* (T) can be obtained as the infimum of all constants c >~ 0 such that [trace(ASoBT)l <~ c [IAllot(So)lIBll regardless of how we choose the spaces X0, Y0 and the operators B 6 ~(Y, Y0), So E 93(Y0, X0), A ~(X0, X). This characterization is quite flexible, since the finite rank operators A S o B T , So B T A, B T A So and T A So B have the same trace. If 9A and ~3 are Banach ideals, then 9A _c ~3 and fl ~< c o~ imply ~3" _c 93* and ~* ~< c #*. In particular, it follows from 9..1 _c 93** that 93* = 93"**, isometrically. Analogous relations hold within the setting of vollkommene Folgenriiume, in the sense of K~3the [9 1] and Toeplitz. 4.7.
As a basic example, we describe the ideal 9t~* -- ~"t' ml a x 9Grothendieck defined
1-integral, or simply integral, operators T ~ ~ ( X , Y) by the following property. Viewing (Tx, y*) as a bilinear form on X x Y*, he assumed the linear form induced on X | Y* to be continuous with respect to the plus petit | raisonnable [59, Chapter I, pp. 126-127], [60, pp. 9, 12]. This just means that there is a constant c >~ 0 such that [trace(ST) I ~< cllSI] for S 6 ~(Y, X). The 1-integral operators form a Banach ideal 31 with the ideal norm ll (T) := infc. Although the inequality [trace(ST)l ~< c[ISII looks much simpler than the inequality Itrace(ASoBT)[ <~ c[]Allot(So)llBll considered in the previous paragraph both
J. Diestel et al.
458
are equivalent for oe : II 9 II. Hence t3" = ~ 1 . The formula 9t~ -- 13 follows from the fact that I~1 : I ~ for all operators between finite dimensional spaces, and tilT* - - ~1 is then obvious. m-1
m-1
4.8. For every quasi-Banach ideal 9A, we have ~ m a x __ ~ o 9,1 o ~ . This formula can be used to define the maximal hull also in the case when 9.1 is any ideal. Similarly, we get a minimal kernel by letting ~min .__ ~ o ~ o ~ . It turns out that ~min C ~ C ~max i m p l i e s ~min __ ~ m i n and ~2[max -- ~ m a x . If 9,1 is a Banach ideal, t h e n ~ m i n ( x , Y) consists of all operators T E/3(X, Y) that can be written in the form CO
(3O
such that
T -- Z B k T k A k k=l
Z Ilnkll~(Zk)llAkll < k=l
and X Ak> Ek Tk> Fk 8k> y. Here El, E2 . . . . and F1, F2 . . . . are supposed to be finite dimensional. Moreover, P-dmin is a Banach ideal under the ideal norm (3O
~min(T) "-" inf ~
IIB~II~(T~)IIAklI,
k=l
where the infimum ranges over all representations described above. This criterion and its obvious generalization to p-Banach ideals implies that, in analogy to 4.5, a minimal quasiBanach ideal is uniquely determined if we just know the values of its ideal quasi-norm for all operators between finite dimensional spaces. As an example and a counterpart of ~'~Tax - - 21, we get ~min~,l- - r . Surprisingly, the ideals Pip are not only maximal, but also minimal.
5. Summing operators and some of their relatives
Standard references: [35,73,124], [135, Chapter 8]. 5.1. A classical theorem due to Dirichlet (1829) asserts that all rearrangements of a scalar series Y ~ l ~ converge (even to the same sum) if and only if Y ~ l I~k[ < ~ - I n other words, unconditional and absolute convergence of series coincide. Looking at series in a finite dimensional Banach space and replacing the absolute value I" I by the norm I1" II does not change the statement. However, troubles ensue if we pass to the infinite dimensional setting. Still, absolute convergence implies unconditional convergence, but the converse cannot be generally true: look at orthogonal series in 12. Mazur and Orlicz asked in The Scottish Book [108, Problem 122] if every infinite dimensional Banach space contains a series that converges unconditionally, but not absolutely. A positive answer to this question was given by Dvoretzky and Rogers [44]. On the other hand, Orlicz [ 118] proved already in 1929 that every unconditional convergent series ) - ~ - 1 fk in L p satisfies Y~_-i [Ifk II2 < ~ if 1 ~< p ~< 2 and ~ - - 1 IIA IIp < c~ if 2 <~ p < ~ ; see 6.1 and 9.4.
Operator ideals
459
5.2. It turned out to be extremely fruitful to extend such investigations to operators. By analogy with the notion of complete continuity, we may consider operators that improve the convergence of specific sequences or series. Let 0 < p ~< q < cx~. An operator T E ~ ( X , Y) is said to be absolutely (q, p)-summing, or just (q, p)-summing, if it takes weak lp-sequences (xk) of X to strong (absolute) lqsequences (Txk) of Y; precisely: if ((xk,x*)) ~ lp for all x* E X*, then ([[Txkl[) ~ lq. It is important to realize the local character of this concept, which is hidden in this definition. In fact, by the closed graph theorem, T is (q, p)-summing if and only if there exists a constant c ~> 0 such that, for any choice of n E 1~ and Xl . . . . . Xn E X,
1/p IlTxkll q k=l
~< c
sup ](xk,x*)lP IIx*ll~
We put ~q,p(T) : = infc, where the infimum is taken over all constants c ~> 0 with the above property. The class of (q, p)-summing operators, denoted by ~q,p, is an injective and maximal Banach ideal for 1 ~< q < cx~ and a q-Banach ideal for 0 < q < 1. Most important are the ideals ~3p :-- ~]3p,p whose members will be called p-summing. In view of the preliminary remarks, we stress that an operator is 1-summing, or just summing, if and only if unconditionally converging series are carried to absolutely converging series. If 1/p - 1/q <~ 1/r - 1/s and p ~< r, then ~q,p c ~s,r" The 1-parameter scale ~3p increases strictly on [1, cx~), but is constant on (0, 1); see [109]. Absolutely 1-summing operators were introduced by Grothendieck as applications prdintdgrales (semi-intdgrales ?~) droite [59, pp. 160-161], [60, p. 33]. He also considered operators T 6 ~ ( X , Y) that admit a factorization of the form T : X --+ C (K) --+ H --+ Y. In a communication addressed to the ICM in Moscow (1966), Pietsch identified these applications prohilbertiennes gauches, also called applications ayant la propridtd de prolongement hilbertien, as absolutely 2-summing operators [60, p. 36]. Most important:/t'2 proved to be the natural ideal norm on ~]32. At the same congress, Mityagin and Pelczyfiski [ 114] defined the general concept of a absolutely (q, p)-summing operator. 5.3. The central part of the theory of p-summing operators was developed by Pietsch [130]. In particular, he established a criterion for an operator to be p-summing, which has become a standard result in Banach space theory. Its proof can be found in many monographs: [35, p. 44], [73, p. 55], [135, p. 232], [188, p. 47], [192, p. 203] and in the Johnson and Lindenstrauss article [80]. Recall that Bx,, the closed unit ball of X*, is a compact Hausdorff space in the weak* topology. So, thanks to the Riesz representation theorem, positive linear functionals on C(Bx,) are just given by regular Borel measures. Pietsch's criterion may be formulated as follows.
J. Diestel et al.
460
An operator T 6 s Y) is p-summing if and only if there exists a constant c >/0 and a (regular) Borel probability # on B x , such that
IlZxll <~ c
for all x E X.
](x,x*)[ p d/ztx*)
(5.3a)
X*
In this case, rrp(T) is the infimum, and even the minimum, of all constants c >~ 0 for which such a measure can be found. In order to make the support of/z small, we may pass to any weak* compact subset W of B x , such that Ilx[I - SUpx,cW [(x,x*)l. For example, if X is C ( K ) , then the set of all Dirac measures 6~" f ~ f ( ~ ) with ~ E K has this property. For 1 ~< p < oo, we get the factorization T
X
J
>Y
> loc(Br*) (5.3b)
C(Bx,)
I
> L p ( B x , , Ix)
Indeed, let A take x to (x, .), viewed as a continuous function on B x , , while J sends y to ((y, y*)), viewed as a bounded family on Br,. Moreover, consider the canonical map I " C ( B x , ) --+ L p ( B x , , lz). Then inequality (5.3a) means that []JTxl[ <~ c [llAxlLpl] for all x E X. Hence, I A x w-~ J T x defines an operator on a linear subset of L p ( B x , , lz) and, by the metric extension property of loo(Br,), we can find B such that J T -- B I A and Ilnll ~< c. From (5.3a) and (5.3b) we easily conclude that p-summing operators are completely continuous and weakly compact: ~3p c ~ and 213p ___~[i'. 5.4.
The kth Rademacher function rk takes alternately the values + 1 and - 1 on the
dyadic intervals of length 1/2 ~. That is, rk(t) :-- ( - 1 ) j-1 whenever t 6 A~j) .= [_St_, ) j - 1 ffj and j = 1 . . . . . 2 k. A particular case of Khintchine's inequality says that, for ~l . . . . . ~n E ]~, r/
I~kl 2
Z ~ k r l ~ ( t ) dt.
~< ~/2
(5.4a)
k=l
k=l
The constant ~/~ is best possible [184] and [96]. Moreover, since (r~) is an orthonormal system in L2[0, 1), we have Parseval's equality
I~kl 2 k=l
-
(So
2
n
k=l
~kr~(t)
d')
1/2
(5.4b)
Observe that p ' t w-, (r~(t)) defines a Borel measurable map from [0, 1) into the closed unit ball of l ~ - l~, equipped with the weak* topology. Look at the image of Lebesgue's measure under p to see that (5.4a) and (5.4b) are in fact special cases of (5.3a) applied
Operator ideals
461
to the embedding l : l l --+ 12. Hence this map is 1-summing, and afortiori 2-summing. We even have rrl (I : ll --+ 12) = v/2 and ~r2 (I : ll --+ 12) -- 1. These examples show that
q3p ~ J~. 5.5. Let 1 ~< p, q, r < c~ and 1 / r - - 1/p + 1/q. Then HOMer's inequality for summing operators asserts that
ST E 9i3r(X, Z)
and
rtr(ST) <<,rrp(S) rtq(T),
whenever T E q3q(X, Y) and S E ~
ST E 9tl (X, Z)
and
Z). In the case p = q = 2 we even have
vl (ST) <~/t'2(S) rt2(T).
We mention that the composition formula gtp o gtp C ~D"o ~ = ~ yields an ideal theoretic proof of the Dvoretzky-Rogers theorem discussed in 5.1: Pp -- F. 5.6. In the rest of this section, # will denote a probability measure 9 Sometimes # is d e f n e d on an abstract measurable space M; however, should the underlying set be a compact Hausdorff space K, then we a s s u m e / z to be a regular Borel probability. The canonical maps I : L ~ ( M , / z ) --~ Lp(M, #) and I : C ( K ) --~ Lp(K, #) belong to ~13p. Moreover, in both cases, the p-summing norms are equal to 1. As discrete counterparts, we consider diagonal operators D t : l ~ -+ lp induced by sequences t E [p. Then ltp(Dt) = IIDtll = Iltllpll. 5.7.
Let 1 ~< p < cx~. An operator T : X --+ Y is p-nuclear if there exists a factorization
of the form T 9X a ~ l~ Dt> lp B ~ Y with t E [p. These operators form the Banach ideal r the ideal norm being given by vp(T) : = infllAII IIDt II IIBll, where the infimum extends over all admissible factorizations. In the limiting case p - cx~, we use co instead of lp. Operators T E 91p(X, Y) are characterized by the property that they admit a p-nuclear
representation oo
T --
rk x k | Yk, k--1
:~ . . ~ E S :~ and Yl, Y 2 , . . . :r II ~ 1 and x~,x 2, E Y satisfy the conditions sup~ Ilx~ II((yk, y*)) I lp, II ~< Ily*ll for all y* 6 Y*. In addition, (rk) E lp. The connection between p-summing and p-nuclear operators is given by trace duality: 9tp - ~3p,. To preserve this formula in the limiting case p - 1, we let g t ~ " - ,!3. In this way the correspondence [p +-~ ~13p extends to p -- cx~. It is clear that the counterpart of c0 in such a setting is ~3, the ideal of completely continuous operators. In the language of tensor products, 2-nuclear operators were first introduced by Saphar [163]. Later, Chevet [26] and Saphar [164] generalized this concept to exponents 1 ~< p < cx~.
where
462
J. Diestel et al.
5.8. Let 1 ~< p ~< cx~. An operator T e s Y) is said to be strictly p-integral if it factors through a map I : C ( K ) ~ L p ( K , lz). These operators form a Banach ideal. As an example we mention the embedding map from l l into co, which can be factorized in the form l "/l R*> L ~ [ 0 , 1 ) I > LI[0,1) R > c0, where R ' f ~ - + (f o f ( t ) r k ( t ) d t ) . M o r e important are the p-integral operators that are supposed to admit a factorization of the following form: X
C(K)
T
Ky
>Y
I
> Y**
> L p ( K , #)
The lower arrow may be replaced by L ~ ( M , #) I> L p ( M , / ~ ) , where (M,/z) is an abstract probability space. This definition yields the Banach ideal 2p equipped with the ideal norm tp(T) := inf IIBII ]IA[[; the infimum ranges over all decompositions as above. If p = 1, then we get the 1-integral operators, which were already considered in 4.7. The formulas ~max __ ,'~p and ~min = q~j~p reflect the philosophy of passing from finite or infinite - -p --p sums to integrals. Sophisticated examples show that when p ~ 2 there exist p-integral operators that fail to be strictly p-integral [ 153]. Of course, the reader may wonder why we prefer the more p --- ~.~ p , . complicated concept. The answer is given by trace duality: ~3p - - "~ Jp, and ~J '~'* Persson and Pietsch [ 126] defined strictly p-integral operators. Later, Pietsch [ 132] observed that regularity is indispensable for a smooth theory. So, following Grothendieck's pattern for p = 1, the natural embedding Ky was incorporated in the definition of operators T ~ 3 p ( X , Y). ~inj
5.9. Diagram (5.3b) implies that ~Jp -- ~]3p, and every p-summing operator T" C ( K ) --+ Y is even strictly p-integral. The following example due to Petczyfiski [ 123] indicates the fruitful interplay between ideal theory and harmonic analysis. Given any subset A of Z, we denote by C a (~1") and L pA ('T) the subspaces of C ('I[') and L p (T), respectively, that are spanned by all characters e int with n E A. Then the p-summing embedding map I" cA (T) ~ LpA (T) is p-integral if and only if L A p ('~') is complemented in L p ('IF). Since, for p -~ 2, there are A's failing this property, ~ p is properly contained in glp. 5.10. The case p - 2 is something special. The Banach ideals of strictly 2-integral, 2-integral and 2-summing operators coincide isometrically. Moreover, Jr2 (T) = v2 (T) for 2-nuclear operators T. The formula ~]3~ = ~]32 can be rephrased by saying that ~]32 is selfadjoint. The sole defect is a missing theorem of Schauder type. Indeed, the embedding from 11 into 12 is 1-summing though neither its predual nor its dual is p-summing for any 0
Operator ideals
463
T 6 ~ 2 ( X , Y). This is an immediate consequence of the fact that the spaces L ~ ( M , #), which appear in strictly 2-integral factorizations, have the metric extension property 9 Next, we show that the ideal of 2-summing operators not only has a beautiful theory; it is also extremely useful. Garling and Gordon [51] observed that zr2(E) = ~ for every n-dimensional Banach space E. An elementary proof of this fact is due to Kwapiefi; see [135, p. 385]. Using a strictly 2-integral factorization of the identity map yields an operator T such that
IE 9E T ) 12n T - I ) E and 11T1111T - i l l ~< x/-d. In this way, we have obtained John's theorem [77], which states that the Banach-Mazur distance d (E, l~) is at most v/ft. Moreover, if E is a subspace of X, then the Hahn-Banach theorem for 2-summing operators provides us with a projection P 6 s E) such that IIPI] ~< nr2(P) ~< ~ (Kadets-Snobar theorem) [81]. K6nig [89] observed that the square ~ 2 o ~ 2 admits a trace, which can be obtained as the sum of the eigenvalues of the operators under consideration; see also [137, p. 182]. We stress that this ideal coincides isomorphically with the product ~ 2 o r which was already discussed in 2.7. 5.11. One of the basic problems is to determine the Hilbert space components of a given ideal 9A. For example, we have ~ p ( H ) = ~ 2 ( H ) if 0 ~< p < ~ ; see [122]. The situation is more complicated for (q, p)-summing operators with p < q. If 1/p - 1/q >~ 1/2, then ~ q , p ( H ) = ,~(H). The case 1/p - 1/q < 1/2 splits into two subcases: On the one hand, ~q,p(H) -- ~13r,2(H ) = ~r(H) for 1/r - 1/2 = 1/p - 1/q and 0 < p ~< 2. On the other hand, it can be shown that ~q,p(H) = ~2q/p,q(H) whenever 2 < p < cxz. Some of these results are proved by interpolation theory [138]. However, the inclusion ~q,p(H) C ~J2q/p,q(H) was obtained by considering random matrices; [9] and [10]. Use of probabilistic techniques to prove existence results has become a staple in the arsenal of functional analysis. 5.12. Ideal quasi-norms ot and fl defined on different ideals 9A and ~3 cannot be compared in general. The situation improves when the considerations are restricted to finite rank operators. Then we always find constants cn ~> 1, which may also depend on a and fl, such that
or(T) <~cn fl(T)
whenever rank(T) ~< n.
It is of interest to know best possible values of the numbers Cn, because their asymptotic behaviour may be used to measure the distance between a and ft. Here is a short list of known and useful relations from [24], which hold for operators with rank(T) ~< n:
Otp(T) <~n llp-1/q Otq(T)
i f 0 < p < q ~< cx~,
r
if0
~ n 1/p []T]]
1,
tp(T) ~ n max{1/p'l/2} []T[]
i f l ~< p ~< c~,
tp(T) ~ n jl/p-1/21 lrp(T)
i f l ~< p ~< cx~.
J. Diestel et al.
464
It is worthwhile to point out some analogies with other fields. For example, if x -- (~k) has at most n non-zero coordinates and 0 < p < q ~< o~, then H61der's inequality implies that [[xllp[I <~ n 1/p-1/q [[xllq[I. For trigonometric polynomials P with deg(P) ~< n and 1 ~< q < p <~ c~, NikolskiY's inequality says that [[PILp[[ <<.Cp,q n l/q-lIp [[PILq [[. As in the theory of function spaces such results can be used to prove embedding theorems. Applying the inequality zr2(Ak) ~< 2k/2[[A~[[ to a dyadic (2,1)-decomposition (3.5a) gives
oo oo ~2(T) ~ Z ~2(Ak) ~< Z 2k/2llAkll" k=0 k=0 Hence 9---[2,1 ~ ~]~2" This result is the best possible. Indeed, it follows from Dt 9-12,q(12, 11) <:~ t E [1,q and Dt 6 ~]32 (/2, 11) <:~ t 6 [1 that 9---[2,1+~ ~ ~ 2 if e > 0; see [137, pp. 64 and 110]. Similarly, ~p(T) ~ n 1/p IIT[[ for rank(T) ~< n and 0 < p ~< 1 yields 9ap _c ~p. 5.13. An ideal quasi-norm a is only defined for operators contained in the underlying ideal 9,1. Trivially, we could put or(T) := (x~ whenever T ~ 9,1. But this is quite unsatisfactory. Petczyfiski [ 125] made a much better proposal, one which we entertain presently. Recall that the 2-summing norm Jr2(T) is the least constant c ~> 0 such that
1/2 ~< c k=l
sup
.(Xk,X*). 2
IIx*II~<1 k--1
for n E N and xl . . . . . Xn E X . If T r ~ 2 ( X , Y), then we can find no constant that works for all n. However, for a fixed n we can always find some c ~> 0 such that the above inequality obtains so long as we consider no more than n members of X. Let g~n)(T) denote the infimum of such constants. What results is a non-decreasing sequence (Jr~n) (T))
with ~ n ) (T) ~< ~
IITll, whose growth measures just how non-2-summing the operator T is. All identity maps of infinite dimensional spaces X are examples of the worst case: g~n) (X) -- X/rn. On the other hand, for the summation operator, ~r~n) ( r ' l l --+ 1~) behaves like log n. The most beautiful and powerful theorem along this line was proved by Tomczak-Jaegermann [ 187]: n'2(T) ~< ~/2jr~n)(T)
wheneverrank(T) ~< n.
6. Lp-factorable operators Standard references: [35, Chapter 9], [135, Chapter 19], [188, w 6.1. So far we have mainly dealt with small ideals; that is, ideals 9,1 for which Ix ~ 9,1 implies dim(X) < oo. Whenever A -- F, good spectral properties as known from the RieszSchauder theory may be expected. Now we turn to the study of large ideals. This means
Operator ideals
465
that A contains infinite dimensional spaces too. Here classification of Banach spaces is the major objective. For example, q32, l gives P2,1, the class of spaces with the Orlicz property; see 5.1. 6.2.
Let 1 <~ p < c~. An operator T : X --+ Y is strictly L p-factorable if there exists A
B
a factorization of the form T : X > Lp (M, #) > Y, where (M, #) is some measure space. Among others, we may take a set IT equipped with the counting measure. From our experience with p-integral operators we know that using Kr T instead of T is advantageous. This leads to the definition of an Lp-factorable operator: K r T :X
a
B
Lp(M, Iz) > Y**. We get the maximal Banach ideal 13p equipped with the ideal norm ~,p(T) :---- infllBll IIAII, where once more the infimum extends over all factorizations as above. If p = cx~, then we factor through a suitable space C(K), which may be replaced by L ~ ( M , #) or l~(]l). However, the ideal of strictly L~-factorable operators is smaller than that of strictly C-factorable operators. Note that 13~ -- 3 ~ . The maximality of 13p is by no means obvious. Deep representation theorems from the theory of Banach lattices, due in the main to Bohnenblust [13] and Kakutani [83,84] in tandem with ultraproduct techniques are called for; see also [95, pp. 131-139]. Whereas it is readily seen that .f~dual ~_p -- 13p,, it requires some doing to verify that 13p, -~.~dual p o g3p,. Moreover, if 1 < p < oc, then t32 ___13p. This follows from the fact that 12 is isomorphic to a complemented subspace of Lp[0, 1]. The theory of Lp-factorable operators is essentially due to Kwapiefi [93]. 6.3. Most important is the case p -- 2. Then L2 (M, #) can be replaced by an abstract Hilbert space, and there is no need of the natural embedding Kr. That is, L2-factorable operators are even strictly L2-factorable. Operators admitting a factorization T : X ~ H --+ Y were already investigated in Grothendieck's R~sum~. Lindenstrauss and Petczyfiski [ 104] referred to such operators as Hilbertian. Based on their work, Kwapiefi showed that T E 13(X, Y) is Lz-factorable if and only if there exists a constant c ~> 0 such that 1/2
~-~ trhkTXk h= l
k=l
C
IlXk II2 k-1
for all n E 1~, unitary (n, n)-matrices (O'hk) and xl . . . . . Xn E X. In this case, ~,2(T) -- infc, where the infimum is taken over all constants c ~> 0 with the above property. The ideal 132 is injective, surjective and symmetric. 6.4. The class kp consists of all Banach spaces X for which X** is isomorphic to a complemented subspace of some Lp(M, #). Lindenstrauss [103] found a space in t-1 that has no complemented copy in any L I(M, #). Bourgain and Delbaen [ 16] have constructed an infinite dimensional Banach space X E k ~ with the Schur property. Hence X admits no complemented copy in any C(K). So, for p = 1 and p -- cx~, passage from X to X** is really needed. Of course, for 1 < p < cxz, all spaces in Lp are reflexive, and we have 12 c Lp. In particular, 12 consists of those Banach spaces that are isomorphic to a Hilbert space.
J. Diestel et al.
466
In their seminal papers, Lindenstrauss, Petczyfiski and Rosenthal [ 104,105] studied the class s which may be slightly smaller. More precisely, we have 12p U/~2 : Lp and s N /~2 : F for 1 < p < cx~ and p ~- 2, but/~1 : L 1 , / ~ 2 : L2 and s = L~. 6.5.
An application of trace duality leads to s
o r.,_ dpu a l C_
"~p
. There is also an in-
dual - 1
verse formula, s -- 2p o [~13p ] . By definition of quotient ideals, this means that T e ~ ( X , Y) belongs to ,P_,p(X, Y) if and only if TA e "3p(Xo, Y) regardless of how we choose f t inj
dual - 1
X0 and A e ~dual(Xo.,_ p , X). Passing to the injective hulls, we get ,,p -- ~13p o [gtp ] In the setting of spaces, we end up with the following characterization: a Banach space X is isomorphic to a subspace of some Lp(M, #) if and only if, for any choice of the Banach space X0, every operator T e s X) with a p-summing dual is itself p-summing. When p = 2, this yields an isomorphic characterization of inner product spaces, which was the starting point of Kwapiefi's investigations on Lp-factorization [92]. In 5.10 we have stated ~13p'S lack of symmetry. However, the above criterion is based on just this lack! 6.6.
We conclude this section by presenting another type of factorization. An operator A
B
s Y) is called lattice-factorable if Kr" T can be written in the form Kr T : X >L > Y**, where L is a suitable Banach lattice. Passing from L to L**, we may even suppose that the lattice is order complete. The class of these operators, denoted by s is a maximal Banach ideal with respect to the ideal norm Lust(T) := inf IIBII IIA II. As usual, the infimum ranges over all possible decompositions as described above. This theory goes back to Gordon and Lewis [56]. Solving a problem raised by Grothendieck [60, p. 72], they proved that 1-summing operators need not factor through L1. That its, ~31 ~ s A consequence of this result will be discussed in 12.10. On the other hand, we have gtl o s C ~1. 6.7. Banach spaces belonging to Lust are said to have local unconditional structure in the sense of Gordon and Lewis. This term is justified by the fact that, from the local point of view, those spaces behave just like Banach spaces with an unconditional basis. The most prominent examples of spaces X ~ Lust are the disk algebra [124, p. 25] and the Schattenvon Neumann classes ~5p(12) with p r 2; see [56] and [98].
7. Grothendieck's theorem Standard references: [32, Section 14], [35], [135, Chapter 8], [148, Chapter 5]. 7.1. Henceforth, Lp denotes any member of Lp. In particular, we may think of a classical space Lp(M, #). By defnition, the C(K)'s are included in t_~. Let 9,1 and ~3 be maximal quasi-Banach ideals. Then for all Lp and Lq, we have P,.l(Lp, Lq) c_ ~ ( L p , Lq) if and only if P,.[(lp, lq) c ~ ( l p , lq). This, in turn, occurs pren cisely when fl(T) <, cot(T) holds for T e ,g,(lnp, lq) and n - 1,2 . . . . . where c ~> 1 only
7.2.
depends on ot and ft.
Operator ideals
467
By far, the most striking result of the theory of operator ideals is Grothendieck's theorem, which he obtained as a corollary of his thdorkme fondamental de la thdorie mdtrique de produits tensoriels [60, p. 59]. It can be stated in various ways. The usual 7.3.
form is
t2(L1, L2) = gj31(L1, L2). We prefer, however, to work with inequalities, since this point of view yields additional information about the involved constants: (G1)
~ l ( T ' l 7 --~ l~) <~cllT'l 7 ~
z ll,
(G2) (G3)
Vl ( r "l~ --+ l ~ ) <~ c k 2 ( r "l~ ~ l ~ ) ,
(G4)
,~"a' o,~2(T'l~ -, zg) ~ ~IITZL -+ 1711.
Trace duality tells us that (G1)C~(G2) and (G3)CC,(G4). In order to prove (G3)=~(G1), just embed l~ almost isometrically into some 1u; and (G1)=~(G3) is straightforward. All inequalities hold with the same constant c > 0, and the infimum is denoted by KG; different values obtained in the real and complex case will be distinguished by the superscripts R and C. The exact Grothendieck constants are still unknown; see [32, pp. 174 and 269]: 1.27323... < K~ < 1.40492... < 1.57079... < K~ < 1.78222 . . . . Further we have the estimates
(gl) (g2)
~ ( T ' l ~ ~ l~) ~<~11T "l~ ~ ~l"
The later can be restated as the little Grothendieck theorem:
~(L~,L2)--~2(L~,L2). Obviously, (G1) implies (gj). The equivalence (gl) r rr2(S) for S E/3(1~) and the formula
~2(T)- sup{rrz(TA)" [IA'l~
(g2) follows easily from N2(S*) =
~ E[ ~ 1, n ~ 1~l,
which holds, in particular, for all operators T : E --+ F between finite dimensional spaces. c are known: KgR _ ~ / z r / 2 In this case, the best possible constants, denoted by KgR and Kg, 1.2344... and Kgc _ v/T/zr _ 1.1284 . . . . Finally, we deduce from (G4) that (H)
X2(T'I~ ~ iT) ~ c[IT.z ~ ~ zT[[,
2 ~ KG ~< K 2KH ; see [32, and the smallest constant will be denoted by KH. T h e n Kg p. 173]. It would be interesting to know whether KH is less than KG.
J. Diestel et al.
468
7.4. Grothendieck's theorem can also be stated in the form ,~2 o ,~1 ~__ ~1~1" We even have ,~2 : ~ 1 o ~11 . In particular, s X) -- ~131(11, X) implies that X e / 2 . On the other hand, let ~5'Eb " - s o g t 1. Then X e GTh if and only if s = ~31 (X,/2). In this case X is said to satisfy Grothendieck's theorem. Kislyakov [86] and Pisier [143], [148, Chapter 6] showed that this property is shared by all quotients of L1 modulo a reflexive subspace N. If, moreover, N is infinite dimensional, then L 1/N ~ L1. Hence L1 is properly included in GTh, and we have obtained a new and interesting class of Banach spaces. Bourgain [ 15] proved one of the most spectacular results in this area: L 1 ('~) / H1 (T) E GTh. 7.5. Grothendieck's theorem has many applications in Banach space geometry. We present but two of them. A classical result asserts that every Banach space X is isometric to a subspace of some C (K) and to a quotient of some l l (]I). Interchanging the roles, Grothendieck asked: What spaces X are simultaneously quotients of a C ( K ) and subspaces of an L1, say via the surjection Q : L ~ --+ X and the injection J : X --+ L 1? Then J I x Q E ~ ( L ~ , L 1) = ~ 2 ( L ~ , L1). Injectivity and surjectivity of s imply that Ix e s So X is isomorphic to a Hilbert space. Let (x~) be an unconditional normalized basis in l l. Then every x E l l admits a unique representation x -- ~ = 1 ~x~ such that the right-hand sum converges unconditionally. From l l e P2,1 it follows that x w-> (~k) takes l l into 12. By Grothendieck's theorem, this map is 1-summing. Hence ( ~ ) e ll. Conversely, every sequence ( ~ ) E l l defines an element x -- ~ - 1 ~kx~ in 11. So all unconditional normalized bases of l l are equivalent to the standard basis.
8. Concrete operators 8.1. Determining precisely when a concrete operator of prescribed form belongs to a given ideal often provides considerable enlightenment about the operator, its domain and codomain, and the ideal. For example, every continuous kernel K defined on the unit square induces a nuclear integral operator g(t) ~ fd K (s, t)g(t) dt from C[0, 1] to itself, but may fail to be nuclear as an operator from L2[0, 1] to itself. We are going to present a small but tasty selection of concrete operators, which make the subject of ideal theory more attractive. 8.2. The simplest examples are identity maps I'lnp --+ lq. These operators belong to t/ every quasi-Banach ideal 9.1 and o t ( I ' l p/7 ~ lq) is well-defined for the underlying ideal quasi-norm ct. The asymptotic behaviour of these quantities as n --+ oo offers valuable information, in particular, when we compare different quasi-Banach ideals. Here are two typical examples:
Vl
.
(I " lp --+ l
q) = {" 1-1/p+l/q 11
rtp(I " lq, ---->lq) ~ (n logn) 1/q
i f l <<.q <~ p <<.cx:, i f l <<.p <<.q <<.c~. ifl~
469
O p e r a t o r ideals
8.3. Next, consider diagonal operators Dt : (~k) w-~ (rk~k) generated by a sequence t = (rk) E c0. If 1 ~< q ~< p ~< oo, then we need t E lr with 1/r = 1/q - 1/p in order to guarantee that Dt maps lp into lq. Of special interest is the case DZ'(~k) w-~ (k-X~k) with ;~ > 0. The limit order )~(9A, p, q) of an ideal 9A is defined to be the infimum of all )~ > 0 such that D ~ E 9d(lp, lq). Note that Dt E Pd(lp, lq) whenever t E lr, where 1/r > )~(gA, p, q). If 9A is a quasi-Banach ideal with respect to a , then n )~(9A, p, q) - inf{)~ > 0: ot(I'Ipn --+ lq) <. cn z for some c > 0}.
As an example, we give the limit order )~(~13r , p, q). Viewed as a function of 1 / p and 1/q, it is illustrated in the following diagrams [135, pp. 312-313]" l~
2
1/q
l/q
8 9~§
l/r
~1 - ~1 + ~ 1 1/2
I,
l/r I
p*
p*
1/p
lip
D,
l/r*
l/r*
_ )~
"
1 _jr_ ( l / r * - l l p ) ( l l q - 1 / r ) l12-11r
r
8.4. Closely related to diagonal operators are embedding maps of function spaces. As indicated in 1.10, properties of I ' W 2 (s ~ L p ($72) can be used in the theory of partial differential operators. Given a b o u n d e d domain s in R n with a smooth boundary, )~ > 0 and 1 ~< p < c~, then Wpz (S-2) denotes the Sobolev space if )~ E N and the SlobodetskiY
space if )~ r N. If p - c~ and 0 < )~ < 1, then we mean by W)(S-2) the space of all H61derLipschitz continuous functions: CZ(s The limit order )~se(PA, p, q) is d e f n e d to be the infimum of all )~ > 0 such that the e m b e d d i n g map from Wpz (S-2) into Lq (S-2), which exists whenever )~/n > m a x { 1 / p - 1/q, 0}, belongs to 9A. It was shown by K6nig [88] that, for quasi-Banach ideals, both limit orders are connected as follows:
)~se(PA, p, q) - )~(9A, p, q) + 1/p - 1/q. R o u g h l y speaking, the inclusion I E f21(W~p(X-2),Lq(S-2)) and the asymptotic estimate
ot(I "lpn __+ lq)n < c n ~- 1/p+ 1/q for n - 1, 2,. .. are 'almost' equivalent. 8.5.
Let (M, # ) and (N, v) be ~r-finite measure spaces. If 1 ~< p, q < ec, then a Hillex N --+ C such that
Tamarkin kernel is a / z x v-measurable function K ' M 1/p (fNIfM
IK(s, t)lq d v ( t ) l P / q d # ( s ) )
< (X).
470
J. Diestel et al.
Then
TK "g(t) --+ fN K(s, t)g(t) dr(t) defines a p-summing operator from Lq.(N, v) into Lp(M, #); for p -- q -- 2, we get the classical Hilbert-Schmidt operators. 8.6.
The convolution operators
Tf "g(t) ---->f 9 g(s) -- ~
mf+
f (s - t)g(t) dt
are especially important examples of integral operators 9If f 6 C (qI'), then it follows from T f 9L l (7~) r f> C (7~) I > L1 (T) that Tf E 9tl (L1 (T)) C ~2(L1 ('IF)). For f E Cz(T) we can do better. Approximation by trigonometric polynomials gives En ( f ) ~< c n -z, which in turn implies aen(Tf" LI(T) -+ C(T)) ~ c n - A . Hence Tf E 9132 o 9.lp,c~(Ll(T)) with 1 / p - - )~. Finally, employing the theory of eigenvalue distributions [90, pp. 213-214], [137, p. 310], we arrive at a well-known result of Bernstein [ 11 ]" if f E C z (~) and 1/2 < ~. < 1, then f has an absolutely and uniformly convergent Fourier series.
8.7. Let 0" D -+ D be a analytic function on the open unit disk D. Then the composition operator
Co " f (w ) --> f o O(z) "- f (d#(z)) maps every Hardy space Hp (D) into itself. Shapiro and Taylor [ 180] were the first to find criteria that guarantee that a composition operator belongs to a given ideal. For instance, if (1 -14~1) - 1 has integrable boundary values and 1 ~< p < oo, then C 4) ~ O43p(Hp(D)). In the case 2 ~< p < oo and only in this case the condition above is also necessary. We know that Ce maps Hp (D) into H ~ (ID) if and only if ~b(II)) is contained in a disk r D with radius r less than 1. An equivalent property is the compactness of C~ : H ~ (D) --+ H ~ (D). Moreover, under this assumption, C4):Hp(D ) ~ Hp(D) is nuclear and even a member of 9.1o. Using Nevanlinna's counting function
N4)(w)'--
Z
l~
~1
for w --# 0 (0)
ZEr (tO)
Shapiro [ 178] answered the question of how much r has to compress ID into itself in order to insure that C~ compresses bounded subsets of Hp (I3) into relatively compact ones: Cr ~ ~(Hp(]D)) r
lim Nr Iwl~ 1 log
1
Iwl
=0.
Operator ideals
471
We stress that the right-hand condition is independent of 1 ~< p < ~ . In the case of H2 (D), this criterion was extended by Luecking and Zhu [107]: dA(w)
Ccb E ~r(H2(D)) ~ fD[Ncb(t~ r/2
(1 - Iwl2) 2
log ~
<ex~
if0
where dA(w) denotes the area measure. The symbol 4~ of a compact composition operator C 4, always has a unique fixed point z0 6 D such that I~P'(z0)[ < 1. If 4~'(z0) r 0, then the spectrum consists of 0 and a rapidly decreasing sequence of simple eigenvalues: 1, 4~'(z0), 4r (z0) 2 . . . . . Otherwise, we only get {0, 1}. Composition operators have been investigated in other spaces of analytic functions. In many respects, Bergman spaces Ap(D) can be used to replace Hardy spaces and, sometimes, they behave better [192, p. 23]. The most general result obtained so far is due to Domenig [39]. He constructed for every analytic function 4) :D --+ D and 1 ~< p, q < a sequence t 6 [~ such that, for every maximal Banach ideal 9,1, the statements C~ 9A(Ap(D), Aq(D)) and Dt c 9,1(lp, lq) are equivalent. For more information about composition operators, we refer to [28,29,76,177] and [ 179]. 8.8. Specific components of Banach ideals provide a useful arsenal of non-classical Banach spaces. The most prominent examples are the Schatten-von Neumann spaces ~r(/2), which fail to have local unconditional structure for r ~ 2; see 6.7. The components ~r(lp, lq) were also investigated; see [100,149] and [172].
9. Rademacher type and cotype Standard references: [35, Chapter 11], [113, Part II], [135, Chapter 21], [139, Chapter 4], [148, Chapter 3], [175], [188, w 25]. 9.1. Let (rk) be the sequence of Rademacher functions defined in 5.4, and assume that 1 ~< p ~< 2 ~< q < oo. Then Parseval's equality (5.4b) implies that
I~klq
dt
k=l
~<
k=l
I~k]p k=l
t
Just as in the theory of summing operators, we may ask what happens when the scalars ~l . . . . . ~n E K are replaced by arbitrary elements Xl . . . . . xn of a Banach space X and the absolute value l-I by the norm I[. 11. The upper and the lower part of the preceding inequality are now used separately to introduce new classes of operators. An operator T 6 ~ ( X , Y) is said to have Rademacher type p if there exists a constant c/> 0 such that, for any choice of n 6 1~ and Xl . . . . . xn E X,
• Is0
k--1
Txkrk (t)
2)j2 dt
~< c
[Ixk[Ip k=l
t
.
(9. la)
J. Diestel et al.
472
We put p'rp(T) "= infc, where the infimum is taken over all constants c ~> 0 with the above property. The class of these operators, denoted by 9lff;p, is an injective, surjective and maximal Banach ideal. On the other hand, the inequality 2
IITxk IIq
~< c
k=l
k--1
1/2
d'1
~xkrk(t)
(9.1b)
yields operators of Rademacher cotype q. An injective and maximal Banach ideal, denoted by 9lffq, is obtained in this way; the ideal norm prq being defined in an analogous fashion. We readily see that 9q;'Ep~ C ~C~p2 if 1 ~< P2 < pl ~< 2 and ~C~ql C ~C~q2 if 2 ~< qi < q2 < oo. Clearly, 9~ff71 = .!3, and it is convenient to put 9~ff~ : - s 9.2. The Kahane inequality is obtained by extending the classical Khintchine inequality to the vector-valued case" for 0 < s < ec, there are constants As and Bs such that, regardless of the Banach space X, the number n 6 N and the elements x l . . . . . Xn ~ X, we have
As
(s0
2
xkrk(t) k=l
1/2
d't
<,
(s0 (s0
xkrk(t)
k=l
<~ Bs
s),s II ) dt
2
~_~xkrk(t) k=l
1/2
dt
Consequently, the quadratic average in the above definitions can be replaced by any Lsaverage. Sometimes, it is advantageous to use the same exponents on both sides of (9. l a) and (9. l b), respectively. Equivalent, but non-normalized ideal norms are produced in this way on 9l~gp and 9lffq. 9.3. Our definitions can be repeated with random variables other than the Rademacher functions. For instance, take a sequence of independent standard Gaussian random variables defined on some probability space, and use in (9.1a) and (9.1b) quadratic Gaussian averages. Then the preceding strategy leads to the Banach ideals ~ g p and ~Sffq. The operators so obtained are said to be of Gaussian type p and Gaussian cotype q, respectively. As for type, there is nothing new: ~5ffp = 9l~gp, with equivalent norms. On the other hand, in the case of cotype, we only get l~l~q C 9cl;l~q C ~l~q§ for e > 0. Further variants of type are obtained from any sequence of independent, identically distributed s-stable random variables with 0 < s < 2 [97, Chapter 5]. 9.4. Investigations on type and cotype originally started on the level of spaces. In this setting we have not only GTp -- a T p, but also GCq -- aCq. It is readily seen that
L p e RTp if 1 ~ p ~< 2
and
L q E RT2 if 2 ~< q < oo,
(9.4a)
L p E a c 2 if 1 ~< p ~< 2
and
Lq E RCq if 2 ~< q < oo,
(9.4b)
Operator ideals
473
and these statements are best possible. Moreover, ~,-~dual C ~ p * ; this inclusion is strict, since l l E RG2, but neither co nor l ~ have Rademacher type for any 1 < p ~< 2. From l l 6 RC2 we also conclude that ~ q fails to be surjective. It can be shown [185] that ~ l ~ q C ~ q , l " We even have RCq -- Pq,1 for 2 < q < oo, while RG2 C P2,1. Therefore (9.4b) generalizes a classical result of Orlicz [ 118]" L p E P2,! if 1 ~< p ~< 2
and
L q E P q,1 if 2 ~< q < c~.
9.5. The notions of type and cotype emerged around 1972: 9 Hoffman-JCrgensen [68] studied sums of independent vector-valued random variables. 9 Dubinsky, Petczyfiski and Rosenthal [40] showed that all operators from C(K) into a Banach space X are 2-summing if X* has subquadratic Gaussian averages. 9 Maurey [109] proved that, for 0 < p < 1, all ideals ~13p coincide, and, in particular, 9 Kwapiefi [94] characterized inner product spaces via an abstract two-sided Khintchine inequality. These are the historical roots of the pioneering work of Maurey and Pisier [112] who introduced type and cotype as we use it today. 9.6. In our language, Kwapiefi's amazing Hilbert space characterization takes the form of a formula: RT2 A RC2 = L2. In terms of operator ideals, a more general version is available: ~ I ~ 2 o ~lq~'7~"2 = ~2- In particular, Maurey showed that every operator from a Rademacher type 2 space into a Rademacher cotype 2 space factors through a Hilbert space. This implies that ~ ( L q , Lp) = ,~2(Lq, Lp) whenever 1 ~< p ~< 2 ~< q < oo. Luckily, since every such space L p embeds into some L1, the excluded case q = cx~ is covered by Grothendieck's theorem: ~ ( L ~ , L1) = ~ 2 ( L ~ , L1). In order to get a unified approach, one may ask whether ~ ( X , Y) = ~ 2 ( X , Y) if X* and Y are of Rademacher cotype 2. Under the additional assumption that X or Y has the approximation property, Pisier [ 144], [ 148, p. 41 ] gave indeed an affirmative answer. Moreover, if X is a C*-algebra, then X* 6 RG2, and the required formula even holds for all Y ~ RG2; see [ 186,75]. However, in the general case, Pisier [146], [148, p. 143] exhibited a counterexample. 9.7. If ~ ) ~ 1 ~ "-- ~ 2 o ,~x~1, then DPR is the class of all Banach spaces Y such that , ~ ( L ~ , Y) -- ~ 2 ( L ~ , Y) for every L ~ , and the Dubinsky-Petczyfiski-Rosenthal theorem says that RT~ _ DPR. Maurey [110, p. 116] improved this result by showing that RC2 ___ DPR. On the other hand, DPR _ P2,1. Since RC2 is a proper subclass of P2,1, we only know that DPR belongs somewhere in between. In the case when Y E RCq and 2 < q < oe, a weaker result is true:
~ ( L ~ , Y) -- ~ q , 2 ( L ~ , Y) = ~,~q+e(Lc~, Y) if e > 0, but not necessarily if e - 0.
J. Diestel et al.
474
9.8. The theory of Rademacher cotype and that of p-summing operators are intimately related due mainly to discoveries of Maurey [110]. Here are the basic results underlying these relations [35, pp. 222-224]" ~ 1 (X, Y) ~---~ 2 ( X , Y)
if X 6 FIC2 and Y E L,
~.~I (X, Y) -- ~.~p(X, Y)
if X 6 FiCq and Y E L, where 1 < p < q* < 2,
~ 2 ( X , Y) -- ~ q (X, Y)
if X E I and Y E FIC2, where 2 ~< q < oo.
Consequently,
~'~1 (X, Y) = g3p(X, Y)
if X, Y E FIC2 and 1 < p < oo.
We stress that, by trace duality, the relation ~ I ( X , Y ) - ~ 2 ( X , Y ) t~ec(Y, X) = ~132(Y, X). Hence 13(L~, X) = ~ 2 ( L o o , X) if X 6 FIG2.
translates into
9.9. In 5.13 we have associated with ~r2 the sequence of the ideal norms ~r~n) , which are defined for all operators. This construction will now be repeated in the case of pr2. Let (n) p32 (T) "-- infc, where the infimum is taken over all constants c >~ 0 such that inequality (9.1a) holds for p -- 2 and fixed n. Then
pr~ n) (T) <. v/-ff IITII. The
operators T with
limn~c~ p'r 2(n) (T)/~/-ff = 0 form the injective, surjective, symmetric and closed ideal 9q2E. An operator T E s Y) does not belong to 9~t~E if and only if there exists a constant c ~> 0 with the following property [5]" For n -- 1, 2 . . . . . we can find elements xl n),..., X(nn) ~ X such that
~ k=l
~x~ ~
whenever ~1 . . . . . ~n ~ K. k=l
k=l
This implies that Mn "--span{xl n) . . . . . X(nn) } and span(Txl n) ..... Tx~ n)} are uniformly isomorphic to l~', and cllTxll >>Ilxll for all x ~ M,,. In the setting of spaces, X r FIT means that X contains the l~ 'S uniformly. Spaces in FIT are also referred to as B-convex; see also 11.5. On the other hand, we conclude from 1.18 that an operator T E s Y) does not belong to O lt if and only if there exist a constant c >t 0 and a sequence of elements x t, x2 . . . . ~ X such that oo
k=l
OQ
~< ~-~ I~kl ~
~k Txk
The criteria above and the formula 9~t~E -and global concepts.
whenever (~k) E l i .
~[~uper illustrate
Already Pisier [140] observed that the sequence
P~2(mn) (ST) <<.plr~m) (S)p'c~n) (T)
(pr~n))
the connection between local
is submultiplicative:
for T E ~ ( X , Y) and S 6 ~(Y, Z),
Operator ideals
475
and he employed this to prove U1
dt
~< c sup (xk,x*) 2 IIx*ll<~l k-1
We put 7rp (T) := infc, where the infimum is taken over all constants c ~> 0 with the above property. The class of p-summing operators, denoted by ~3p, is an injective and maximal Banach ideal. We have ~13p ___~]3p if 1 ~< p < cx~, and ~132(X, Y) = ~3p(X, Y) if X 6 L and Y e RC2. Replacing the Rademacher functions by Gaussian random variables yields ~]3• the Banach ideal of y-summing operators. In fact, we obtain nothing new: !:13• = ~13p. The underlying ideal norms ~r• and rrp are different, but equivalent. For every finite dimensional Banach space E and e > 0, the Dvoretzky dimension D(E, e) is defined to be the largest n such that we can find an n-dimensional subspace En of E whose Banach-Mazur distance to l~ is less than 1 + e. The main result, implicitly proved by Pisier [147, Chapter 1] says that
a(e)r~y(E) <~v / D ( E , e) <~b(e)rry(E), where the constants a(e) > 0 and b(e) > 0 do not depend on E; see also [113]. The following asymptotic relations are due to Linde and Pietsch [ 102], while the corresponding results for D(l N, e) were found by Figiel, Lindenstrauss and Milman [49]:
rr• N) ~ N 1/2 if l ~
and
rr•
N)~N
1/q if 2 ~
Moreover, Jr• (l~) N • x/log N. This is the worst case, since rr• (EN) ~ C x/log N for all N-dimensional Banach spaces E N and a universal constant c > 0. As an immediate consequence we obtain the celebrated Dvoretzky theorem [43]: every infinite dimensional Banach space contains almost euclidian subspaces of finite dimension as large as we please. 9.11. The S~minaire Laurent Schwartz 1969/70 was devoted to applications radonifiantes. This subject is based on the fact that cylindrical measures on Banach spaces may behave very badly. The most prominent example is the standard Gauss measure Y on L2 [0, 1], which fails to be ~r-additive. So the question arose whether the quality of a cylindrical measure improves under certain transformations. It can be shown that the integration
operator S" f (t) --+
f0 Sf (t) dt
J. Diestel et al.
476
is y-summing as a map from L2[0, 1] into CZ[0, 1] whenever 0 < ~ < 1/2 and, as a consequence, it turns out that the Wiener measure, which is the image of ~, under S, extends to a Radon measure on the Borel a-algebra of CX[0, 1].
10. U M D operators, Haar type, and uniform convexity Standard references: [33, IV], [139, Chapter 7, Section 8.5]. For k -
10.1. by
X~j) (t)
1, 2 . . . . and j - 1, 2 . . . . . 2 k - l , the Haar functions are defined on [0, 1)
:=
2j-2
2j-1
-+-2(k-l)/2
if
--2 (~-~)/2
if 2j -- 1 < ~ t < - -2 j 2~ 2 k' otherwise.
0
2k
~< t <
2~
In order to get an orthonormal basis of L2[0, 1), the function X0~l) (t) :-- 1 must be added. An X-valued Haar polynomial (of length n), n
2 ~-1
+ZZ
,
k=l j=l is a function f , " [0, 1) --+ X that takes constant values on the dyadic intervals A~j) := [
,
) with j = 1 . . . . . 2n. In other terms, f , must be Dn-measurable, where Dn denotes
the algebra generated by
A (1)
, ..
A(2") . . , ,._.~n
Writing
m
fm " = ~ d k
with do "-- x 0( 1 ) and dk
"--
--2k-I Z j = I x~J)x~J)
k=0
yields a Walsh-Paley (dyadic) martingale (fo .... , fn). Indeed, if m < n, then fm is the conditional expectation of fn with respect to 79m. The functions do . . . . . d, so obtained are said to be martingale differences. 10.2. Let 1 < p < oo. We refer to T E ~ ( X , Y) as a UMD operator if there exists a constant c ~> 0 such that
LTd,(t) k=0
I1I (SoL dt
e~d~(t)
<<.c
1/p dt
(10.2a)
k=0
for n E N, any choice of X-valued martingale differences do . . . . . dn and arbitrary signs eo,..., en = -+-1. The class of these operators, denoted by )AgJIX), does not depend
Operator ideals
477
on p and is an injective, surjective, symmetric and maximal Banach ideal; the ideal norms II.p(T) "-- infc being equivalent, but non-normalized for p 7~ 2. Burkholder [21] showed that I X p ( I K ) - m a x { p - 1, p * - 1}, which is the unconditional constant of the Haar basis in Lp[O, 1); see also [192, p. 63]. The notation UMD stems from the fact that inequality (10.2a) expresses an unconditionality property f o r martingale differences. UMD spaces were considered for the first time by Maurey [ 111, Expos6 2]. He already observed that L p E UMD if 1 < p < cx~. However, the real break-through only came a few years later. In a series of papers Burkholder [18,19,22] found a characterization by a geometric property, called ~"-convexity, and he discovered the connection with the vectorvalued Hilbert transform to be discussed in 11.4. 10.3. Let 1 ~< p ~< 2. An operator T 6 t3(X, Y) is said to have Haar type p if there exists a constant c ~> 0 such that /7
y ~ Tdk (t)
<. c
(s01
Ildk(t)ll p dt
(10.3a)
k=0
k=O
for n E N and any choice of X-valued martingale differences do . . . . . dn. The class of these operators, denoted by 9A'Ep, is an injective, surjective and maximal Banach ideal. Its ideal norm OtTp(T) "-- infc is non-normalized for p 7~ 2. Clearly, ~1'~1 - - ,t~. Setting 2k-1
do
.
0.
and dk . .
2-(k-1)/2 Z
-
XkXk
"(j)
--
Xkrk
with Xl . . . . . x/7 E X,
j=l
inequality (10.3a) is equivalent to (9.1a). Hence 9.1~gp __ 9Vgp. The summation operator I7 :ll --+ l ~ shows that this inclusion is proper. Indeed the difference is quite large, since there exist spaces in FIT2 that have no Haar type p > 1; see [72] and [ 150]. Put a~/7) (T) "-- infc, where the infimum is taken over all constants such that inequality (10.3a) holds for p -- 2 and fixed n. Then a~/7) (T) ~< ~/n + 1 lITII. The operators T with limn--,~ ot~/7) ( T ) / ~ / B -- 0 form the injective, surjective, symmetric and closed ideal 9.1ff7. If 1 < P2 < Pl ~< 2, then 9A'Epl C ~'['~p2 C ~.1'~. As in the context of Rademacher functions, a Banach ideal ~9.11~qformed by operators of Haar cotype q can be introduced for 2 ~< q < ec, and a closed ideal 9.1ff is obtained in analogy with 9~ff. But we now encounter a situation of immaculate duality: p~l . . .r'Cf . dual p - 9 A f f p . and p_l . . .ord . qual --9.1~gq,. In the limiting case, we even have 9~l"~dual _ f21 ~ - - ~ 1 ~ . - - 9.1 l~. dual .
J. Diestel et al.
478
10.4. The order relations between the ideals treated in Sections 9 and 10 are illustrated in the following diagram, where the arrows indicate strict inclusions:
~'~2
~'~P
)-~J~
Pdl~q
~'7~"2
~lq~ 2
~,[l~2
10.5. Let 1 ~< p <~ 2. An operator T 6 ~ ( X , Y) is called uniformly p-smooth if there exists a constant c >~ 0 such that l l y + Txll p + IlY- Txll p 2 Let 2 ~< q < cx~. An operator T 6 s constant c ~> 0 such that
_ ilyllP)1/p
cllx II for all x 6 X and y 6 Y.
Y) is called uniformly q-convex if there exists a
x++x_
T x + - T x - II <~c ( llx+ llq -+2
q)l/q
for all x+ 6 X.
10.6. We are also interested in the limiting cases of p-smoothness as p--+ 1 and of q-convexity as q --+ ~ . An operator T e s Y) is called uniformly smooth if for every e > 0 there exists 6 > 0 such that IlY+ Txll + I l Y - Txll
- 1 ~ EIIxll
whenever IlYll- 1 and Ilxll ~ ~.
An operator T 6 ~ ( X , Y) is called uniformly convex if for every e > 0 there exists 6 > 0 such that
Tx+2
ii~<e
whenever
Ilxi II -
1 and
2 x++x ii> ~ 1 - 3 .
10.7. The operators just defined form one-sided ideals that are related to the Haar type and cotype ideals in the following way: 9 T ~ 9A~Ep(X, Y) if and only if Y admits a renorming such that T becomes uniformly p-smooth.
Operator ideals
479
9 T E PA~q(X, Y) if and only if X admits a renorming such that T becomes uniformly q-convex. These criteria remain true in the limiting cases as well. 10.8.
The following examples should be compared with (9.4a) and (9.4b).
Lp E ATp if 1 ~< p ~< 2
and
Lq EAT2 if2~
(10.8a)
L p E A02 if 1 < p ~< 2
and
Lq E AC,q if 2 <~ q < cx~.
(lO.8b)
The easy parts follow from Clarkson's inequalities I l f + g l L p l l p-+- IIf - gltpll p ~ 2(llflLpll p + llglLpll p)
ifl~p~2
and IIf + gltq IIq -+- IIf - gltq IIq ~ 2(llfltq IIq + Ilgltq IIq)
if 2 ~< q < cx~,
whereas the rest is more involved. 10.9. It turns out that PA~ = PAX coincides with ~ s u p e r , the ideal of super weakly compact operators. The local nature of this concept is stressed by the following criterion: An operator T E ~ ( X , Y) fails to be super weakly compact if and only if it factors n An> T Bn> the finite summation operators uniformly. That is, Z'n "l 1 X ~Y In such that SUPn IIB~ II IIAn II < ~ , where
~:. : ( ~ ) k=l
l <~m<~n
The global counterpart of this property was already discussed in 1.12 when we observed that the infinite summation operator Z :ll --+ l ~ is C~-universal. 10.10. The historical starting point of all renorming results stated above is James's theory of super reflexivity [71]. Enflo [45] showed that every super reflexive space can be given an equivalent uniformly convex norm, and Pisier [142] developed a quantitative approach in terms of Walsh-Paley martingales. Moreover, Beauzamy [6] introduced the ideal of uniformly convexifiable operators. The inclusion 3.193ED C 9A~Eextends a result of Pisier [ 141, Annexe 2] which says that every UMD space is super reflexive. 10.11. As already mentioned in 1.15, the (C(K), Y)-components of ~3 and ~/3' coincide. Diestel and Seifert [36] and Pe~czyfiski [120] observed that the same is true for the ideals ~3~5 and ~5c0 defined in 1.16 and 1.18; see also [37, p. 160]. This list can be enriched by adding the corresponding super ideals.
J. Diestel et al.
480
Let 9,1 be a quasi-Banach ideal, then an operator T 9 s Y) belongs t o ~inj, the injective hull of 9,1, if and only if there exist an operator To 9 s Y0) and a function N:IR+ --+ IR+ such that IITxII ~ N(e)llToxll-4- ellxll --inj
see [74, pp. 473-476]. Obviously, ~1
for x 9 X and e > O; ---~inj
__C"'2 ~ ~
-----~]'' However, using Niculescu's
characterization [117] of weakly compact operators on C(K), we get -'~[nJ(c(K), Y) -~ 3 ( C ( K ) , Y) for all Banach spaces Y. Hence the (C(K), Y)-components of the ideals ~. inj -~ inj ~J1 , ""2 , ~
and ~ 3 coincide. This observation connects, among others, the non-
ultrapower-stable ideal ~
with the ultrapower-stable ideal
~. inj D 1
.
So
"" T 9 ~Z~3(C(K), Y) ==~ T 9 ~;nJ(C(K), Y) ==~T u 9 ~J1. i n J ( c ( K ) t ~ y U ) :=~ TU 9 ~ ( C ( K ) U , yU) :=~ T 9 ~:Ijsuper(c(K), Y),
and all weakly compact operators on spaces C(K) are even super weakly compact. Moreover, using the ideal 9 l ~ defined in 9.9, Beauzamy [7] showed that ~]1~super C ~]~ 0 ~J~"~ C f13~ C ~ . Hence ~ j ~ s u p e r _ _ ~ s u p e r . Let X be a reflexive quotient of C(K), and fix a surjection Q : C ( K ) --+ X. Then Q e ~ 3 S"per, which in turn implies that X is super reflexive. However, super reflexive spaces have finite Rademacher cotype. Therefore Q is q-summing and hence strictly q-integral for some 2 ~< q < oo. In conclusion: we find a probability measure # such that X becomes a quotient of Lq (K, #). This is an ideal theoretic proof of the dual version of a theorem due to Rosenthal [158].
11. Extension of classical theorems to vector-valued functions
Standard references: [37,135]. 11.1. A major goal of functional analysis consists in investigating how much of scalar variable theory might be extended to functions taking their values in Banach spaces. Some examples have been discussed earlier: the relation between absolute and unconditional convergence of series, the theory of Rademacher type and cotype as well as its counterpart in the context of Haar functions. Ensuring the validity of classical results in the vectorvalued setting often requires to impose quite strong restrictions on the underlying Banach spaces. In this section, we will address such problems with respect to the Fourier transform, the Hilbert transform, the Riesz projection, the Rademacher projection and the RadonNikod3~m theorem. 11.2. As one of the first steps Bochner [12] developed his Integration von Funktionen, deren Werte die Elemente eines Vektorraumes sind. For 1 ~< p < oe, we denote by [Lp(M, #), X] the Banach space of all Bochner p-integrable X-valued functions defined on a a-finite measure space (M, #); see also [37, Chapter II]. As usual, functions are
Operator ideals
481
identified when they coincide almost everywhere. In a natural fashion, the algebraic tensor product L p ( M , #) @ X can be viewed as a (dense!) subset of [Lp(M, #), X]. Some of the deepest questions in Banach space theory have their roots in the following simple situation: if S : L p ( M , #) --+ Lq(N, v) and T :X --+ Y, then
Z
fk | x~ w-~
k--1
Sfk @ Txl( k=l
defines a map
S | T : L p ( M , #) @ X --+ Lq(N, v) @ Y; when does there exist a continuous extension
[S, T] : [Lp(M, #), X] --+ [Lq(N, v), Y]? 11.3.
The Hausdorff-Young theorem asserts that the Fourier transform
f'" f (t) ~ fA(s) "-- F
~ f (t) e ist dt OG
induces an operator from L p (R) into L p. (]R) if 1 ~ p ~< 2. Accordingly, T 6 E(X, Y) is said to have Fourier type p if the extension [9r , T] exists. These operators form the injective, surjective, symmetric and maximal Banach ideal ~57p. Its ideal norm t#'rp(T) := 11[f', T]:[Lp(IR), X] ~ [Lp, (IR), Y]]], is non-normalized if 1 < p ~< 2. Babenko and Beckner [8] have computed the exact value of ~'rp(/K) = I1~-: Lp(]R) ~ Lp*(]R)II. The following characterization is more flexible and displays, in particular, the local nature of the concept under consideration. An operator T 6 ~ ( X , Y) has Fourier type p if and only if there exists a constant c ~> 0 such that, for any choice of n 6 1~ and x l . . . . , Xn ~ X,
(1
fi
Txk e ikt
dt
~< c
k-1
IIx~ IIp k--1
Putting 4}1:p(T) "-- infc yields an equivalent ideal norm. Banach spaces of Fourier type were first considered by Peetre [ 119] who used tools from interpolation techniques to show that L p and Lp. are contained in FTp. The fundamental Kwapiefi theorem [94] asserts that / 2 -- FT2. However, it seems likely that ~2 is a proper subclass of ~ff72. 11.4.
The periodic Hilbert transform is defined by
_
7-/" f (t) ~-+ f (s) "-- ~
f (t) cot 7/"
(s_,) 2
dt.
J. Diestel et al.
482
If f e L p (~) and 1 < p < c~, then the right-hand integral exists almost everywhere in the sense of Cauchy's principal value, and the Riesz theorem tells us that 7-/yields an operator from Lp (qF) into itself. An operator T 6 ~ ( X , Y) is said to be compatible with the Hilbert transform if the extension [7-/, T] exists. In this way we get another injective, surjective, symmetric and maximal Banach ideal ~ E , which does not depend on the choice of p. All ideal norms Xlrp(T) : = 1117-/,T]: [Lp(7~),X] ~ [ L p ( ~ ) , Y]ll are equivalent, but non-normalized for p -~ 2. Pichorides [127] has computed the exact value of X rp(IK) = 117-t: Lp(T) Lp(qF)ll. Local characterizations are available as well. For example, T e ~ ( X , Y) is in ~ E if and only if there exists a constant c ~> 0 such that
( io"
11 Z Txk c o s kt k=l
' )"' ( i0 • dt
~< c
rr
1 --
2
1/2
xk sin kt
k=l
for n e N and Xl . . . . . Xn e X. Again the smallest constant provides an equivalent ideal norm. The same ideal can be obtained by working, instead of the periodic Hilbert transform, with the Hilbert transform on the real line, some discrete versions thereof, or the related Riesz projections. If 1 < p < ex~, then L p 6 HT, but l l, l ~ ~ H'I'. The remarkable Burkholder-Bourgain theorem [20,14] asserts that UMD = HT. We do 9
not know whether this result extends to operators: JI93tX) -- ,9'E. 11.5.
Khintchine's inequality implies that the Rademacherprojection
~" f (t) w-> Z
(fo'
f (t)rk(t) d t ) rk(s)
k=l
maps L p[O, 1) into itself whenever 1 < p < ex~. An operator T e ~ ( X , Y) is compatible with the Rademacherprojection if there exists a continuous extension [7~, T]. The injective, surjective, symmetric and maximal Banach ideal 9~t~13 so obtained does not depend on the choice of p. All ideal norms prtp(T) := [117~, T]:[Lp[O, 1), X] --+ [Lp[0, 1), Y][[ are equivalent, but non-normalized for p # 2. A local characterization reads as follows: an operator T e s Y) belongs to 9~r if and only if there exists a constant c ~> 0 such that
(s0 s
k=l
)
T fn(t)rk(t) dt rk(s)
as
~< c
IIf~ (t) llp dt
for n 6 N and all X-valued Haar polynomials fn; see 10.1. Then the smallest admissible constant coincides with the ideal norm prrp defined above.
Operator ideals
483
Let us return to the relation r_ __p ual C 91ffp, stated in 9.4. Though we do not have equality, there is something like a reverse: 91~dual o 9 1 ~ C 91(~'.p. Moreover, 91gt C 91~. In the setting of spaces, we even have FIP = FIT; a formula which is usually phrased by saying that K-convexity and B-convexity are the same; see [ 145] and 9.9. 11.6. Next, we look for a vector-valued version of the Riesz representation theorem. Given a compact Hausdorff space K and a Banach space X, we want to represent operators A : C ( K ) --+ X by integration against X-valued measures. There are a few problems in the general case, but weak compactness of A is characterized by the existence of a regular X-valued Borel measure m on K such that
A f -- s
f (~)dm(~)
for all f 9 C(K).
Moreover,
A 9 gt 1(C(K), X) ~ m has a finite variation Iml, A 9 011 (C(K), X) ~ m has a derivative with respect to [ml. The latter condition means that there is a function a 9 [L 1(K, #), X] such that
m(B) = fB a(~) dlml(~)
for all Borel sets B in K.
An operator T 9 t2(X, Y) is said to have the Radon-Nikodym property if A 9 ~ 1 (C[0, 1], X) implies TA 9 011 (C[0, 1], Y). The same holds then with any C(K) instead of C[0, 1]. A large variety of equivalent properties is known; see [37, pp. 217-219]. For example, T f T carries absolutely continuous functions f :[0, 1] --+ X to functions g : [ 0 , 1] >X >Y that are differentiable almost everywhere. In this case, we have b
g(b) - g(a) --
fa
g' (t) dt
for0 ~< a
The operators with the Radon-Nikod))m property form an injective and closed ideal, 910t. From l l E FIN and co, loc ~ FIN it follows that 9101 and 9101 dual are incomparable. We have ~ C 9101 C ~ r and ~ C ~r C l~[1. Since all reflexive spaces are in FIN, a deep theorem of Grothendieck [59, p. 132] follows: ~ o ~1 -- 9 t l , isometrically. If ~ e p denotes the closed ideal of all operators with separable range, then ~5r dual C ~r and Stegall [182] proved that 919l duat- ~ r o l~r -1 . In particular, FIN dual coincides with the class of Asplund spaces, which are defined by the property that their separable subspaces have separable duals. The ideal 919t dual behaves better than 919t. It is not only injective, but also surjective; fJlgt dual has the factorization property, whereas 919t even fails to be idempotent; see [ 182, 53]. Moreover, there are C9101 dual -universal operators. For example, we may take the map
484
J.
Diestel et al.
from ll into L ~ [ 0 , 1) given by (~k) ~ Y~'~--1~kXk. Here (Xk) is any enumeration of the Haar functions, normalized with respect to the L~-norm. The difference between 9~9t and ~r disappears when passing to the associated super ideals; in both cases we get ~super. Beginning in 1936 with a pioneering paper of Clarkson [27], the study of the RadonNikod3~m property was mainly carried through on the level of spaces. There is a geometric part based on the concept of dentability [154]; see [37, Chapter VII, pp. 208-219] for a historical survey. Our presentation has emphasized the analytic point of view, which is closely related to vector measures. Operators with the Radon-Nikod3~m property were first considered by Reinov [152] and Linde [101].
12. Operator ideals and tensor products, the approximation property Standard references: [32,34], [37, Chapter VIII], [ 135, Chapter 10], [ 106,162,167]. 12.1. In this final section, we describe how ideal theory grew out of the metric theory of tensor products. Nowadays it seems that the novice outstripped the master. Let us compare some pros and cons: 9 The language of operator ideals is more appropriate for applications. 9 Operators can be composed, which is not so easy for tensors and bilinear forms. 9 Tensor product techniques yield minimal and maximal ideals, but there are plenty in between. This is in particular true when we consider closed ideals. 9 9tl is the least ideal that can be obtained from a tensor norm. However, smaller ones are needed, for example, in trace theory. 9 Tensor products provide a deeper insight into trace duality. 9 Tensor products proved to be an indispensable tool for the understanding of all phenomena related to the various approximation properties. Since mathematics is not a one-way street, it may well happen that somewhere in the future this evaluation will be refuted. 12.2. In 1936, Murray and von Neumann [115, Chapter II] invented direct products of Hilbert spaces to be used as a tool in their fundamental studies of rings of operators; their notation: H1 | | Hn. Two years later, the name tensor product was introduced by Whitney [ 191 ] when he investigated abelian groups: If the linear spaces are topological, a topology may be introduced into the tensor product . . . . In the case of Hilbert spaces, there is a natural definition of the topology in the product. In the intermediate case of Banach spaces, probably the norm I~1 may be defined as the lower bound of numbers ~ Igillhil for expressions y~ gi 9hi of Or. 1 1 This definition was suggested to me by H.E. Robbins.
Around 1950, Schatten [165,167] and Ruston [160,161] developed the basic theory. The most significant contributions, however, are due to Grothendieck. His main achievements in this field include
Operator ideals
485
9 introducing localization techniques in the theory of tensor products and in the theory of Banach spaces as a whole; 12.6, 9 enriching the theory through his 14 natural tensor norms; 12.10, 9 stressing the significance of approximation properties and giving various equivalent reformulations; 12.8, 9 pointing out the use of Banach spaces in the theory of (nuclear) locally convex spaces. 12.3. The theory of tensor products was developed on three levels. The simplest way is to work on a fixed algebraic tensor product. On the other hand, we can consider all X | Y's with X 6 t_ and Y 6 !_ simultaneously. More elementary, it is possible to start with tensor norms that are only given on all finite dimensional domains: E | F. We begin by describing the most general approach. By a tensor norm (| we mean a rule r that assigns to every element u = ~ = l xk | Yk in the algebraic tensor product X | Y of any pair of Banach spaces a nonnegative number o~(u) such that the following conditions are satisfied: (TN1)
a(x @ y) --Ilxll IlYll for x 6 X and y E Y.
(TN2)
ot(u + v) ~< or(u) + or(v) for u, v ~ X | Y.
(n
ot ~ A x k (TN3)
@ Byk
)
~ IIAII Ilnll~
(n)
k--1
~xk
Q Yk
k=l
17
for ~ x k
| y~ ~ X N Y, A ~ t2(X, Xo) and B E t3(Y, Yo).
k=l
Since (TN3) implies that a()~u) = I~l~(u) for u 6 X | Y and )~ 6 K , we have indeed a norm. For greater clarity, we shall sometimes write ot(ulX | Y) instead of ~(u). The objective of the theory is to investigate the Banach spaces X | Y that are obtained by completing X @ Y with respect to any tensor norm a. The duals (X | Y)*, formed by bilinear functionals, are considered as well. 12.4. If (TN3) is supposed to hold only for functionals instead of arbitrary operators A and B, we get
Z ( x * , x k ) ( Y k , Y*) <~ IIx*ll IlY*ll~
xk | Yk
/7
for Z x k | Yk 6 X @ Y, x* 6 X* and y*
y*.
k=l
In the terminology of Schatten, a norm on a single algebraic tensor product is called a crossnorm if it satisfies (TN1). He also used condition (TN~). Following Grothendieck, we refer to a norm with both properties as reasonable. Schatten defined the greatest crossnorm, now the projective tensor norm, /7
rr(u) "-- i n f ~ k=l
Ilxkll IlYkll,
J. Diestel et al.
486
where the infimum ranges over all representations of u, and the least reasonable crossnorm, now the injective tensor norm,
e(u) " - s u p {
~-~(x*, xk) (Yk,Y*) 9 IIx*ll ~ 1, Ily*ll ~ 1 }. k=l
That is, for any reasonable norm t~, we have e(u) <<.or(u) <<.rt(u). We know that L 1(M, #) | X = [L 1(M,/z), X] for every measure space (M,/z). On the other hand, C ( K ) | X coincides with [C(K), X], the Banach space of all X-valued continuous functions on a compact Hausdorff space K. More generally, if 1 ~< p < c~, then Bochner's norm
IlflLpll " -
IIf(~)ll p d/z(~)
restricted to L p ( M , lZ) | X is reasonable. We even have
~-~A | Tx~ILp
~< IITII
k--1
•
fk | Xk [ Lp
k=l
for T ~ 13(X, Y), f l . . . . , fn E L p ( M , lZ) and xl . . . . . Xn ~ X. On the other hand, if 1 < p, q < c~, then there exist operators S" L p ( M , lZ) --+ L q ( N , v) and Banach spaces X such that F/
H
~
(12.4a) k=l
does not hold for any constant c > 0, independent of fl . . . . . fn ~ L p ( M , Iz) and x 1. . . . . Xn ~ X. Hence Bochner's norm can be derived from a tensor norm only for p -- 1. It was just this deficiency that called forth the introduction of the ideals 9l~gp, 9,1~gp, ~ g p , tlgJED, etc. For example, if S" lp --+ L2[0, 1) is given by (~k) w-~ Y~'~--1 ~krk, then (12.4a) defines spaces of Rademacher type p; see 9.1 and 11.2. 12.5. Already Schattten observed that it is advantageous to define crossnorms not only on a single X | Y, but on all algebraic tensor products. In this case, he spoke of a general crossnorm. However, without any additional property, crossnorms defined on different domains X N Y were unrelated and could be chosen at random. So we need some connecting link, and this is just condition (TN3) in 12.3. Schatten referred to crossnorms satisfying this inequality as uniform [167, p. 29], [60, p. 9]. Gradually, it turned out that (TN3) and (IN3) are the decisive keys of both, the theory of tensor products and the theory of Banach operator ideals.
Operator ideals
487
12.6. We stress that Grothendieck used the term tensor norm (| only in the finite dimensional setting. However, in a subsequent step he constructed extensions of such ot's. This process will be described next. L e t u -- ~ = 1 x~:Nyk E X | Y. If M 6 DIM(X) and N 6 DIM(Y) such that Xl . . . . . Xn U_ M and Yl . . . . . Yn E N , then u can also be considered as an element of M | N. We put A
ot ( u l X @ Y) "-- i n f a ( u l M @ N), where the infimum ranges over all M and N described above. In this way we get the greatest extension of or, which is defined on all tensor products X @ Y. The least extension is given by
v
/( n
ot ( u l X N Y) "-- sup t~
i
~Q~xk|
X/M N Y/N
)
" N 6 COD(Y)
"
k=l
12.7. Let F "-- {E, F . . . . } denote the set of finite dimensional Banach spaces. Identifying E* | F and 13(E, F) shows that, in this case, the theories of tensor norms and ideal norms can be translated into each other. In particular, the tensor norms ~r and e correspond to the ideal norms t l - P l - - P7 and II" INrespectively. In the infinite dimensional setting, this relationship becomes much more complicated. For example, there exist operators T such that tl (T) < P l ( T ) and v~ (T) < v ~ ( T ) , respectively. Hence ideal norms over F admit different extensions over L; and the same phenomena occur for tensor norms. This situation will now be discussed in detail. Our starting point is some bisexual 0t as indicated in the middle of the following picture: V
I
Jot
~.[max
ol
m,
least extension
I
maximal ideal
tensor
I ideal norm
finite dimensional setting
A
Qot
~min
greatest extension
minimal ideal V
A
Then, by 12.6, 4.5 and 4.8, the extensions a and a as well as the Banach ideals ~[max and f21min are uniquely determined. Every tensor u -- Z ~ = I Xk* | Yk ~ X* @ Y defines an operator S" X -+ Y by 17
S ' x ~--~ X,Xk)Yk. k--1
J. Diestel et al.
488
It turns out that the map u w-~ S extends continuously to a metric injection
Jot" X* ~ ~ Y ---+ f2[max ( x , Y);
(12.7a)
see [32, p. 205]. Since ~max (X, Y) may contain non-approximable operators, the range of Jot is in general a proper subset. In this way, we can identify X* ~ e Y and ~ ( X , Y). The situation changes significantly when t~ is replaced by &. Then the continuous extension of u w-~ S becomes a metric surjection
Qot" x* ~
Y --+ }2[min(x, Y);
(12.7b)
see [32, p. 290]. This explains why Grothendieck refers to operators in ~2[min as a-nuclear. In view of the relationship between OIl and 31, it is justified to say that operators in ~max are a-integral. In the R~sum~, however, we find the term application de type or. Furthermore, with every functional u* 6 (X* | Y)* we associate an operator T ' Y --+ X** via (x*, Ty) -- (x* Q y, u*)
forx*6X*andyEY.
The correspondence u* --+ T is an isometry between (X* ,~A y)* and 9A*(Y, X**). ~ot If Qot in (12.7b) is one-to-one, then the predual X* | Y can be identified with
~,[min(x, Y). Unfortunately, this does not always happen. However, if Y has the approximation property, then everything becomes fine. In this case, the equation (2[min(x, Y)* "-9,1" (Y, X**) means the following. Since S ~ (2[min(x, Y) is approximable, an operator S ~176 can be defined by viewing S** as a map from X** into Y. Then S~176~ OIl(Y) for T E 9,1"(Y, X**), and we may put (S, T) "-- trace(S~176 This formula is the historical source of trace duality; see [176] and [32, p. 293]. 12.8. A Banach space X is said to have the approximation property if the following (equivalent) conditions are satisfied: 9 For every compact subset K and every e > 0 there exists T 6 ~ ( X ) such that mix T x II ~< e for x E K. 9 The natural map from X* ~ X into X* ~ e X or s is one-to-one. 9 The natural map from Y* ~ X into Y* ~ e X or s X) is one-to-one for all Y E k. 9 vl (T) -- v~ (T) for all T ~ ~ ( X ) . 9 Vl (T) -- v~ (T) for all T E ~(Y, X) and all Y 6 L. 9 The functional T ~-~ trace(T), defined for finite rank operators, admits a continuous extension to 9 t 1(X). 9 ~(Y, X) - .~(Y, X) for all Y ~ L. It seems to be open whether we may add: ~ ( X ) - ~ ( X ) . A Banach space X has the metric approximation property if, in the first condition of the above list, T 6 ~ ( X ) can be chosen such that IJTJl ~< 1. Now we are able to describe precisely the relationship between t l, vl and v 1o .
489
Operator ideals
t l (T) = P l (T) for all T 9 ~(Y, X) and all Y 9 k if X is reflexive. 9 Vl (T) - v~ (T) for all T 9 ~(Y, X) and all Y 9 k if and only if X has the approximation property. 9 t l (T) -- v~ (T) for all T 9 ~(Y, X) and all Y 9 k if and only if X has the metric approximation property. Approximation properties were introduced by Grothendieck in 1955. He also discovered various reformulations. For a long time many experts believed that all Banach spaces would have this property. So it occurred as a big surprise when Enflo [46] constructed a counterexample. Later, Szankowski [183] showed that even ~(12) fails this property. 9
12.9. In general, the natural map from X | Y to X | Y is neither one-to-one nor onto. Grothendieck conjectured that isomorphy only occurs when at least one of the spaces is finite dimensional [59, Chapter II, p. 136], [60, p. 74]. But this is not the case! Pisier [146] constructed an infinite dimensional space X such that X | X - X | X, algebraically and topologically. In addition, he arranged X and X* to have Rademacher cotype 2. Translating this result in the language of ideals yields ~ ( X * , X) -- 9tl (X*, X), and John [78] observed that .~(X, X * ) - 9tl (X, X*). Moreover, ~ ( X ) - 9tl (X). Loosely speaking, X admits only very few operators. It is still open if .~(X) -- 9tl (X) can be true for some infinite dimensional Banach space X. 12.10. Taking Grothendieck's point of view, the following considerations are carried out for operators acting between finite dimensional spaces. Given ideal norms ot and /3, the symbol ot -< 13 means that r ~< c f l ( T ) for all T 9 t2(E, F), where the constant c >~ 1 does not depend on E 9 F and F 9 F. If ot -3 and 13 -< a, then a and 13 are said to be u n i f o r m l y equivalent: ot • ft. For example, from Grothendieck's t h d o r k m e f o n d a m e n t a l we get the formulas 1r2 ~ ~.2 o ~.oc,
Ndual X ~1 o ~2
and
gdual o 1r2 x ~1 o ~oc.
In 2.2 and 4.6 we have defined the rules a w-> Otdual Ol ~ Olinj, Ol ~ Olsur and c~ ~ or* which assign to every given ideal norm ot a new ideal norm Olnew. The following list contains two more examples, namely ot ~ ot ext and a ~ ot lift, where the superscripts refer to ' e x t e n s i o n ' and 'lifting' otdual(T) "-- or(T*), otadJ(T) "-- or* ( T ) -- sup{ Itrace(ST)l" a ( S " F --+ E) ~< 1 }, otinJ(T) "-- sup{ot(BT)"
liBf--> l~[ I <<.1,
a'"r(T) - - sup{u(TA)" [IA " l'~ --+ oteXt(T)
El
1, n -- 1, 2 . . . . },
9-- inf{ot(To)llAIl" T -- ToA, E
wlift(T) " - i n f { l l B l l c ~ ( T o ) "
T--BTo,
n - 1,2 . . . . },
E
A >1~
To> F, n~>rank(T)},
To> 11n
B
> F, n ~> rank(T) }.
Taking the operator norm as a starting point, Grothendieck observed that there exists a minimal set of 14 n a t u r a l ideal n o r m s , which is stable with respect to ot ~ ot dual . . . . .
490 ot ~
J. Diestel et al. ot lift provided that we identify uniformly equivalent members. Thanks to the relations
(oldual)SUr__ (olinj)dual,
oldual) dual -- Ol,
(oladj)ext-- (olinj) adj,
(oladj) adj = ~,
(olSUr)sUr--otSUr '
(olinj) inj = otinj,
(~dual)lift
(Olext)dual,
(oladj)lift--- (olsur) adj, (Olext)ext-.-Olext '
(~lift)lift--ollift '
it suffices to check the invariance under a ~-+ Otdual, Oll----> Otadj and a ~ ot inj. In the following diagram the arrows point from the smaller ideal norms to the larger ones, in the sense of the preodering -<. There are no further relations as indicated; for example, we know from 6.6 that ~.l s ~1.
lift1 t dual
7[ 1
7[
dual
o 7[2
t 7~dual
7[2
lgdual
192
~adj
t axis of symmetry
~.1
i2
O~
Otdual
II II axis of symmetry
~2
t ~ S(X)U F
II. II We stress that the above considerations were originally formulated in the language of tensor norms; [60, p. 37], [32, p. 362]. This shows why the R d s u m d was the germ-cell of ideal theory.
References [ 1] D.E. Allekhverdiev, On the rate ofapproximation of completely continuous operators byfinite dimensional operators, Azerbajzhan. Goz. Univ. Uchen. Zap. (Baku) 2 (1957), 27-37 (Russian). [2] I. Amemiya and K. Shiga, On tensor products of Banach spaces, Kodai Math. Sem. Report 9 (1957), 161-178. [3] A. Baernstein, On reflexivity and summability, Studia Math. 42 (1972), 91-94. [4] S. Banach and S. Saks, Sur la convergence forte dans le champ LP, Studia Math. 2 (1930), 51-57. [5] B. Beauzamy, Opdrateurs de type Rademacher entre espaces de Banach, S6m. Maurey-Schwartz, Expos6s 6 et 7, l~cole Polyt6chnique, Paris (1975/76). [6] B. Beauzamy, Opdrateurs uniform~ment convexifiant, Studia Math. 57 (1976), 103-139.
Operator ideals [7] [81 [91 [101 [11] [121 [131 [141 [15] [161 [171 [18] [191 [20]
[21] [22] [23] [241 [25] [26] [271 [28] [29] [301 [311 [321 [33] [341 [351 [36] [371 [381
491
B. Beauzamy, Propridtd de Banach-Saks, Studia Math. 66 (1980), 227-235. W. Beckner, Inequalities in Fourier analysis, Ann. of Math. 102 (1975), 159-182. G. Bennett, Some ideals of operators on Hilbert space, Studia Math. 55 (1976), 27-40. G. Bennett, V. Goodman and C.M. Newman, Norms of random matrices, Pacific J. Math. 59 (1975), 359365. S. Bernstein, Sur la convergence absolue des s~ries trigonomdtriques, C. R. Acad. Sci. Paris 158 (1914), 1661-1664. S. Bochner, Integration von Funktionen, deren Werte die Elemente eines Vektorraumes sind, Fund. Math. 20 (1933), 252-276. E Bohnenblust, An axiomatic characterization of L p spaces, Duke Math. J. 6 (1940), 627-640. J. Bourgain, Some remarks on Banach spaces in which martingale difference sequences are unconditional, Ark. Mat. 22 (1983), 163-168. J. Bourgain, New Banach space properties of the disc algebra and H ~ Acta Math. 152 (1984), 1--48. J. Bourgain and E Delbaen, A class of special s spaces, Acta Math. 145 (1980), 155-176. J. Bourgain and J. Diestel, Limited operators and strict cosingularity, Math. Nachr. 119 (1984), 55-58. D.L. Burkholder, A geometric characterization of Banach spaces in which martingale difference sequences are unconditional, Ann. Probab. 9 (1981), 997-1011. D.L. Burkholder, Martingale transforms and the geometry of Banach spaces, Probability in Banach Spaces (Medford, 1980), Lecture Notes Math. 860 (1981), 35-50. D.L. Burkholder, A geometric condition that implies the existence of certain singular integrals of Banachspace-valued functions, Conf. on Harmonic Analysis in Honor of Antoni Zygmund (Chicago, 1981), Vol. I, Wadsworth, Belmont (1983), 270-286. D.L. Burkholder, An elementary proof of an inequality of R.E.A.C. Paley, Bull. London Math. Soc. 17 (1985), 474-478. D.L. Burkholder, Martingales and Fourier analysis in Banach spaces, Probability and Analysis (Varenna, Italy, 1985), Lecture Notes Math. 1206 (1986), 61-108. J.W. Calkin, Two-sided ideals and congruences in the ring of bounded operators in Hilbert space, Ann. of Math. (2) 42 (1941), 839-873. B. Carl, Inequalities between absolutely (p, q)-summing norms, Studia Math. 69 (1980), 143-148. B. Carl and I. Stephani, Entropy, Compactness and the Approximation of Operators, Cambridge Univ. Press (1990). S. Chevet, Sur certains produits tensoriels topologiques d'espaces de Banach, C. R. Acad. Sci. Paris, S6rie A 266 (1968), 413-415. J.A. Clarkson, Uniformly convex spaces, Trans. Amer. Math. Soc. 40 (1936), 396-414. C.C. Cowen, Composition operators on Hilbert spaces of analytic functions: A status report, Proc. Symp. Pure Math., Vol. 51, Part I, Amer. Math. Soc. (1990), 131-145. C.C. Cowen and B.D. MacCluer, Composition Operators on Spaces of Analytic Functions, CRC Press (1995). D. Dacunha-Castelle and J.L. Krivine, Applications des ultraproduits it l'dtude des espaces et des algkbres de Banach, Studia Math. 41 (1972), 315-334. W.J. Davis, T. Figiel, W.B. Johnson and A. Petczyfiski, Factoring weakly compact operators, J. Funct. Anal. 17 (1974), 311-327. A. Defant and K. Floret, Tensor Norms and Operator Ideals, North-Holland Math. Studies, Vol. 176, Amsterdam (1993). R. Deville, G. Godefroy and V. Zizler, Smoothness and Renorming in Banach Spaces, Longman, Harlow (1993). J. Diestel, J. Fourie and J. Swart, The metric theory of tensor products, Grothendieck's Rdsumd revisited, in preparation. J. Diestel, H. Jarchow and A. Tonge, Absolutely Summing Operators, Cambridge Univ. Press (1995). J. Diestel and C.J. Seifert, The Banach-Saks ideal, Comment. Math. 21 (1979), 109-118, 343-344. J. Diestel and J.J. Uhl, Vector Measures, Math. Surveys, Vol. 15, Amer. Math. Soc., Providence, RI (1977). A.C. Dixon, On a class of matrices of infinite order and on the existence of "matricial" functions on a Riemann surface, Trans. Cambridge Phil. Soc. 14 (1901), 190-233.
492
J. Diestel et al.
[39] T. Domenig, Composition operators on weighted Bergman spaces and Hardy spaces, Thesis, Univ. Ziirich (1997). [40] E. Dubinsky, A. Petczy6ski and H.P. Rosenthal, On Banach spaces X for which 172 ( s X) = B ( s X), Studia Math. 44 (1972), 617-648. [41] N. Dunford and B.J. Pettis, Linear operations on summablefunctions, Trans. Amer. Math. Soc. 47 (1940), 323-392. [42] N. Dunford and J.T. Schwartz, Linear Operators, Vol. I, Interscience, New York (1958). [43] A. Dvoretzky, Some results on convex bodies and Banach spaces, Proc. Symp. on Linear Spaces, Jerusalem (1961), 123-160. [44] A. Dvoretzky and C.A. Rogers, Absolute and unconditional convergence in normed linear spaces, Proc. Nat. Acad. Sci. USA 36 (1950), 192-197. [45] P. Enflo, Banach spaces which can be given an equivalent uniformly convex norm, Israel J. Math. 13 (1972), 281-288. [46] P. Enflo, A counterexample to the approximation problem in Banach spaces, Acta Math. 130 (1973), 309317. [47] P. Erd6s and M. Magidor, A note on regular methods of summability and the Banach-Saks property, Proc. Amer. Math. Soc. 59 (1976), 232-234. [48] H. Fakhoury, Sur les espaces de Banach ne contenant pas l! (1~), Math. Scand. 41 (1971), 277-289. [49] T. Figiel, J. Lindenstrauss and V.D. Milman, The dimension of almost spherical sections of convex bodies, Acta Math. 139 (1977), 53-94. [50] V.R. Gantmacher, Ober schwache totalstetige Operatoren, Mat. Sbomik 7 (49) (1940), 301-307. [51] D.J.H. Garling and Y. Gordon, Relations between some constants associated with finite dimensional Banach spaces, Israel J. Math. 9 (1971), 346-361. [52] M. Girardi and W.B. Johnson, Universal non-completely-continuous operators, Israel J. Math. 99 (1997), 207-219. [53] N. Ghoussoub and W.B. Johnson, Counterexamples to several problems on factorization of bounded linear operators, Proc. Amer. Math. Soc. 92 (1984), 233-238. [54] E.D. Gluskin, Y. Gordon and A. Pajor, in: Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), to appear. [55] I.C. Gohberg and M.G. Krein, Introduction into the Theory of Non-Self-Adjoint Operators in Hilbert Space, Nauka, Moscow (1965) (Russian); Engl. transl.: Amer. Math. Soc., Providence, RI (1969). [56] Y. Gordon and D.R. Lewis, Absolutely summing operators and local unconditional structures, Acta Math. 133 (1974), 27--48. [57] A. Grothendieck, Sur une notion de produit tensoriel topologique d'espaces vectoriels topologiques, et une classe remarquable d'espaces vectoriels li~e gl cette notion, C. R. Acad. Sci. Paris 233 (1951), 1556-1558. [58] A. Grothendieck, Sur les applications lin~aires faiblement compactes d'espaces du type C(K), Canad. J. Math. 5 (1953), 129-173. [59] A. Grothendieck, Produits tensoriels topologiques et espaces nucl~aires, Mem. Amer. Math. Soc. 16 (1955). [60] A. Grothendieck, R~sum~ de la th~orie m~trique des produits tensoriels topologiques, Bol. Soc. Mat. Sao Paulo 8 (1956), 1-79. [61] A. Grothendieck, La th~orie de Fredholm, Bull. Soc. Math. France 84 (1956), 319-384. [62] A. Grothendieck, The trace ofcertain operators, Studia Math. 20 (1961), 141-143. [63] S. Heinrich, Finite representability and super-ideals of operators, Dissertationes Math. 172 (1980). [64] S. Heinrich, Ultraproducts in Banach space theory, J. Reine Angew. Math. 313 (1980), 72-104. [65] S. Heinrich, Closed operator ideals and interpolation, J. Funct. Anal. 35 (1980), 397-411. [66] D. Hilbert, Grundziige einer allgemeinen Theorie der linearen Integralgleichungen (Vierte Mitteilung), Nachr. Wiss. Gesell. G6tt., Math.-Phys. K1. (1906), 157-227. [67] E. Hille, Functional Analysis and Semi-Groups, Amer. Math. Soc., New York (1948). [68] J. Hoffman-JCrgensen, Sums of independent Banach space valued random variables, Aarhus Universitet, Matematisk Institut, Preprint Series 1972/73, No. 15. [69] A. Horn, On the singular values of a product of completely continuous operators, Proc. Nat. Acad. Sci. USA 36 (1950), 374-375.
Operator ideals
493
[70] C.V. Hutton, On the approximation numbers of an operator and its adjoint, Math. Ann. 210 (1974), 277280. [71] R.C. James, Some self-dual properties of linear normed spaces, Symposium on Infinite Dimensional Topology (Baton Rouge, 1967), Princeton Univ. Press, NJ (1972), 159-175. [72] R.C. James, Nonreflexive spaces of type 2, Israel J. Math. 30 (1978), 1-13. [73] G.J.O. Jameson, Summing and Nuclear Norms in Banach Space Theory, London Math. Soc. Student Texts, Vol. 8, Cambridge Univ. Press (1987). [74] H. Jarchow, Locally Convex Spaces, Teubner, Stuttgart (1981). [75] H. Jarchow, On weakly compact operators on C*-algebras, Math. Ann. 273 (1986), 341-343. [76] H. Jarchow, Some functional analytic properties of composition operators, Quaestiones Math. 18 (1995), 229-256. [77] E John, Extremum problems with inequalities as subsidiary conditions, Courant Anniversary Volume, Interscience, New York (1948), 187-204. [78] K. John, On the compact non-nuclear problem, Math. Ann. 287 (1990), 509-514. [79] W.B. Johnson, A universal non-compact operator, Coll. Math. 23 (1971), 267-268. [80] W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84. [81] M.I. Kadets and M.G. Snobar, Certain functionals on the Minkowski compactum, Mat. Zametki 10 (1971), 453-458 (Russian). [82] S. Kakutani, Iteration of linear operations in complex Banach spaces, Proc. Imp. Acad. Tokyo 14 (1938), 295-300. [831 S. Kakutani, Concrete representation of abstract (L)-spaces and the mean ergodic theorem, Ann. of Math. (2) 42 (1941), 523-537. [841 S. Kakutani, An abstract characterization of (M)-spaces, Ann. of Math. (2) 42 (1941), 994-1024. [851 T. Kato, Perturbation theory for nullity, deficiency and other quantities of linear operators, J. Analyse Math. 6 (1958), 261-322. [86] S.V. Kislyakov, On spaces with 'small' annihilators, Sem. Leningrad. Otdel. Mat. Inst. Steklov., Vol. 65 (1976), 192-195 (Russian). [87] H. von Koch, Sur quelques points de la th~orie des d~terminantes infinis, Acta Math. 24 (1901), 89-122. [881 H. K6nig, Grenzordnungen von Operatorenidealen, Math. Ann. 212 (1974), 51-64, 65-77. [89] H. K6nig, A Fredholm determinant theory for p-summing maps in Banach spaces, Math. Ann. 247 (1980), 255-274. [90] H. K6nig, Eigenvalue Distribution of Compact Operators, Birkh~user, Basel (1986). [91] G. K6the, Neubegriindung der Theorie der vollkommenen Riiume, Math. Nachr. 4 (1951), 70-80. [92] S. Kwapiefi, A linear topological characterization of inner product spaces, Studia Math. 38 (1970), 277278. [93] S. Kwapiefi, On operators factorable through Lp spaces, Bull. Soc. Math. France, M6m. 31-32 (1972), 215-225. [94] S. Kwapiefi, Isomorphic characterizations of inner product spaces by orthogonal series with vector valued coefficients, Studia Math. 44 (1972), 583-595. [95] H.E. Lacey, Isometric Theory of Classical Banach Spaces, Springer, Berlin (1974). [96] R. Latata and K. Oleszkiewicz, On the best constant in the Khinchin-Kahane inequality, Studia Math. 109 (1994), 101-104. [97] M. Ledoux and M. Talagrand, Probability in Banach Spaces, Springer, Berlin (1991). [98] D.R. Lewis, An isomorphic characterization of the Schmidt class, Compositio Math. 30 (1975), 293-297. [99] V.B. Lidski~, Non-self-adjoint operators with a trace, Dokl. Akad. Nauk SSSR 125 (1959), 485-488 (Russian). [100] EK. Lin, B-convexity of the space of 2-summing operators, Israel Math. J. 37 (1980), 139-150. [101] W. Linde, An operator ideal in connection with the Radon-Nikodym property of Banach spaces, Math. Nachr. 71 (1976), 65-73. [1021 W. Linde and A. Pietsch, Mappings of Gaussian measures of cylindrical sets in Banach spaces, Theory Probab. Appl. 19 (1974), 445-460. [103] J. Lindenstrauss, On a certain subspace ofl I , Bull. Acad. Polon. Sci., S6r. Math. 12 (1964), 539-542.
494
J. Diestel et al.
[104] J. Lindenstrauss and A. Petczyfiski, Absolutely summing operators in 12p-spaces and applications, Studia Math. 29 (1968), 275-326. [105] J. Lindenstrauss and H.E Rosenthal, The 12p spaces, Israel J. Math. 7 (1969), 325-349. [ 106] H.P. Lotz, Grothendieck Ideals of Operators on Banach Spaces, Lecture Notes, Univ. of Illinois, Urbana (1973). [107] D. Luecking and K. Zhu, Composition operators belonging to the Schatten ideals, Amer. J. Math. 114 (1992), 1127-1145. [108] R.D. Mauldin (ed.), The Scottish Book, Birkh~iuser, Boston (1981). [109] B. Maurey, Demonstration d'une conjecture de A. Pietsch, C. R. Acad. Sci. Paris, S6rie A 274 (1972), 73-76. [110] B. Maurey, Th~orkmes de factorisation pour les op~rateurs lin~aires gt valeur dans les espaces L p, Ast6risque 11 (1974). [ 111 ] B. Maurey, Systkme de Haar, S6m. Maurey-Schwartz, Expos6 2, l~cole Polyt6chnique, Paris (1974/75). [112] B. Maurey and G. Pisier, S~ries de variables al~atoires vectorielles ind~pendantes et propri~t~s g~om~triques des espaces de Banach, Studia Math. 58 (1976), 45-90. [113] V.D. Milman and G. Schechtman, Asymptotic theory offinite dimensional normed spaces, Lecture Notes in Math. 1200, Springer, Berlin (1986). [114] B.S. Mityagin and A. Petczyfiski, Nuclear operators and approximative dimension, Proc. ICM, Moscow (1966), 366-372. [115] EJ. Murray and J. von Neumann, On rings of operators, Ann. of Math. (2)37 (1936), 116-229. [ 116] J. von Neumann, Some matrix-inequalities and metrization of matrix-spaces, Tomsk Univ. Rev. 1 (1937), 286-300; see also Collected Works, Vol. IV, 205-219. [ 117] C. Niculescu, Absolute continuity in Banach space theory, Rev. Roumaine Math. Pures Appl. 24 (1979), 413-422. [118] W. Orlicz, Ober unbedingte Konvergenz in Funktionenriiumen, Studia Math. 4 (1933), 33-37, 4147. [119] J. Peetre, Sur la transformation de Fourier des fonctions a valeurs vectorielles, Rend. Sem. Mat. Univ. Padova 42 (1969), 15-26. [120] A. Petczyfiski, Banach spaces in which every unconditionally converging operator is weakly compact, Bull. Acad. Polon. Sci., S6r. Math. l0 (1962), 641-648. [121] A. Petczyriski, On strictly singular and strictly cosingular operators, Bull. Acad. Polon. Sci., S6r. Math. 13 (1965), 31-36, 3741. [122] A. Petczyfiski, A characterization of Hilbert-Schmidt operators, Studia Math. 28 (1967), 355-360. [123] A. Petczyfiski, p-integral operators commuting with group representations and examples of quasi-pintegral operators which are not p-integral, Studia Math. 33 (1969), 63-70. [ 124] A. Petczyfiski, Banach Spaces of Analytic Functions and Absolutely Summing Operators, Regional Conf. Series in Math., Vol. 30, Amer. Math. Soc., Providence, RI (1976). [125] A. Petczyfiski, Geometry of finite dimensional Banach spaces and operator ideals, Notes in Banach Spaces, Univ. of Texas Press, Austin (1980). [126] A. Persson and A. Pietsch, p-nukleare und p-integrale Abbildungen in Banachriiumen, Studia Math. 33 (1969), 19-62. [ 127] S.K. Pichorides, On the best values of the constant in the theorem ofM. Riesz, Zygmund and Kolmogorov, Studia Math. 46 (1972), 164-179. [ 128] A. Pietsch, Einige neue Klassen von kompakten linearen Abbildungen, Rev. Roumaine Math. Pures Appl. 8 (1963), 427-447. [ 129] A. Pietsch, Nukleare lokalkonvexe Riiume, Akademie-Verlag, Berlin (1965); Springer, Berlin (1972). [130] A. Pietsch, Absolut p-summierende Abbildungen in normierten Riiumen, Studia Math. 28 (1967), 333353. [131] A. Pietsch, Ideale von Operatoren in Banachriiumen, Mitteilungen Math. Gesellschaft DDR (1968), 1-13. [132] A. Pietsch, Adjungierte normierte Operatorenideale, Math. Nachr. 48 (1971), 189-211. [133] A. Pietsch, Theorie der Operatorenideale (Zusammenfassung), FSU Jena (1972). [134] A. Pietsch, Ultraprodukte von Operatoren in Banachriiumen, Math. Nachr. 61 (1974), 123-132. [135] A. Pietsch, Operator Ideals, Deutscher Verlag der Wissenschaften, Berlin (1978); North-Holland, Amsterdam (1980). [136] A. Pietsch, Factorization theorems for some scales ofoperator ideals, Math. Nachr. 97 (1980), 15-19.
Operator ideals
495
[137] A. Pietsch, Eigenvalues and s-Numbers, Cambridge Univ. Press (1987). [ 138] A. Pietsch and H. Triebel, Interpolationstheorie fiir Banachideale von beschriinkten linearen Operatoren, Studia Math. 31 (1968), 95-109. [139] A. Pietsch and J. Wenzel, Orthonormal Systems and Banach Space Geometry, Cambridge Univ. Press (1998). [140] G. Pisier, Sur les espaces de Banach qui ne contiennent pas uniform~ment de l~, C. R. Acad. Sci. Paris, S6rie A 277 (1973), 991-994. [ 141 ] G. Pisier, Un exemple concernant la super-reflexivitY, S6m. Maurey-Schwartz, Annexe 2,13cole Polyt6chnique, Paris (1974/75). [ 142] G. Pisier, Martingales with values in uniformly convex spaces, Israel J. Math. 20 (1975), 326-350. [143] G. Pisier, Une nouvelle classe d'espaces de Banach v~rifiant le th~orkme de Grothendieck, Ann. Inst. Fourier 28 (1978), 69-90. [144] G. Pisier, Un th~orkme sur les op~rateurs lin~aires entre espaces de Banach qui se factorisent par un espace de Hilbert, Ann. Sci. 13cole Norm. Sup., 4 e S6rie 13 (1980), 23-43. [145] G. Pisier, Holomorphic semi-groups and the geometry of Banach spaces, Ann. of Math. 115 (1982), 375392. [146] G. Pisier, Counterexamples to a conjecture ofGrothendieck, Acta Math. 151 (1983), 181-208. [147] G. Pisier, Probabilistic methods in the geometry of Banach spaces, Probability and Analysis (Varenna, Italy, 1985), Lecture Notes Math. 1206 (1986), 167-241. [148] G. Pisier, Factorization of Linear Operators and Geometry of Banach Spaces, Regional Conf. Series in Math., Vol. 60, Amer. Math. Soc., Providence, RI (1986). [149] G. Pisier, A remark on I-I2(1p , lp), Math. Nachr. 148 (1990), 243-245. [150] G. Pisier and Q. Xu, Random series in the real interpolation spaces between the spaces vp, GAFA 1985/86, Lecture Notes Math. 1267 (1987), 185-209. [ 151 ] D. Przeworska-Rolewicz and S. Rolewicz, Equations in Linear Spaces, Warszawa (1968). [152] O. Reinov, Operators oftype RN in Banach spaces, Dokl. Akad. Nauk SSSR 220 (1975), 528-531 (Russian). [153] O. Reinov, Approximation properties of order p and the existence of non-p-nuclear operators with a p-nuclear second adjoint, Math. Nachr. 109 (1982), 125-134. [154] M.A. Rieffel, Dentable subsets of Banach spaces, with application to a Radon-Nikodjm theorem, Functional Analysis, Gelbaum, ed., Proc. Conf. (Irvine, CA, 1966), Thompson, Washington (1967), 71-77. [155] F. Riesz, Les Systkmes d'Equations Lin~aires gt Une Infinit~ d'Inconnues, Gauthier-Villars, Paris (1913). [156] E Riesz, Ober lineare Funktionalgleichungen, Acta Math. 41 (1918), 71-98. [ 157] J.R. Ringrose, Compact Non-Self-Adjoint Operators, Van Nostrand, London (1971). [158] H.P. Rosenthal, On subspaces of L p, Ann. of Math. 97 (1973), 344-373. [159] H.E Rosenthal, A characterization of Banach spaces containing 11 , Proc. Nat. Acad. Sci. USA 71 (1974), 2411-2413. [160] A.E Ruston, On the Fredholm theory of integral equations for operators belonging to the trace class of a general Banach space, Proc. London Math. Soc. (2) 53 (1951), 109-124. [161] A.E Ruston, Direct products of Banach spaces and and linear functional equations, Proc. London Math. Soc. (3) 1 (1951), 327-384. [162] A.E Ruston, Fredholm Theory in Banach Spaces, Cambridge Univ. Press (1986). [163] E Saphar, Applications d puissance nucl~aire et applications de Hilbert-Schmidt, Ann. Sci. t~cole Norm. Sup., 3 e S6rie 83 (1966), 113-151. [164] E Saphar, Produits tensorielles topologiques et classes d'applications lin~aires, C. R. Acad. Sci. Paris, S6rie A 266 (1968), 526-528. [165] R. Schatten, On the direct product of Banach spaces, Trans. Amer. Math. Soc. 53 (1943), 195-217. [166] R. Schatten, The cross-space of linear transformations, Ann. of Math. (2) 47 (1946), 73-84. [ 167] R. Schatten, A Theory of Cross-Spaces, Annals of Math. Studies, Vol. 26, Princeton Univ. Press (1950). [168] R. Schatten, Norm Ideals of Completely Continuous Operators, Springer, Heidelberg (1970). [169] R. Schatten and J. von Neumann, The cross-space of linear transformations, Ann. of Math. (2) 47 (1946), 608-630; 49 (1948), 557-582. [170] J. Schauder, Ober lineare, vollstetige Funktionaloperationen, Studia Math. 2 (1930), 183-196.
496
J. Diestel et al.
[171] E. Schmidt, Zur Theorie der linearen und nichtlinearen Integralgleichungen, Math. Ann. 63 (1907), 433476; 64 (1907), 161-174. [172] C. SchUtt, Unconditional bases in Banach spaces of absolutely p-summing operators, Math. Nachr. 146 (1990), 175-194. [173] I. Schur, Uber die charakteristischen Wurzeln einer linearen Substitution mit einer Anwendung auf die Theorie der Integralgleichungen, Math. Ann. 66 (1909), 488-510. [ 174] I. Schur, Ober lineare Transformationen in der Theorie unendlicher Reihen, J. Reine Angew. Math. 151 (1920), 79-111. [175] L. Schwartz, Geometry and Probability in Banach Spaces, Lecture Notes in Math. 852, Springer, Berlin (1981). [176] H.U. Schwarz, Adjungierte Operatorenideale, Math. Nachr. 55 (1973), 293-308. [177] J.H. Shapiro, Composition Operators and Classical Function Theory, Springer, Berlin (1993). [178] J.H. Shapiro, The essential norm of a composition operator, Ann. of Math. 125 (1987), 375-404. [ 179] J.H. Shapiro, Composition operators and Schroeder's functional equation, Studies on Composition Operators, Jafari et al., eds, Contemporary Math., Vol. 213 (1998), 213-228. [180] J.H. Shapiro and P.D. Taylor, Compact, nuclear, and Hilbert-Schmidt composition operators, Indiana Univ. Math. J. 23 (1973), 471-496. [181] B. Simon, Trace Ideals and their Applications, London Math. Soc. Lecture Note Series, Vol. 35, Cambridge Univ. Press (1979). [182] C. Stegall, The Radon-Nikodym property in conjugate Banach spaces, Trans. Amer. Math. Soc. 206 (1975), 213-223; 264 (1981), 507-519. [183] A. Szankowski, B(H) does not have the approximation property, Acta Math. 147 (1981), 89-108. [184] S.J. Szarek, On the best constants in the Khinchin inequality, Studia Math. 58 (1976), 197-208. [185] M. Talagrand, Cotype of operators from C(K), Invent. Math. 107 (1992), 1-40. [ 186] N. Tomczak-Jaegermann, The moduli of smoothness and convexity and the Rademacher averages of trace classes Sp (1 < p < e~), Studia Math. 50 (1974), 163-182. [187] N. Tomczak-Jaegermann, Computing 2-summing norms with few vectors, Ark. Mat. 17 (1979), 273-277. [188] N. Tomczak-Jaegermann, Banach-Mazur Distances and Finite Dimensional Operator Ideals, Longman, Harlow (1989). [189] L. Weis, On the surjective (injective) envelope of strictly (co-)singular operators, Studia Math. 54 (1976), 285-290. [190] H. Weyl, Inequalities between the two kinds of eigenvalues of a linear transformation, Proc. Acad. Sci. USA 35 (1949), 408-411. [191] H. Whitney, Tensor products of Abelian groups, Duke Math. J. 4 (1938), 495-528. [192] P. Wojtaszczyk, Banach Spaces for Analysts, Cambridge Univ. Press (1991). [193] K. Yosida, Mean ergodic theorem in Banach spaces, Proc. Imp. Acad. Tokyo 14 (1938), 292-294.
CHAPTER
12
Special Banach Lattices and their Applications
S.J. Dilworth Department of Mathematics, University of South Carolina, Columbia, SC 29208, USA E-mail: dilworth @math. sc. edu
Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Lorentz L p,q spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . n into Lp,q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 9 Embedding ~q and ~p 4. Orlicz spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Containment of gp and the Nikishin factorization theorem . . . . . . . . . . . . . . . . . . . . . . . . . 6. Some probabilistic applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7. Embedding L w, q into Lq . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
H A N D B O O K OF T H E G E O M E T R Y OF B A N A C H SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 497
499 499 505 511 515 519 524 527
This Page Intentionally Left Blank
Special Banach lattices and their applications
499
1. Introduction
This article discusses certain Banach lattices of importance in analysis, particularly the Lorentz and Orlicz spaces. Special Banach lattices arise naturally in probability theory and in many areas of analysis: in interpolation theory, in Fourier analysis, and in functional analysis in the theory of absolutely summing operators, for example. They furnish a rich supply of interesting Banach spaces with nontrivial properties. Consequently, their Banach space structure has attracted considerable attention, and over time a large and diverse body of results has accumulated. The first section after this Introduction is a sketch of the basic theory of the Lorentz Lp,q spaces, covering duality, convexity and concavity, isomorphic classification of weak Lp spaces, and the connection with real interpolation. Section 3 discusses the subspace structure of the Lp,q spaces, particularly the possibility of isomorphically embedding g p and ~q into Lp,q. Section 4 is a sketch of the basic theory of Orlicz spaces, covering the Luxemburg and Orlicz norms, the duality theory of N-functions, isomorphic containment of co and g~, and the Matuszewska-Orlicz indices and their role in interpolation theory. The next section discusses the structure theory for subspaces of Orlicz sequence spaces. Here we also present the Nikishin factorization theorem, which is an important application of the space weak L p to functional analysis. This result is then used to extend to Orlicz spaces a famous result of Aldous stating that every subspace of L1 contains some g p subspace. Section 6 presents two applications of Orlicz and Lorentz spaces to probability theory. The first is a theorem of Rodin and Semenov identifying the closed linear span of a Bernoulli sequence in exponential Orlicz spaces. The second is a version of Rosenthal's inequality for moments of sums of independent mean zero random variables, a result which has had important applications to Banach space theory. The final section discusses the existence of isometric or isomorphic embeddings of Lorentz spaces into L p spaces. This article is intended to be a self-contained introduction to the area. Proofs are given of the key points of many of the results presented, although very technical issues have usually been side-stepped. These considerations as well as personal taste have inevitably influenced the selection of topics. Many results which could not be presented in detail are discussed briefly in the endnotes which conclude each section. The interested reader will find there a partial guide to the extensive literature in this area. Finally, a few words of thanks: to William B. Johnson and Joram Lindenstrauss for the invitation to write this article, and to my collaborators. Special thanks to Neal Carothers for help with the bibliography and for sending me his seminar notes which shaped the sections on the Lp,q spaces.
2. Lorentz
Lp,q spaces
For a measurable function f on (0, o<~) its distribution function df is given by df(t) = m(lf(x)l > t) (0 < t < ~ ) , where m denotes Lebesgue measure. The decreasing rearrangement of f is given by f*(x) = inf{t ~> 0: df(t) <<,x}. Note that f * is a non-negative, right-continuous, decreasing function on (0, cx~) with the same distribution function as f .
S.J.Dilworth
500
For 0 < p < c~ and 1 ~< q < c~, the Lorentz space Lp,q(1), where I is a subinterval of (0, c~), is the collection of all measurable functions f on I such that
Ilfllp,q -
(fo
f*(t) q d(tq/p)
)l/q=p(fo q
c~f*(t)q t q-1 dt
Integration by parts yields a useful alternative expression for
Ilfllp,q =
(foCX~(df(t)) q/p
d(tq)
)l/q
< cx~.
(1)
IIf IIp , q "
)l/q
(2)
For 1 < p < cx~, the space weak Lp, denoted L p , ~ , consists of all measurable f on I for which Ilfllp,~ -- sup tl/p f* (t) = suptdf(t) 1/p < ~ t>0
(3)
t>0
After identifying functions which are equal almost everywhere, the L p,q spaces become quasi-Banach spaces when equipped with the quasi-norms (1) and (3). For 0 < q < p, observe that the weight function w(t) - - t ( q / p ) - I is decreasing, and so in the range 0 < q < p
Ilfllp,q = sup
(f f(~r(t)) q d(t q/p) )'/q
(4)
where the supremum is taken over all measure-preserving transformations of (0, ~ ) . (For the range p < q < cx~, the weight is increasing, and so the supremum in (4) should be replaced by an infimum.) The sequence space (q < OQ) consists of all sequences (an)n~=l equipped with the norm
~p,q (cx~ )l/q II(an)l[p,q--~akq(kq/P-(k-1)q/p)
(s)
k=l
The sequence space ~ p,~ (weak e p) is equipped with the quasi-norm
II(an)llp,~ - sup nl/p a n.*
(6)
Applying Minkowski's inequality in the Lebesgue space Lq(d(tq/P)), we see from (4) that II. IIp,q satisfies the triangle inequality when 1 ~< q < p. For 1 < p < q < cx~, however, the triangle inequality is not satisfied and (1) defines only a quasi-norm. Nevertheless, this quasi-norm is equivalent to a norm, a consequence of 'Hardy's inequality'.
Special Banach lattices and their applications
501
THEOREM 1. For q ~ 1 and ~ > O, and for every non-negative measurable function g on
(0, cx~), we have g(u) du
t -~-1 dt
<~ q
[tg(t)] q t -~-1 dt
Ol
.
(7)
PROOF. Jensen's inequality applied to the probability measure d ( ( u / t ) ~/q) on [0, t] gives
g(u)du (fo')~ -,(~ -- fo'g(u)ul--4"d((u/t) ~/q) )~ Ol
< ~ ( q_Ol) q t ~ fot g(u)quq-~ d((u/t) ~/q) --
t ~ ( 1 - ~/)
g(u)qu q-~+q-1 du.
Thus,
foe~ (fot g(u) du)qt-~-I dt
~(q)~'f0 ~ t -l -4"(f0'~ug(u )1)'-+"du qu
-
Ol
1--or+ ~
=
(q)q-' ji~[ )]
=
(~)q fo~ (ug(u))qu-l-~du"
ug(u
q U-
q
(f~
-4
t
_1 _ ~
dt
)
-4 dt du
For t > 0, let f** (t) = (1 / t) fo f * (u) du, and define
[]fll(p,q) --
(fo ~ f**(t) d(tq/p) )'~q
(8)
It is readily verified that ( f + g)** ~< f** § g**, and so by Minkowski's inequality (8) defines a norm for all 1 < p < cx~ and 1 ~< q < e~. Moreover, an application of (7) with ot = q(1 - ( I / p ) ) yields
Ilfllp,q ~ [[fll(p,q) ~
P
p-1
Ilfllp,q.
(9)
For the ranges p ~< 1 and q < 1, however, L p,q is a non-locally convex quasi-Banach space. Under the natural pairing
( f ' g) -- fI f (t)g(t) dt,
S.J. Dilworth
502
the Banach space dual of Lp,q c a n be identified with Lp,,q,, where pl and q' are the H61der conjugate indices of p and q, and the dual norm is equivalent to II 9IIff,q'. More precisely, for 1 < p, q < cx~ there exists a positive constant c (depending on p and q) such that
c IIg IIp',q' ~ IIg IIL~,q ~ IIg IIp',q'.
(10)
The constant c arises from an application of Hardy's inequality which occurs in the proof of (10). In particular, the dual norm is not isometrically a n L p,q norm when q > 1. For Lp,1, on the other hand, the dual space is isometrically isomorphic to weak Lp, equipped with the norm: IIf II(p', cx3) - - sup t 1 / p' f** (t). t>O
( 11 )
The L p,q spaces arise naturally in the real interpolation method which we briefly review. An interpolation couple is an ordered pair (X0, X1) of Banach spaces of which both members are continuously embedded into some common larger topological vector space. For each f belonging to the sum space Xo 4- X1, and for each t > 0, the K-functional is defined thus:
K ( f , t; Xo, X1) = inf{ Ilgllxo + tllhllx," f - g 4- h}.
(12)
For 0 < 0 < 1 and 0 < q < cx~, the interpolation space (X0, X1)O,q consists of all f E X0 4- X1 such that
]lfllo,q --
(fo
[t - 0
K(f, t; Xo, X l ) ] q
< cx~.
(13)
For 0 < 0 < 1, the interpolation space (X0, X1)o,~ consists of all f 6 X0 + X1 such that [[fll0,e~ - supt -~ K ( f , t) < c~. t>0
(14)
For fixed t > 0, the K-functional defines a norm on X0 + X1. Thus, by Minkowski's inequality, (13) and (14) define norms on the interpolation spaces (Xo, X1)o,q and (Xo, X1)o,~. Completeness of 11. [[O,q can be shown to follow from the completeness of the norms of X0 and X1. Suppose that (Y0, Y1) is another interpolation couple and that T is a linear operator which simultaneously maps X0 into Y0 and X1 into Y1 continuously, with operator norms M0 and M1, respectively. The Marcinkiewicz interpolation theorem asserts that T is a continuous linear operator from (X0, X1)O,q into (Y0, Y1)O,q, and that the operator norm of T on (X0, X l)O,q is bounded by M 1-~ m ~ The couple (L 1, L ~ ) is especially important for the Lp,q spaces. It is easily seen that
K(f,t; L1,L~)--
fo t f * ( u ) d u - t f * * ( t ) .
Special Banach lattices and their applications
503
Hence for the couple (L1, L ~ ) , we have
IlfllO,q
-
(fo
[tl_ Of**(t)] q -idt
)
1/q -
(q)l/q Ilfll(p,q)
for p - 1/(1 - 0). The following more general result will be used in Section 3. THEOREM 2. Let 0 < Pl < P2 <~ cx~, 0 < 0 < 1, and 0 < q <~ ~ . Then, up to an equivalent norm, [Lpl (I), Lp2 (I)]o,q -- Lp,q(I), where 1/p - (1 - O)/pl + O/p2. For all q, the L p,q norm behaves like the L p norm under dilation: i.e., for )~ > 0, For each fixed p E (1, cx~), we obtain a scale of Banach spaces indexed by q E (1, ~ ) . To compare the spaces in this scale, let us compute the norm of the step function y~'km__1 akZn,k, where zn,k is the indicator function of the interval [(k - 1)/n, k/n]"
[]f('/)Ol[p,q --)l/pllf(.)llp,q.
m
m
Z ai Zn,i k=l P,q
-f -- f
lak Iq Zn,i d(t q/p) k=l
k=l
ak,q -- ak+l) *q
Zn,i ] d ( t q/p) i=1 m
*q - ak+l)kq/P -- y ~ a k*q (k q/p
-
k=l
(k - 1)q/P).
k=l
The last expression can be shown to be a decreasing function of q by arguing by induction on m. Thus, for 1 ~< ql ~< q2 ~< ~ , we obtain the following monotonicity property of the L p,q norms: Ilfllp,~ ~ Ilfllp,q2 ~ Ilfllp,q~ ~ Ilfllp,1.
(15)
A similar summation by parts argument shows that if X is a rearrangement-invariant function space on I such that IIZE IIx = m ( E ) 1/p for every indicator function XE, then
Ilfll(p,~) ~ Ilfllx ~ Ilfllp,1.
(16)
Thus, Lp,e~(1) (equipped with the norm I1" II(p,~)) and Lp,1 (I) are the largest and smallest rearrangement-invariant Banach function spaces on I whose norms coincide with the L p norm on indicator functions. Observe at this juncture that it is often convenient to use (1), as we may, to define Lp,q ($2, •, #) for an arbitrary a-finite measure space. Many of the results discussed above extend straightforwardly to this more general setting. Let us now mention a typical application of the Marcinkiewicz interpolation theorem. Recall that the Fourier transform f ~-+ f maps L1 (Nn) continuously into L ~ ( R n ) , and
S.J. Dilworth
504
that by Plancherel's theorem it is an isometry on L2(Rn). Observe that from (15) it follows that Lp,,p(I) C Lp,(l) for 1 < p < 2, where 1/p + 1/p' = 1. Applying Theorem 2 we obtain the following refinement of the classical HausdorffYoung inequality (replacing L p, by the smaller space L p,, p). COROLLARY 1. Let 1 < p < 2. The Fourier transform maps Lp(~ n) continuously into L p,, p (]Rn). Next we consider the Banach lattice properties of convexity and concavity for L p,q. THEOREM 3. If 1 <. q ~ p < eC, then Lp,q is q-convex, i.e., for all (~ )i=l n in Lp,q, we have
IJSI q
)'Jq ("ll llqq )
l/q
<.
i=1
(17)
i=1
P,q
If the f i 's are disjointly supported then there is also a lower p-estimate:
/>
i=1
(18)
IIf/ IlpP,q
i=1
P,q
For p < q, Lp,q is q-concave and satisfies an upper p-estimate for disjoint functions. PROOF. First we use the triangle inequality in Lp/q,1 to prove q-convexity:
)'Jq q
If/I q
i=1
k
~ lfil q i=1
P,q
/7 IIIj~lqllp/q,~
p/q,1
i=1
=
Z ]If/" 11qp,q" i--1
For the lower p-estimate, we use (2) and the reverse triangle inequality in Lq/p(d(tq)) (note that q / p < 1)" tl
~f/ i=1
,
p,q
--
(f0
(d~n
fi (t)) q/p d(tq)
>>.~l[df,(t)llLq/~ i--1
)
p/q
~d~(t) i--1
Lq/p(d(tq))
~ ]]f/llp,q. i=1
We conclude this section with a discussion of the isomorphic classification of Lorentz spaces. It is well-known that Lp(0, 1) and Lp(O, cx~) are linearly isometric Banach spaces. For Lp,q spaces, however, it was shown in [22,23] that (for p ~ q and q < cx~) Lp,q(O, 1) and L p,q(O, CX)) are not even isomorphic as Banach spaces. Nevertheless, using the Petczyfiski decomposition method and some intricate combinatorial arguments, Leung [90]
Special Banach lattices and their applications
505
proved the surprising result that the three classical weak L p spaces are all isomorphic to each other. THEOREM 4. Let 1 < p < cx~. The spaces gp,~, Lp,~(0, 1) and Lp,~(O, Cx3) are isomorphic Banach spaces. Notes. Hardy [58] proved the first version of Theorem i (for sequences). The spaces Lp,1 and weak L p were first introduced by Lorentz [ 100] with somewhat different notation as the spaces A(ot) (c~ = 1/p) and M(o~) (or = 1 - 1/p), respectively. The Lp,q spaces for I ~< q ~< p were introduced in [101] as members ofthe larger class of Lw,p spaces (see Section 7). The Lp,q notation was introduced by Calder6n (see [117, p. 129]). See [69] and its references for linear topological and interpolation-theoretic properties of the Lp,q spaces. The injective isometries of Lp,1 (0, 1) were determined by Carothers and Turett [31 ]. For the real method of interpolation and the Marcinkiewicz interpolation theorem (as well as the Lp,q spaces) see the texts [10] and [11]. Corollary 1 is taken from [135]. Theorem 3 was proved by Creekmore [38] who used it to determine the type and cotype of the Lp,q spaces. An example of Pisier [99] shows that Lp,q (1 ~< q < p) is not p-concave. In particular, L2,q is not of cotype 2. An interpolation argument is given in [53] to show (for 1 < p < 2 and 1/p + 1/pl = 1) that Lp,p, has Fourier type p. The isomorphic classification of the weak Lp spaces Lp,~(S-2, z~, lZ), where ($2, ZT, #) is an arbitrary measure space, was carried out by Leung in a series of papers [88,90-92]. If (S2, Z', #) is countably generated and o--finite then the corresponding weak L p space (when infinite-dimensional) is isomorphic to either s or ep,~ [91]. For an uncountable purely atomic measure space there are five distinct classes of isomorphism types [91]. The weak L p spaces corresponding to purely non-atomic measure spaces which are not countably generated are classified in [92]. In [89] it is proved that s (1 < p < cx~) contains a complemented Hilbertian subspace. Kalton [77] proved that (for 1 ~< p < 2) L p, ~ (0, 1) contains a complemented subspace isometrically isomorphic to L p (0, 1). It was proved in [39] that weak L1 has a nontrivial dual space. Lotz [ 100] proved that weak L p (1 < p < oo) is a Grothendieck space, i.e., that weak and weak-star convergence coincide for sequences in the dual of weak L p. The seminar notes [21] influenced this discussion of the Lp,q spaces.
3. Embedding s and s t i into L p,q The results of this section will be proved for Lp,q(O, ~ ) (denoted simply Lp,q) as this case requires rather careful reasoning and is more general than both L p,q (0, 1) and ~p,q. First we prove the existence of subspaces isomorphic to s The following result gives sufficient conditions for the existence of a sequence equivalent to the the unit vector basis Of s PROPOSITION 1. Suppose that 0 < p, q < c~. Let (fn) be a sequence of unit vectors in Lp,q such that f * ~ 0 pointwise. Then some subsequence of (fn) is equivalent to the unit vector basis of (q.
S.J. Dilworth
506
PROOF (Sketch). The hypothesis ensures that we can find a subsequence of (fn) which is arbitrarily close in the L p,q n o r m to a disjoint sequence of compactly supported functions. Thus, without loss of generality, we assume that the fn's are disjoint and compactly supported. Let (en) be a decreasing null sequence of positive numbers, and let Un = {Ifnl < en}, gn = fnXu,, and hn = f n - gn. Clearly, either limsup Ilgnllp,q > 0 or limsup Ilhn IIp,q > 0. If the former is the case, then (passing to a subsequence) we may assume that inf Ilgn IIp,q - - ~ > 0. We shall show that (gn) has a subsequence equivalent to the unit vector basis of s proving this first for the range p < q. By Theorem 3, there is a lower q-estimate for disjoint functions (in fact, L p , q is q-concave), and so
lanl q
•
~<
an gn
n=l
n=l
P,q
Obtaining an upper estimate will be the heart of the proof: to this end, choose (for each n ~ 1) Sn > 0 such that fSn+m(Un)
Ig*(t - Sn)] q d
(t q/p
q
(19)
) -- 2llgnllp,q.
J Sn
Clearly, Sn ~ 00, since Ign I < 6n ~ O. So (passing to a subsequence) we may assume that Sn + m(Un) < Sn+l. Now (19) implies the existence of a measure-preserving transformation cr of (0, ec) such that cr([Sn, Sn + m(Un)]) -- Un and f Sn+l
Ign(cr(t))l q d
(tq/p
q
) - 2llgnllp,q.
,1sn
Hence, by (4) for p < q, we have
• n--1
oo
an gn
<~ P,q
f0
oo
~
la~lqlgn(Cr(t))l q d(t q/p)
n=l
0(3
f Sn+1
= ~lanl q
Ig,(~(t))l q d(t q/p) a Sn
n=l o(3
oo
= 2 y ~ lan Iq Ilgn IIq,q ~< 2 Z n--1
lan Iq.
n=l
This completes the proof of the upper q-estimate. An analogous argument shows that if lim sup [Ihn IIp,q > 0, then (hn) has a subsequence equivalent to the unit vector basis of s Finally, the case q < p can be handled in a similar fashion: here the upper estimate follows from Theorem 3, and the lower estimate is obtained as above by considering a suitable measure-preserving transformation. [3 The following theorem is the main result on the containment of s subspaces: the closed linear span of every disjointly supported sequence in Lp,q contains a subspace isomorphic to
s
Special Banach lattices and their applications
507
THEOREM 5. Suppose that 0 < p, q < c~. Let (fn) be a disjointly supported sequence of
unit vectors in Lp,q. Then [fn] contains a subspace isomorphic to g,q. PROOF. After passing to a subsequence, we may assume by Helly's selection theorem [111, p. 221] that ( f * ) converges pointwise (except for a countable set of points) to some right-continuous decreasing function g, i.e., that fn -- gn -+-hn, where gn* -- g and hn ~ 0 in measure. If g = 0, then the result follows at once from Proposition 1. So we shall assume that m(lgl > 23) > 26 for some 6 > 0. Since hn --+ 0 in measure, we may also assume that m(lhnl > en) < en, where (e,) is some sequence of positive numbers satisfying ~nC~=l En < 6. This yields m(lfnl > 3) > ~ for all n. We shall prove the following claim: given e > 0 and k ~> 1 there exist n ~> k and ~--,n+N N ~> 1 such that F - z--,i=n+l fi satisfies m ( l F I / > 2ellFllp,q) ~< 2e. An obvious induc,~-~n k+ l tion will then yield a sequence Fk - ',z---,i=nk+l (V'~k+' J~)/ll ---,i=~ +1 f/II P,q such that Fk --+ 0 pointwise. Then Proposition 1 will produce an ~q subsequence of (Fk), which proves the theorem. To establish the claim, fix e > 0 and k ~> 1. Since the f~'s are disjointly supported, and since m(Ifnl > 6) > 6 for each n, we get
ellFllp,q >>,~3llx[o,N~)llp,q - eB(N6) 1/p -- (31+(1/P)g)N 1/p.
(20)
On the other hand, since IIgx[o,e/Nlllp,q --+ 0 as N --+ c~, we have for all sufficiently large N (depending on e and 6) ( 3 1 + ( 1 / P ) ~ ) N 1/p ~
-N- ) E
/p
IIgx[o,e/N) llp,q (21)
Ilgx[o,e/N)llp,~ >- g(e/N).
Combining (20) and (21) we obtain e][F]]p,q > / g ( e / N ) for all sufficiently large N. Now choose n ~> k such that Y~i=n+l Ei < E. Then for all sufficiently large N, we have n+N
m(IFl>ZellFIIp,q)
--
~
m(IJ~l>ZellFIIp,q)
i=n+l
n+N
<" Z
n+N
m(]gi] > e]]F]]p,q) + Z
i--n+l
n+N
n+N
<~ y ~ m(]gi] > g ( e / N ) ) + ~ i=n+l
m([hi] > e[]F]]p,q)
i--n+l
~i
i=n+l
(since e II F II p,q >/ gi)
<~ N ( e / N ) + e -- 2e. This completes the proof of the claim.
E]
S.J. Dilworth
508
Embedding s into L p , q is quite a different story: we shall show that there is no embedding of s into L p,q if p 5~ 2, q < cx~, and p ~ q. But first we give an answer to the local version of this question by providing a quantitative estimate of the dimension of s k subspaces of/~np,q" n ~or 1 < p
L
--
akek
k--1
a ; q (k q/p -- (k - 1) q / p )
p,q
)"q
k--1
By the Mean Value Theorem
L
a;q(k(q/P )-11
akek
k=l
P,q
(22)
k=l
where the constants of equivalence depend only on p and q. Let 13 > 0 and set m = [nr The theorem is equivalent to the statement (which we shall prove) that the spaces s contain uniformly complemented subspaces that Set v = ~ i =mnl i - 1 / P e i and, for 1 ~< j <~ n, set vj = ~im=l i - 1 / P e i + ( j _ l ) n . O b s e r v e that the vj's are disjointly supported and that they all have the same distribution. Thus, since L p,q satisfies a lower p-estimate (Theorem 3), we have are uniformly isomorphic to s n
[[Vl ]]p,q j=l
]aj] p j=l
P,q
n To complete the proof t h a t (l)j)j= 1 is uniformly equivalent to the unit vector basis of s /7 it suffices to prove the upper estimate
•
ajoj
j=l
~< CIIvl Ilp,q p,q
lajl p j--1
with C depending only on p and q. By (22)
i-q/Piq/p-I
I[vllqp,q ~ i=1
) /7
and so it is enough to prove that [[ ~ j = l
,~ logmn ~< (1 +/3) logn -~ Ilvl Ilq,q,
ajvj[]p,q
To this end, suppose that aj ) 0 and that ~ = l
n
~ [IVl]p,q whenever ~-'~j=l [ajl p -- 1.
ajP -- 1. Fix 1 <<.i <<.mn, and, for each
509
Special Banach lattices and their applications
1 ~<j ~
Kj -- max{k: ajk -1/p > i -1/p}. /7 Then Zj=I
K j represents the total number of coordinates of the
vector Y ~/7j = I p
a j v j which
n
exceed i -1/p (-- v(i), the ith coordinate of v). Thus, i -- ~-~i%1 iaj > Y~j=I K j , and so
ajvj
(i) <. i -1/p -- v(i),
j=l g/
for each i, which yields 11Zj=l a j v j [[p,q ~ [[IJllp,q as desired. The proof of uniform complementation is straightforward. Let vj be a Hahn-Banach norming functional for vj with the same support as vj, and set P/7 -- Y~=I (', v~)vj/llvj [[p,q. Then P/7 is a projection onto [vj] nj = l ' and
]]P,(x)] pP,q ~ Ll<x j-1
*>1p ~ Ilxllp,q p ,
, Vj
since Lp,q satisfies a lower p-estimate. Finally, the case p < q follows by duality.
V1
We end this section by proving the important fact that there is no embedding of g p into L p,q (except when p -- 2, p = q, or q -- oc). The proof requires the concept of a disjoint sum of functions. Given finitely many real-valued functions fl . . . . . fn defined on [0, 1] (or, more generally, defined on some probability space (s 27, P)), we define the disjoint sum, denoted }-~in 1 O f , , to be any function f on (0, oc) for which d f - ~i~=1 dfi; for example, take f (t) -- ~-~in=l fi (t - i + 1 ) X [ i _ l , i ) ( t ). The disjoint sum concept is helpful for formulating certain norm inequalities and will also be used in Section 6. PROPOSITION 2. Let 0 < p < 2 and 0 < q <~ oc. There exists C > 0 (depending only on p and q) such that
<~C i=1
P,q
Lej~ i=1
(23) P,q
for all n ~ 1 and for all fl . . . . . fn in Lp,q(O, 1). PROOF. Consider the linear mapping Tn on Lp(O, oc)defined by T n ( f ) - ~--~i%l fi, where ~ ( t ) - f ( t + i - 1)X[0,11(t). Then, for 0 < p < 1, since II" lipp satisfies the triangle inequality, we have
I T n ( f ) Ipp <~
LIj~I i=1
p
/7
-
Lej~ i=1
-II/11 pp. P
(24)
510
S.J. Dilworth
By the Marcinkiewicz interpolation theorem (Theorem 2), Tn is bounded on each L p,q space, for 0 < p < 1 and 0 < q ~< e~, with norm C 2 depending only on p and q. Thus, for 0 < p < 2, we can apply (24) in the space Lp/2,q/2 to obtain
/2)'J2ii
1/2 _.
1/2
~
s
p,q
p/2, q/2
i=1
= c s
o fi p/2, q/2
i=1
.
p,q
i=1
THEOREM 7. Suppose that 0 < p, q < cx~, that p =/=2, and that p ~ q. Then gp is not
isomorphic to a subspace of Lp,q (0, ~ ) . PROOF. We begin with the range 0 < p < 2 and p < q, which is the most interesting and delicate case of the theorem. Arguing for a contradiction, suppose that (j~) is equivalent to the unit vector basis of p. By approximation we may assume that each fi is a step function supported in a finite interval. Let (gi) be a disjointly supported sequence with gi having the same distribution as f/ for each i. Since (fi) is an unconditional basic sequence, the Maurey-Khintchine inequalities yield 1/2
lail p
s
i--1
(25)
ai fi
i-1
P,q
i=1
P,q
For each fixed n we may suppose, by applying a dilation (which is an isometry in Lp,q), that each j~ is supported on [0, 1], and hence that Z i n= I aigi has the same distribution as Y~in=1 EDai ft'. Thus, by Proposition 2 for the left-hand inequality, and the upper p-estimate for disjoint functions (Theorem 3) for the right-hand inequality, we have
in)
112 2 2
ai fi i=1
~ C ( ~_~ lai lp
aigi
<<.c
p,q
i=1
p,q
lip (26)
i=1
Combining (25) and (26) we see that (gi) is a disjoint sequence in Lp,q that is equivalent to the unit vector basis of g p, which contradicts Theorem 5 since g p does not contain a subspace isomorphic to gq. Let us now consider the case q < p < 2. Using Theorem 5 to dispose of the case of a disjointly supported sequence, a routine 'Kadets-Petczyfiski' argument [73] (cf. the proof of Theorem 14 below) shows that it suffices to prove that there is no strongly-embedded subspace of Lp,q(O, 1) (one for which the L0-topology and the Lp,q-topology coincide) isomorphic to g p. But by (15) every such strongly-embedded subspace would embed isomorphically into L p,q,(O, 1) for all q~ > p > q, which we have shown to be impossible.
Special Banach lattices and their applications
511
For the remaining case p > 2, it suffices to observe that every strongly-embedded subspace of L p,q (0, 1) is isomorphic to a Hilbert space because of the continuous inclusion Lp,q(O, 1) c L2(0, 1). F1
Notes. Proposition 1 is from [24]. Its analogue for Lorentz sequence spaces was proved in [95] (see also [6]) and for Lp,q(O, 1) in [51]. Levy [93] proved that the interpolation space [X0, X1]o,q contains ~q whenever X0 A X1 is not closed in X0 + X1. Theorem 6 is due to Carothers and Flinn [30] who also gave the upper bound of C(p, q)n/(logn) I1/p- 1/ql for the dimension of uniformly complemented s k subspaces n . Theorems 5 and 7 are taken from [24] (see also [22]). The fact that for p < 1 o f ~p,q and p < q there is no embedding of s into Lp,q(O, 1) was proved by Kalton [76]. Carothers [19,20] proved that, for p > 2 and 1 ~< q <~ p, the only rearrangementinvariant spaces on [0, 1] which are isomorphic to subspaces of Lp,q (0, 1) are Lp,q (0, 1) and L2(0, 1) (the case p - - q > 2 was proved by Johnson, Maurey, Schechtman and Tzafriri [70]). He also proved (using results of O'Neil [ 118]) that this result does not extend to the range p < q (or p < 2).
4. Orlicz spaces
Let p : [ 0 , ~ ) --+ [0, cx~) be a non-decreasing right-continuous function satisfying p(0) = limt~0+ p(t) = 0 and p(cx~) = l i m t ~ p(t) = ~ . Then t
49(t) --
L
p(u) du
is a continuous convex function satisfying limt~0+ ck(t)/t = 0 and l i m t ~ ck(t)lt = cx~. We say that 4~ is an N-function. The Orlicz class Lg (0, e~) consists of all measurable functions f on (0, cx~) for which
M
oo
L
cb(If ( x ) l ) d x < ~ .
The Orlicz space L~(0, cx~) (denoted simply L~) is the linear subspace generated by L~ (0, cxz) in the space of measurable functions. It consists of all f such that M4~(kf) < cxz for some k > 0. The set {f E L~: M4~(f) <<.1} is a convex set (from the convexity of 4~) which absorbs L~. Its gauge functional Ilfllq~ - inf{k > 0" M4~(f/k) <<.1}
(27)
defines the Luxemburg norm on L~, which is a complete norm enjoying the 'Fatou property', i.e., if fn increases pointwise to f as n --+ cx~, then [Ifn [l~ --+ I[f ]l~. The collection of functions f such that k f E L4~ for all k > 0 is a closed subspace of L~ denoted L~. The simple integrable functions form a dense subspace of L~, and L~ consists precisely of the functions f E L~ for which the norm is absolutely continuous, i.e., if Ifnl <<.If] and fn ~ 0 pointwise, then Ilfn II~ --+ 0.
S.J. Dilworth
512
More generally, given any cr-finite measure space (X, Z, #), one can define the corresponding Orlicz class Lg (X) and Orlicz spaces L~ (X) and L~ (X). Of special interest and importance are X = [0, 1] with Lebesgue measure and X = N with counting measure. The latter gives rise to the Orlicz sequence space g4~ with norm given by
II(an)ll~ - inf
~(lanl/k) < oo .
>
(28)
n=l
Let q(t) = sup{u: p(u) <~t} be the right-continuous inverse of p. Since q satisfies the same assumptions as p, t
~(t) --
L
q(u) du
is an N-function, called the complementary function to q~ (denoted also ~b*). Observe that 4~ is also the complementary function to 7t, so that 4~** = q~. The key result relating q~ and 7t is Young's inequality: for all s, t > 0, we have
st ~ ~(s) + 7t(t)
(29)
with equality if and only if either t = p(s) or s = q(t). (When p(t) is continuous and strictly increasing, one can easily check this inequality by sketching a graph of p(t) and interpreting (29) in terms of areas under graphs.) Young's inequality yields a useful alternative expression for the complementary function 7t: 7t(t) = supst - 4~(s).
(30)
s>0
The associate space (L4~)I is the collection of measurable functions g for which the Orlicz norm
IIg II~ = sup
If0
Ifgl dx: IIf I1~ ~ 1
/
(31)
is finite.
THEOREM 8. L V, = (L4)' and
Ilgllr ~ Ilgll~ ~ 211gll~ for all g ~ L~v. PROOF. To prove the right-hand inequality, suppose that Ilfll~ ~ 1 and Ilgll~ ~ 1 (so that M ~ ( l f l ) ~< 1 and Mo(lgl) ~< 1). Applying Young's inequality, with s = If(x)l and
Special Banach lattices and their applications
513
t = Ig(x)l, and then integrating yields
f
oo Ifgl dx <<.M ~ ( I f l ) + MT, (Igl) ~< 2,
and so [Igll# ~< 2. For the left-hand inequality, the Fatou property allows us to assume that g is a simple integrable function with I[gllV~- 1. Let f ( x ) --q(lg(x)l), and note that IIf [l~ < ex). From the case of equality in (29), we have
I f (x)g(x)l = q(Ig(x)l)lg(x)l
-
4>(IS(x)l)+ 7x(Ig(x)l).
(32)
Observe that from (27) and from the convexity of 4b, the Luxemburg norm of any function h satisfies Ilhll~ ~< max(l, M
1 + Mr
>~ Ilfllq~ ~
L
Ifgl dx = M r
+ Mo(g).
So M~l(g) <~ 1, i.e., Ilgll~ ~ 1. Fatou's Lemma now gives (Lo)' = L#. Observe that ]IXAI[o -- 1/cb-l(1/m(A)) --+ 0 as m(A) --+ O. Since the simple functions are dense in L~, the Radon-Nikodym Theorem yields the following corollary of Theorem 8. COROLLARY 2. (L~,
I1" I1~)* is isometrically isomorphic to (LTx, I1" I1~).
The N-function 4~ is said to satisfy the A2~ condition (respectively A~~ condition) if
limx~oCb(2x)/cb(x) < oo (respectively limx_+oodp(2x)/dp(x) < ex)). It is easily seen that L~ (0, ex)) = L~ (0, ex)) if and only if 4~ satisfies b o t h / t o and/t~o. Suppose, for example that 4b fails the '42oo-condition. Then there is a sequence Mn "Poo with y~'n~176~(Mn)/4J(2Mn) < 1. Let (An) be a sequence of disjoint measurable sets with m ( a n ) - 1/4b(2Mn). Then
Mq$ ~ M n X A n ~=l
--
~<1, n=l 4~(2M~)
II ~-~n--1-+-(MnXAn)II~ ~ 1. On the other hand, Mcb(2MnXAn)- 1, and so IlMn XZn II - 1/2. Setting fn = Mn XAn it follows that
and so
1
~ll(an)ll~
oo
Z
-+-anfn
II(an)ll
n=l
for all (an) E g-oo, so that (fn) is equivalent to the unit vector basis of ~7oo (respectively c0) in L~ (respectively L~). These considerations in combination with Theorem 8 help to prove the following theorem.
514
S.J. Dilworth
THEOREM 9. Let cb be an N-function. The following conditions are equivalent: (a) 4) satisfies the A 0 and A ~ conditions; (b) L4,(O, o c ) - L~ (0, oo); (c) L4,(O, o0) is separable; (d) L4,(O, oo) does not contain a subspace isomorphic to f ~ ; (e) L~ (0, co) does not contain a subspace isomorphic to co. COROLLARY 3. L4, (0, oo) is reflexive if and only if both d? and 4/satisfy the A 0 and A ~
conditions. Analogous results hold for L4, (0, 1) (respectively s for N-functions satisfying the A ~ (respectively A2~ condition. The following Matuszewska-Orlicz indices provide useful information about the Banach lattice structure and interpolation properties of Orlicz spaces (for simplicity we consider only the case of L4, (0, 1)):
ot~ = sup{p>O: ~
--
sup ~()Otp < cxz} ; ~,t~>l ~b()~t)
i n f { p > O : i n f qb()~)tp > 0 } . k,t/> 1 q~()~t)
(33) (34)
Note that ot~ ~3~. The full duality between 4~ and 7t and (30) yield 1
oty ~- ~
1
1
1;
1
/~y ~- ~c%
1.
(35)
It can be shown (see [99]) that ot~ is the supremum of the set of p's for which L4, (0, 1) satisfies an upper p-estimate for disjoint functions, and that fl~ is the infimum of the set of q's for which L4,(0, 1) satisfies a lower q-estimate. From the general theory of Banach lattices described in [99] it then follows that L4, (0, 1) is super-reflexive if and only if 1 < ot~ ~< fl~ < oo. By (35) the latter condition holds if and only if both O~and !k satisfy the A ~ condition, and thus (by Corollary 3) L4, is reflexive if and only if it is super-reflexive. The indices ot~ and fl~ can also be shown to coincide with the reciprocals of the Boyd indices [13] of L4,(0, 1), which gives rise to the following useful interpolation theorem (a variation of the Orlicz interpolation theorem). THEOREM 10. Suppose that 0 < p < ~ <. ~ < q <<.o~. Then every linear operator T which is boundedsimultaneously on Lp(0, 1)and on Lq(0, 1) is also bounded on L4,(0, 1). Finally, let us discuss what happens when the convexity assumption on 4~ is relaxed. A function 4~" [0, c~) --+ [0, ~ ) is said to be a ~-function if 4~ is continuous, increasing, and satisfies ~b(0) - 0. An F-norm [. [4, is defined on L4, (0, cx~) as follows: If [4, = inf{ k > 0: M4, ( f / k ) <~k ].
Special Banach lattices and their applications
515
Lr (0, ec) is locally bounded if and only if there exists k > 0 such that ~(kt) liminf-> 1 t~0 4~(t)
and
~(kt) liminf~ > 1. ,~oc ~b(t)
(36)
If 4~ satisfies (36) then L4 becomes a quasi-Banach space with quasi-norm II" 114 defined by (27). Notes. For the theory and history of Orlicz spaces we refer the reader to the texts [104] and [110]. For interpolation theorems for Orlicz spaces and a discussion of the Boyd indices see [10,97,99], and [104]. There is an extensive literature on the Banach space geometry of Orlicz spaces: for example, see [2,3] for weak compactness, [78] and [81] for uniform convexity, [37] and [56] for smoothness, [50] and [ 107] for the Kadets-Klee property, and [36] and [86] for normal structure. For isometries of Orlicz spaces, the reader may consult [ 120] and its references. 5. Containment of g p and the Nikishin factorization theorem
A fundamental theorem of Aldous [ 1] states that every closed subspace of L 1(0, 1) contains a subspace isomorphic to g p for some 1 ~< p ~< 2. Using the results of Section 3 and Aldous' theorem one obtains the corresponding result for closed subspaces of Lp,q. THEOREM 1 1. Suppose that 1 < p < oc, 1 <~q < cx~, where p =/: q, and that X is a closed subspace of L p,q (0, ec). (a) If p >~ 2 then X contains an isomorphic copy of g.r for some r E {2, q}. (b) If p < 2 then X contains an isomorphic copy of ~-r for some r E {q } U (p, 2]. Conversely, g-r embeds isomorphically into g p,q (0, 1)for the given ranges of r. The analogue of Aldous' theorem for subspaces of Orlicz sequence spaces had been obtained earlier by Lindenstrauss and Tzafriri [95,96], with the Matuszewska-Orlicz indices oe~ and/~2 (companions to the indices ~ c and/~c defined above) playing a crucial role: ot~ = s u p { p > 0 :
sup ~b()~t) < ec ] 9 0<)~,t~
(37)
/~ -- inflP > 0:
inf 4~0 ~ t ) 0 } . 0<~., t~
(38)
THEOREM 12 ([96]). (a) The Orlicz sequence space g a contains a subspace isomorphic to gp (when p < cx~) or co (when p = co) if and only if t~~ <<.p <<.~ . (b) Every closed subspace X of ga~ contains a subspace isomorphic to co or to g.p.
The theory of stable Banach spaces initiated by Krivine and Maurey [83] provides the machinery for extending Aldous' theorem to a large class of Banach and quasi-Banach
516
S.J. Dilworth
spaces, including the Orlicz spaces Lr (0, 1) when 4~ satisfies the A ~ condition [55]. However, in order to highlight the theory offactorization of linear operators, an important area of functional analysis in which the L p,q spaces arise naturally, we shall take a different approach to one such extension of Aldous' theorem. Let A0 be the collection of all measurable functions on (0, cx~) for which df(t) < cx~ for all t > 0. It is easily checked that A0 is a separable complete metric linear space with metric d ( f , g) =
sup m(E)=l
I f - gl dx. 1 + I f -- gl
The topology induced by this metric is the topology of convergence in measure on (0, cx~). Let L0 denote the closed subspace of A0 consisting of all measurable functions supported on [0, 1]. Recall that a quasi-norm II 9II on a vector space x is a p - n o r m (0 < p ~< 1) if it satisfies ]Ix -Jr-yIIP <~ ]]xl]p -Jr-]]yllP for all x, y 6 X. Now we can prove the fundamental factorization theorem of Nikishin [ 113]. THEOREM 13. Let T be a continuous linear operator from a p-Banach space (X, ]l" II)
into Lo. There exists a non-vanishing g E Lo such that the mapping x w+ ( T x ) / g is continuous from X into Lp,~. In particular, T can be factorized through Lp,~. PROOF. We have to prove the existence of g 6 L0 such that
m ( . T x , > g.,x,. ) tl/p ~t for all x 6 X and for all t > 0. By an exhaustion argument it suffices to prove that for every e > 0 there exists C(e) > 0 and Ae C [0, 1] with m([0, 1] \ Ae) < e such that
( C(e)l,x,,) m s ~ Ae" I(Tx)(s)l > tl/p <~t.
(39)
The fact that T is a continuous operator into L0 simply means that for every t > 0 there exists c(t) > 0 such that m{ITxl > tllxll} < c(t) for all x ~ X, where c(t) --+ 0 as t --+ oo. Observe that if (ei) are independent Bernoulli random variables (defined on a probability space (I2, r , P)) taking the values 4-1 with equal probability, then by symmetry and convexity P (I ~-~in= 1 ai eil >~max lai I) ~> 1/2 for all real scalars (ai). Let (xi) be a sequence in X. Then
m maxlTxil > t
I]xill p i=1
oo
E
siTxi
> t
i=1
2E(m(E
Ilxi
IIp
i=1
oo
i=1
siTxi
>t
6iXi i--1
517
Special Banach lattices and their applications
Given a > 0, select to such that 2c(t0) < e. Then
m m a x l r x i l > to
Ilxill p
(40)
< e.
i=1
Now choose a disjoint sequence (An) of sets of positive measure and a corresponding sequence of unit vectors (Xn) which are maximal with respect to the conditions An {ITxnl >/to/m(An) l/p} for each n >/1. Then maxn>~l IT(m(An)l/Pxn)[ >/to on Un~>l An and 0<3
OO
~_~[Im(A~)'/Px~II p <~~ m ( A n ) n--1
<<.1.
n=l
Thus m ( U An) < e by (40). Let Ae -- [0, 1] \ Un~>l An. From the maximality ofthe family (An), it follows that for each x 9 X and for each t > 0, we have t011x II
m ( s 9 As" I(Tx)(s)l > tl/P
)
which proves (39) with C(e) -- to.
[:3
REMARK 1. Since the inclusion operator Lp,oo(O, 1) r Lq(O, 1) is continuous for 0 < q < p < oo, it follows from the Nikishin Factorization Theorem that every continuous operator from a p-Banach space X (0 < p ~< 1) into L0 factorizes through Lq (0, 1) for each 0 < q < p. Assuming the validity of Aldous' theorem for subspaces of L p (0, 1), when 0 < p ~< 1, we shall extend the result to subspaces of A0. But first let us observe that A0 has a large collection of Banach and quasi-Banach subspaces. PROPOSITION 3. Let r be a continuous strictly increasing C-function. Then L4~ a is isomorphic to a closed subspace of Ao. PROOF. Clearly, A0 is (0, oo) x (0, cx~) with a is a decreasing function g(t) ~ F(s)g(t). Then, 4)
isomorphic to A0, the space of all measurable functions on finite distribution function. Setting F - l ( s ) r (so that F on (0, oo)), consider the mapping T'La~ (0, e~) --+ AO given by for all ot > 0, we have dt =
f0 ( ) f0 ( F- 1
ot Ig(t)l
dt -
m s" F (s) >
dt Ig(t)l
-- (m x m)(F(s)lg(t)[ > o~) - - ( m • rn)(lTgl > ol). This calculation shows that a sequence (fn) in L~ satisfies Ifn Ir --+ 0 if and only if Tfn ---> 0 in A0. Thus, T is an isomorphic embedding.
[]
THEOREM 14. Let X be a quasi-Banach (i.e., locally bounded) subspace of Ao. Then X
contains a subspace isomorphic to co or to g,p for some p 9 (0, e~).
518
S.J. Dilworth
PROOF (Sketch). We may suppose that the topology of X is given by a p-norm II" I1 for some 0 < p ~< 1. Let T :X --+ A0 be an isomorphic embedding. First suppose that there exists M > 0 such that the composition f w-~ (T f)X[0,M] is also an isomorphic embedding. In this case X will be isomorphic (by dilation) to a subspace of L0, and thus, by Remark 1, to a subspace of Lq (0, 1) for all 0 < q < p, whence X will contain a subspace isomorphic to ~r for some r ) p. On the other hand, if the mapping f w-~ (Tf)x[O,M] fails to be an isomorphic embedding for every M > 0, then one can find an increasing sequence (Mn) of positive numbers, and corresponding unit vectors Xn ~ X, such that (after setting f~ = (Txn)X[Mn_~,Mn]) d(Txn, fn) --+ 0 rapidly enough to ensure that (Xn) and (fn) are equivalent basic sequences (see [74]). Moreover, by Helly's selection theorem we may assume that there exists a nonzero decreasing function F(t) on (0, 1] such that fn* --+ F pointwise. Passing to a subsequence to ensure that d (fn* X[0,11, F) --+ 0 rapidly enough, it will follow that (fn) is equivalent to a disjoint sequence (Fn), where each Fn has the same distribution as F. If F is bounded then (Fn) is equivalent to the unit vector basis of co. If F is unbounded, set q~(t) =dF(1/t), for t > 0, and observe that
m
(5)
anFn ) t =
n=l
5
n=l
dF(t/lan)--
5
4)(lanl/t)
n=l
for all t > 0. It will follow that the Orlicz sequence space ( ~ , II 9I10~)is a p-convex quasiBanach space isomorphic to [Xn], and then from the non-locally convex version of Theorem 12 (see [75]) that [Xn] contains a subspace isomorphic to ~r for some r ~> p. D
Notes.
For a proof of Theorem 12 see the texts [97] and [98] (as well as [96]). The set of p's such that g p embeds isomorphically into the Orlicz function space Lq~(0, 1) is determined in [63]. The corresponding problem for the function space L~(0, ec) is solved in [112] and [63]. A characterization of the Orlicz sequence spaces g0 which embed isomorphically into g~ is given in [95], where the problem of complemented g~p'S is also considered. An example is given there of a reflexive Orlicz sequence space which does not contain any gp as a complemented subspace (see also [97] and [98]). See [61-63] (and [60] for examples of Orlicz function spaces L~(0, 1) without any complemented gp) for results on complemented s in Orlicz function spaces. Kalton [75] extended Theorem 12 to the non-locally convex and the non-locally bounded cases and showed that the theory for complemented subspaces is significantly different in the non-locally convex setting. The proof given here of the Nikishin Factorization Theorem is modelled on the proofs in in [76] and [136]. The assumption that X be a p-Banach space can be relaxed (see [76]) to Rademacher type p (0 < p ~< 2). The Lorentz spaces Lp, l(#) and Lp,cc(#) also arise in the factorization theory of absolutely summing operators: Pisier [ 119] (see also the text [40]) proved that a (p, 1)-summing operator from a C(K) space into a Banach space admits a factorization (analogous to the Pietsch Factorization Theorem) through an Lp,1 (/z) space. See [44] for a recent application of this result, where it is used to show that cotype characterizes the H61der-continuity properties of the indefinite Pettis integral.
Special Banach lattices and their applications
519
See the text [57] for the theory of stable Banach spaces and for the proof of the following refinement of Aldous' theorem: a closed subspace X of L1 contains almost isometric copies of gp, where p is the infimum of the type interval for X. Garling [55] proved that L 4,(0, 1) is a stable Banach space if 4~ satisfies A ~ . His arguments readily extend to cover the quasi-Banach case of closed subspaces of L p, for 0 < p < 1, as was observed in [49]. Raynaud [121] proved that (Lp,q, I[" lip,q) (1 ~< q ~< p) and (Lp,q, I[" II(p,q)) ( p > 1) are stable. It follows that every closed subspace of these spaces contains almost isometric copies of ~r for some r. See [55] for a representation theorem for the space of 'types' on L~, and [9] for types on Orlicz and Lorentz sequence spaces. See [8] for results on the stability of quotient spaces and of interpolation spaces. Proposition 3 and Theorem 14 are adapted from [49].
6. Some probabilistic applications One area of analysis in which Orlicz and Lorentz norms arise frequently is probability theory. Indeed, a useful way of studying the integrability of a random series is to determine its closed linear span in an appropriate function space. In this section we present two important results of this type. Khintchine's inequality states that a convergent Bernoulli series has finite moments of all orders, and that the closed linear span in Lp (0 < p < oe) of a Bernoulli sequence is isomorphic to g2. The Orlicz spaces L~q, where ~q(t) - - e ta -- 1, may be used to study the exponential integrability properties of a Bernoulli series. The following result of Rodin and Semenov [ 130] identifies the closed linear span of a Bernoulli sequence in L~q (q > 2) with the Lorentz sequence space g~p,~, where 1/p + 1/q -- 1. THEOREM 15. Let (en) be a sequence of independent Bernoulli random variables defined on a probability space (S-2, r , P). Then, for each q > 2, there exist constants C1, C2 such that
1 - - II(a,)llp,~ C1
C21l(an)llp,~
~-~ an En
(41)
~q
n=l
for all scalars (an), where 1 / p + 1/q -- 1. PROOF. First we prove a fundamental estimate, known as Hoeffding's inequality [67], for the distribution function of a convergent Bernoulli series: L
P
anen > t
~<2exp
n=l
2 ~n~__l a 2
.
(42)
To prove (42), we may assume by homogeneity that Y~n= ~ l an2 -- 1. Observe that by independence of the en'S, we have Eexp t
anen n--1
-
Eexptanen<<, n----I
exp ~ n=l
-exp
,
520
S.J. Dilworth
and so by Markov's inequality exp(t 2)P
<~ exp(t 2/2),
anen > t n=l
from which (and from symmetry) (42) follows. Suppose now that II(an)I] p,~ = 1. An interpolation argument will be given to show that
P
(s
anen > t
<~ 2exp
,
(43)
n=l
for all t > 0, where Cq depends only on q. First, we may assume that (lan I) is a decreasing sequence. Secondly, it suffices to prove (43) for all t > 2q. Let m be the largest integer such that t > 2qm 1/q + 1. Then
•
m
(3o
m
(3O
an 6n
Z
n--1
n=l
n=m+l
n=l
an 8n
n=m+l
o(3
<~ (qm 1/q + l)-t-
Z
anen ,
n=m+l
and so by (42) (since we are assuming that t > 2qm 1/q + 1) P
anen > t
<~ P
n=l
anen > qml/q n=m+l
~< 2exp
2}~n~
(44)
"
But o<3
oo
2 an<~ Z
Z n=m+l
P___~ml-2/p = ~ qm n-2/P<~2_ p q--2
1-2/p
n--m+l
and it follows from (44) that
n=l
which yields (43) since m ~ (1/2q)qtq. Integration by parts and (43) give
E l[/q Ol
anen
--
=
P
oe ~_~ anen > t
d(exp(t q) - 1)
n=l
~<
f0 (,q) 2 exp Cqotq
d(exp(tq ) - 1) <~ 1
Special Banach lattices and their applications
521
for all ot < Otq. This gives the right-hand inequality of (41). For the left-hand inequality, observe that since P (Y~'~n=l m en -- m) -- 2-m there exists Cq > 0 such that
n=l
Oq
n=l
p,oc
where the second inequality follows by direct calculation. Since II(a~)ll- II ~ - - 1 anellOq defines a symmetric sequence space norm, it follows from the sequence space version of (16) that
•
CqII(an)ll p, ~ ,
an 6n
~q
nzl
which completes the proof of (41). REMARK 2. Note that it follows from (42) that for q ~< 2 the linear span of a Bernoulli sequence in L~q is isomorphic to ~2. Our second application is a result of Johnson and Schechtman [72] which extends (to the range 0 < p < 2) Rosenthal's moment inequality for sums of independent mean zero random variables in Lp (p > 2) [131]. First let us introduce the Orlicz spaces Yp =- L49p (0, cx3) (0 < p < ~ ) , where
dpp(t)
- / t2
I tp
for 0 <~ t ~< 1, for 1 < t < cx~.
Observe that Yp = Lp(O, cx~) -+- L 2 ( 0 , ~x)) ( r e s p e c t i v e l y Lp(O, ~ ) 0 L 2 ( 0 , ~ ) ) for 0 < p < 2 (respectively 2 < p < ~ ) . Note also the duality Yp = Yq for 1 / p + 1/q = 1. The following theorem makes use of the disjoint sum notation introduced in Section 3. THEOREM 16 [72]. Let 0 < p < c~. Then for every sequence of independent random variables (fn)n~=l defined on a probability space (1-2, S , P), we have
G fn n=l
p
n=l
,
(45)
~bp
with constants of equivalence depending only on p. PROOF (Sketch). For the range p > 2 there is a short elementary proof (essentially the proof given by Rosenthal [131 ]), which we present. Making the substitutions gn = f 2 and q = p / 2 (so that q > 1), (45) will follow from
~gi i=1
~< max 2 q Z i--1
Ilgi II1,2
I[gi IIq i=1
9
(46)
522
S.J. Dilworth
(The reverse inequality (with 2 replaced by 1) does not require independence but just follows from the positivity of the gi's and an application of H61der's inequality.) To prove (46) observe that ~< 2 q-1 ( f i gq -+- f i ( j ~ / i--1 i=1
( f i ) q gi i=1
gj )q-1 gi ) 9
So by independence of the gi's, we have
~< 2 q-1 ( f i Ilgi IIq + f i ( f i e gi i=1 q i=1 i--1 ~ 2q-1
("
~ I[gi IIq + i=l
Ilgilll j~/ gj -' 9 " qq-1 Ilgi Ill
i=1
t)
1 i:, I11_,t
(by H61der's inequality and the positivity of the gi's) ~< 2 q max
Ilgi IIq,
i=1
Ilgi II1
figi i=1
i=1
q-l), q
which yields (46). The proof in the range 0 < p < 2 requires more elaborate arguments for which the reader should consult [72]. [~ Using the fact (due to Marcinkiewicz and Zygmund [105]) that the norm in L p of a sum of independent mean zero random variables is equivalent to the norm of the square function, we can reformulate the previous theorem. COROLLARY 4. Let (Xn) be a sequence of independent mean zero random variables. Then, f o r 1 <<.p < oo, we have "
E
Xn
"~
n=l
~Xn
n=l
9
4~p
As a last pretty application of Theorem 16, let us see how the closed linear span in L p of a sequence of independent identically distributed random variables may be identified with a certain Orlicz sequence space. COROLLARY 5. Let (Xn) be a sequence of independent identically distributed mean zero random variables with common distribution function F(t). Then, f o r 1 <~ p < oo, we have cx~ E ~
p anXn
p ~ ]](an)J]C~p,F ,
n=l where dPp,F(t) -- f R dpp(tS) dF(s).
Special Banach latticesand theirapplications
523
PROOF.
p) 1/p
oo
E
~
n=l
(~ an Xn
anXn n=l
~bp
--inf{k>O: ~Ec/)p(anXn) k <,1 --inflk>O: ~RfdPP(~-)dF(s)<<]ln'=l = [(an) I4)p,F. Notes.
[]
The proof of Theorem 15 is adapted from the proofs given in [106] and also [87]. See the text [ 14] for a characterization of the i.i.d, sequences for which the estimates in (41) hold. See [25] and [15] for some inequalities for i.i.d, sequences in the Lp,q spaces: it is proved in [25] that for p > 2 the closed linear span of a mean zero i.i.d, sequence is always a complemented Hilbertian subspace, and that in L2,q a mean zero i.i.d, sequence satisfies an upper g2,q-estimate if and only if 0 < q ~< 2. See [108] for probabilistic inequalities involving weak L p norms in the local theory of Banach spaces. Theorem 15 was extended to the case of vector-valued Bernoulli series, with coefficients from an arbitrary Banach space, by Montgomery-Smith and the author [48]. See [48] also for a fundamental two-sided estimate (improving (42)) for the distribution of a vectorvalued Bernoulli series, involving the K-functional for a certain interpolation couple. Theorem 16 can be formulated much more generally for sums of independent random variables in rearrangement-invariant quasi-Banach function spaces [72] (cf. [23] for the special case of Lp,q). See [72] for an important application of Theorem 16 to the uniform approximation property in L p spaces. Another way to obtain Theorem 16 in the range 0 < p < 2 is by a duality argument, using a version of the Burkholder-Davis-Gundy martingale inequalities [18]. We refer the reader to [43] for a full exposition of this approach. Rosenthal's inequality for Lp (p > 2) (first proved to exhibit new complemented subspaces of L p) in some sense characterizes L p. See [14] for a precise formulation of this fact and for further results on the linear span of independent random variables in rearrangement-invariant function spaces. See [71] and the text [114] for martingale inequalities in rearrangement-invariant spaces. The Orlicz space Yp is isomorphic to L p in the range 1 < p < cx~ [70] (and so L p has two representations as a rearrangement-invariant function space on (0, ec)). The 'other' scale of sums and intersections of Lp spaces, Mp = Lp(O, cx~)71L2(0, oo) (respectively Lp(O, oo) -k- L2(0, ~ ) ) for 0 < p < 2 (respectively 2 < p < ec), is closely related to (though isomorphically distinct from) the L p scale. Several Banach space properties of these spaces and of the L2(s spaces are obtained in [41,42], e.g., that there is no isomorphic embedding of L2(~p) into g2(gp) for p r 2. For the structure (e.g., containment of gr(gs) subspaces) of the 'mixed norm' spaces Lp(Lq) see the papers of Raynaud [ 122,123,127].
S.J. Dilworth
524
7. Embedding
Lw,q into Lq
In his seminal work [7] Banach discussed the question of the 'linear dimension' of the classical Banach spaces, that is, the question of the existence (or non-existence) of an isometric or an isomorphic embedding from one space into another. A famous question left open in [7] for many decades concerned the existence of an isomorphic embedding from ~q into Lp for 1 ~< p < q < 2. Surprisingly, the solution to this problem turned out to be probabilistic in nature: using stable processes Bretagnolle, Dacunha-Castelle and Krivine [16] proved that it is even possible to embed Lq isometrically into L p in the range 0 < p < q ~< 2 (cf. [82] for a self-contained proof). In this section we shall consider the question of embedding Lorentz spaces into Lq. It is appropriate to study this question not just for the L p,q spaces but for the wider class of Lw,q spaces, which will now be introduced. Let I denote the interval (0, 1) or the interval (0, cx~), and let w(t) be a positive decreasing weight function defined on I for which fd w(t) d t = 1 and f o w(t) dt = cx~ (the latter condition only for I = (0, cxz)). For 0 < q < c~, the Lorentz function space Lw,q(I) is the space of equivalence classes of real-valued measurable functions f on I for which the following norm (or q-norm if q < 1)
Ilf llw,q -
(f,
f * (t)q w(t) dt
)"q q
1
is finite. For 0 < q ~< p < c~, the weight w(t) - ( q / p ) t F - corresponds to the space Lp,q with equality of norms. For q ~> 1, the condition that w should be decreasing guarantees (and, in fact, is equivalent to) the triangle inequality in Lw,q [ 101 ]. Given the existence of the stable embeddings of L p into Lq, a natural question to consider is the possibility of an isometric embedding of Lw,q (I) into Lq. It turns out that such embeddings exist for only a very restrictive class of weights [46]. THEOREM 17. Let 0 < q < oo and let I be either (0, 1) or (0, c~). There exists an isometric embedding of Lw,q (I) into Lq if and only if one of the following conditions is satisfied: (a) w(t) -- 1, in which case Lw,q (I) -- Lq (I) with equality of norms; (b) I = (0, 1), 0 < q ~< 1, and w(t) is a decreasing linear weight; that is, there exists ot E [0, 2] such that
w(t) =_ w~(t) = 1 + - -t~t. 2 The case q > 1 is easy: the space Lq is smooth, but Lw,q is not, unless w(t) ---- 1. For 0 < q < 1, the proof of Theorem 17 is quite technical as it uses the theory of Fourier transforms of distributions. For q = 1, however, a short geometrical argument can be given for the main step of the proof, which concerns finite-dimensional Lorentz sequence spaces. For a l >~ ... >~an >~0 (not all zero) and q > 0, the expression
II
.....
II -- (a' Cx )" +""
+
'/"
Special Banach lattices and their applications
525
is the norm (or q-norm if q < 1) of an n-dimensional Lorentz space denoted ~n,q. For an infinite decreasing sequence w -- (an) of positive weights, for which ~-~n~=l an -- co, the Lorentz sequence space g w,q is defined similarly. PROPOSITION 4. The s p a c e g.n,1 (n ~ 2) is i s o m e t r i c to a s u b s p a c e o f L1 i f a n d only i f al - a2 -- a2 - a3 . . . . .
an-1 - an.
(47)
PROOF. First we prove by induction that ~n,1 is isometric to a subspace of L1 only if the weights ak satisfy (47). Let E be a Lorentz space of dimension n + 1 with weights a 1. . . . . an+ 1. If E is isometric to a subspace of L 1 then (47) holds by hypothesis, and so it suffices to prove that an-1 - an = an - an+l to complete the induction. In particular, we may assume that a n - l , an and an+l are not all equal. Let B* denote the unit ball of E*. Since the unit ball of E is a polytope, and since E is isometric to a subspace of L1, it follows (see, e.g., [12]) that B* is a z o n o t o p e (that is, a Minkowski sum of line segments). By [12, Theorem 3.3], all of the two-dimensional faces of B* are centrally symmetric. It is easily seen that the extreme points of B* are all the sign-changed permutations of the vector a = (al, a2 . . . . . an+l). In particular, one of the two-dimensional faces of B* has as its vertices all of the vectors obtained by permuting the last three coordinates of a. If an-t = an or if an = an+l, then this face is triangular, which contradicts the central symmetry requirement. If an-1 > an > an+l, then the face is h e x a g o n a l , and the symmetry condition forces an-1 - an = an - an+l as required. Thus (47) is a necessary condition. To show that (47) is also sufficient, one can check that if (47) is satisfied then B* has four kinds of two-dimensional faces: two classes of quadrilateral faces, one class of octagonal faces, and one class of hexagonal faces like the face described above. The first three kinds of faces are automatically centrally symmetric without any condition on the weights, while (47) guarantees that the hexagonal faces are also centrally symmetric. So, by [12, Theorem 3.3] once again, if (47) is satisfied, then B* is a zonotope, and hence E is isometric to a subspace of L1. D The question of embedding Lw,q isomorphically into Lq is answered by the following theorem of Schtitt [132]. THEOREM 18. L e t 1 <<,q < 2. T h e n the f o l l o w i n g are equivalent: (a) Lw,q (0, 1) is i s o m o r p h i c to a s u b s p a c e o f Lq ; (b) Lw,q (0, 1) is a 2 - c o n c a v e B a n a c h lattice; (c) There exists C > 0 such that f o r all x E [0, 1], w e h a v e
fo x w ( t ) t - p / 2 dt <~ Cx -p/2 fo x w ( t ) dt. An interesting corollary of Theorem 18 is the fact that Lp,q embeds into Lq in the range 0 < q ~< p < 2. Let us use Rosenthal's inequality (Theorem 16) to sketch a short proof of this important special case of Theorem 18, following [26].
S.J. Dilworth
526
By Theorem 2, Lp,q is the interpolation space [Lr, Ls]O,q (with equivalence of norms), where 1/ p -- (1 - O) / r + 0/s. Thus,
Ilfllp,q ~
(~ ?1~--OQ
)l/q
2 -nOq g ( 2 n, f ; gr, gs) q
9
(48)
Let us denote by Lr + t 9Ls the Orlicz space Lr -Jr-Ls equipped with the equivalent norm K(t, f ; Lr, Ls). Then it follows from (48) that Lp,q is isomorphic to a closed subspace of (Zn---cx~CX~O ( t r -Jr-2 n 9Ls))q. For )~ > 0, the dilation property of the norm in L p (namely, IIf (Zx)ll p = z - l / p IIf (x)ll p), yields
g ( t , f ; Lr, Ls) = inf{~/rllg(~x)llr + t~l/Sllh(~x)lls" f - g + h}. Setting ~. = t rs/(s-r) (so that t - ~(s-r)/rs)), we have
g ( t , f ; Lr, Ls) -- ~/rinf{llg(~x)llr + IIh(~x)ll~" f - g + h }
=
Lr, Ls).
Thus the mapping f (x) w+ ts/(s-r) f (trs/(s-r)x) defines a linear isometry from Lr + t . Ls onto Lr + Ls. Thus, (Y~n~=_~ ~)(Lr + 2n" Ls))q is isometrically isomorphic to ~.q(Lr + Ls). In particular, if 0 < q < p < 2, then, setting r -- q and s -- 2, we see that Lp,q embeds isomorphically into s + L2). But by (45) of Theorem 16, the mapping oo
n=l
where fn -- f X(n- 1,n) and (Sn) is a sequence of independent random variables each having the uniform distribution on [0, 1], yields an isomorphic embedding of Yq = Lq + L2 into Lq(s Since s embeds into Lq, it follows that Lq(/~2) embeds into Lq(Lq), which in turn is isometrically isomorphic to Lq. For the other ranges of p and q (p r q) it turns out that it is not possible to embed L p,q into Lr (r < cx~). In particular, since L2,q (q ~ 2) does not have cotype 2 (or type 2) there is no embedding of L2,q into Lq. It can also be shown that there is no embedding of Lp,q into L1 when p < q (see [132] or [26]).
Notes. The Lw,q spaces were introduced by Lorentz [ 101] with different notation as the spaces A(4~, q), where ok(t) - fo w(s)ds. There is an extensive literature on the Banach space geometry of these spaces. The following property of the weight is especially important: w is said to be regularif there exists k > 1 such that 4~(2x) > k ~ ( x ) for all x > 0. For q > 1, Halperin [58] proved that Lw,q (I) is uniformly convex if and only if w is regular. The sequence space version of this result together with estimates for the modulus of convexity were obtained by Altshuler [1 ]. On the other hand, the uniform Kadets-Klee property in Lw, 1(0, cx~) was shown to be equivalent to 'uniform concavity' of 4~ [45,47], which is a stronger condition than regularity of w. It is proved in [27] (cf. [28]) that Lw,1 (0, e~) has the Fixed Point Property for non-expansive mappings on weakly compact convex sets when w is strictly decreasing. See [32] for the isometries of Lw,q (0, cx~).
Special Banach lattices and their applications
527
Reisner [ 129] proved that Lw,q(I) is p-concave for some p < cx~ if and only if w is regular. He also gave necessary and sufficient conditions on w for Lw,q to be p-concave and proved that the interval of p's for which Lw,q is p-concave is an open interval (cf. [ 115] for these results and [ 126] for more general results in which the condition that w be decreasing is relaxed). The Boyd indices for Lw,q are computed in [115] and [109]. See [126] for the representation of Lw,q as an interpolation space (Lr L?)o,q (generalizing Theorem 2). It is proved in [116] that a closed subspace of Lw,q (0, 1) either contains almost isometric copies of gq or or is strongly-embedded in L j(0, 1). Garling [54] identified the dual space and proved reflexivity of the sequence space gw,p (p > 1). Allen [4] obtained a simple description of the dual space for regular weight sequences. See [34] for proofs of the Kadets-Klee property and local uniform convexity of gw,p. The symmetric basic sequences in gw,p are characterized in [6]. Necessary and sufficient conditions are given for g w,p to have exactly two nonequivalent symmetric bases and an example is given of a subspace of a Lorentz sequence space with a symmetric basis which is not itself isomorphic to a Lorentz sequence space. Symmetric basic sequences in the dual of gw,p are considered in [33]. Bretagnolle and Dacunha-Castelle [ 17] proved that Lr is isomorphic to a subspace of L 1 if and only if 4) is equivalent to a 2-concave convex 4)-function (recall that 7t is 2-concave if g r ( ~ ) is concave), which is the Orlicz space analogue of Theorem 18. See [133] for an explicit embedding of a given 2-concave finite-dimensional Orlicz sequence space into L 1. The proof utilizes some combinatorial and probabilistic inequalities [84,85] which have also been used to compute projection constants for finite-dimensional Orlicz and Lorentz spaces [84]. Hermindez and Ruiz [65] proved that, for 0 < p < q < cx~, every Orlicz space Lr (0, cx)) whose Boyd index reciprocals lie strictly between p and q is lattice-isomorphic to a sublattice of Lp -+-Lq (see also [52]). Raynaud [123] proved that every p-convex and q-concave Orlicz space L~(0, oc) is lattice-isomorphic to a sublattice of Lp(Lq). See [128] for results on embedding symmetric spaces into L1 (gq), and [64] and [124,125] for further embedding results. A special case of a theorem of Kalton [76] asserts that the Banach space gp(gq) embeds into L0 if and only if 1 ~< p ~< q ~< 2. Briefly, let us mention a way in which Orlicz and Lorentz spaces have been integrated into a single class of spaces. Let 4) be a convex 4)-function and let w be a decreasing weight function. The Orlicz-Lorentz space A~,w(I) is the Banach space of measurable functions f on I for which
IIf II = inf, )~ > O:
f0
$(f*(t)/~.)w(t) dt
J
< e~.
See [81] for uniform convexity, [80] for extreme point structure (see also [29] for the uniqueness of the Choquet integral representation in Lw,1 (0, oc)), [79] for containment of co and g ~ , and [68] for local uniform convexity and the Kadets-Klee property (see also [ 134] for Lw, 1) of Orlicz-Lorentz spaces. See [ 109] and [ 123] for concavity properties and Boyd indices for Orlicz-Lorentz spaces.
References [1] D.J. Aldous, Subspaces of L 1 via random measures, Trans. Amer. Math. Soc. 267 (1981), 445-463.
528
S.J. Dilworth
[2] J. Alexopoulos, De La Vallde Poussin's theorem and weakly compact sets in Orlicz spaces, Quaestiones Math. 17 (1994), 231-248. [3] J. Alexopoulos, On subspaces ofnon-reflexive Orlicz spaces, Quaestiones Math. 21 (1998), 161-175. [4] G.D. Allen, Duals of Lorentz spaces, Pacific J. Math. 77 (1978), 287-291. [5] Z. Altshuler, Uniform convexity in Lorentz sequence spaces, Israel J. Math. 20 (1975), 260-274. [6] Z. Altshuler, P.G. Casazza and B.L. Lin, On symmetric basic sequences in Lorentz sequence spaces, Israel J. Math. 15 (1973), 144-155. [7] S. Banach, Th~orie des Operations Lin~aires, Warsaw (1932). [8] J. Bastero and Y. Raynaud, Quotients and interpolation spaces of stable Banach spaces, Studia Math. 93 (1989), 223-239. [9] J. Bastero and Y. Raynaud, Representing types in Orlicz and Lorentz sequence spaces, Math. Proc. Cambridge Philos. Soc. 107 (1990), 525-538. [10] C. Bennett and R. Sharpley, Interpolation of Operators, Academic Press, New York (1988). [11] J. Bergh and J. L6fstr6m, Interpolation Spaces, Springer-Verlag, New York (1976). [12] E.D. Bolker, A class of convex bodies, Trans. Amer. Math. Soc. 145 (1969), 323-345. [13] D.W. Boyd, Indices for the Orlicz spaces, Pacific J. Math. 38 (1971), 315-323. [14] M.Sh. Braverman, Independent Random Variables in Rearrangement Invariant Spaces, Cambridge University Press, Cambridge (1994). [15] M. Braverman, Independent random variables in Lorentz spaces, Bull. London Math. Soc. 28 (1996), 79-87. [16] J. Bretagnolle, D. Dacunha-Castelle and J.L. Krivine, Lois stables et espaces L p, Ann. Inst. H. Poincarr, Probab. Statist. 2 (1966), 231-259. [17] J. Bretagnolle and D. Dacunha-Castelle, Applications de l'~tude de certaines formes lindaires al~atoires au plongement d'espaces de Banach dans les espaces LP, Ann. Sci. l~cole Norm. Sup. 2 (1969), 437-480. [18] D.L. Burkholder, B.J. Davis and R.E Gundy, Integral inequalities for convex functions of operators on martingales, Proc. Sixth Berkeley Symp. Math. Statist. Probab., Vol. 2, Univ. of California Press, Berkeley, CA (1972), 223-240. [ 19] N.L. Carothers, Rearrangement invariant subspaces of Lorentz function spaces, Israel J. Math. 40 (1981), 217-228. [20] N.L. Carothers, Rearrangement invariant subspaces of Lorentz function spaces, II, Rocky Mountain J. Math. 17 (1987), 607-616. [21] N.L. Carothers, Lorentz Function Spaces, Seminar Notes, Bowling Green State University (1993) (unpublished). [22] N.L. Carothers and S.J. Dilworth, Geometry of Lorentz spaces via interpolation, University of Texas Functional Analysis Seminar Longhorn Notes 1985-86, Univ. Texas, Austin (1987), 107-134. [23] N.L. Carothers and S.J. Dilworth, Inequalities for sums of independent random variables, Proc. Amer. Math. Soc. 104 (1988), 221-226. [24] N.L. Carothers and S.J. Dilworth, Subspaces of Lp,q, Proc. Amer. Math. Soc. 104 (1988), 537-545. [25] N.L. Carothers and S.J. Dilworth, Equidistributed random variables in Lp,q, J. Funct. Anal. 84 (1989), 146-159. [26] N.L. Carothers and S.J. Dilworth, Some Banach space embeddings of classical function spaces, Bull. Austral. Math. Soc. 43 (1991), 73-77. [27] N.L. Carothers, S.J. Dilworth and C.J. Lennard, On a localization of the UKK property and the fixed point property in L w, 1, Interaction Between Functional Analysis, Harmonic Analysis, and Probability (Columbia, MO, 1994), Lecture Notes in Pure and Appl. Math. 175, Dekker, New York (1996), 111-124. [28] N.L. Carothers, S.J. Dilworth, C.J. Lennard and D.A. Trautman, A fixed-point property for the Lorentz space Lp,1 (#), Indiana Univ. Math. J. 40 (1991), 345-352. [29] N.L. Carothers, S.J. Dilworth and D.A. Trautman, On the geometry of the unit spheres of the Lorentz spaces Lw,1, Glasgow Math. J. 34 (1992), 21-25. H ot
//
[30] N.L. Carothers and P. Flinn, Embedding g.p in g.p,q, Proc. Amer. Math. Soc. 88 (1983), 523-526. [31] N.L. Carothers and B. Turett, Isometries on Lp,1, Trans. Amer. Math. Soc. 297 (1986), 85-103. [32] N.L. Carothers, R.G. Haydon and P.K. Lin, On the isometries of the Lorentz function spaces, Israel J. Math. 84 (1993), 265-287.
Special Banach lattices and their applications
529
[33] EG. Casazza and B.L. Lin, On symmetric basic sequences in Lorentz sequence spaces, H, Israel J. Math. 17 (1974), 191-218. [34] EG. Casazza and B.L. Lin, Some geometric properties of Lorentz sequence spaces, Rocky Mountain J. Math. 7 (1977), 683-698. [35] S.T. Chen, Geometry of Orlicz spaces, Dissertationes Math. 356 (1996). [36] S. Chen and Y. Duan, Normal structure and weakly normal structure of Orlicz spaces, Comment. Math. Univ. Carolinae 32 (1991), 219-225. [37] S.T. Chen, H. Hudzik and A. Kamifiska, Support functionals and smooth points in Orlicz function spaces equipped with the Orlicz norm, Math. Japon. 39 (1994), 271-279. [38] J. Creekmore, Type and cotype in Lorentz Lp,q spaces, Indag. Math. 43 (1981), 145-152. [39] M. Cwikel and Y. Sagher, L(p, oc)*, Indiana Univ. Math. J. 21 (1972), 781-786. [40] J. Diestel, H. Jarchow and A. Tonge, Absolutely Summing Operators, Cambridge University Press, Cambridge (1995). [41 ] S.J. Dilworth, Intersection of Lebesgue spaces L ! and L 2, Proc. Amer. Math. Soc. 103 (1988), 1185-1188. [42] S.J. Dilworth, A scale of linear spaces related to the Lp scale, Illinois J. Math. 140 (1990), 140-158. [43] S.J. Dilworth, Some probabilistic inequalities with applications to functional analysis, Banach Spaces (Merano, Venezuela, 1992), Contemp. Math. 144, W.B. Johnson and B.-L. Lin, eds, Amer. Math. Soc., Providence, RI (1993), 53-67. [44] S.J. Dilworth and M. Girardi, Nowhere weak differentiability of the Pettis integral, Quaestiones Math. 18 (1995), 365-380. [45] S.J. Dilworth and Y.-P. Hsu, The uniform Kadets-Klee property for the Lorentz spaces Lw, 1, J. Austral. Math. Soc. (Series A) 60 (1996), 7-17. [46] S.J. Dilworth and A. Koldobsky, The Fourier transform of order statistics with applications to Lorentz spaces, Israel J. Math. 92 (1995), 411-426. [47] S.J. Dilworth and C.J. Lennard, Uniformly Kadets-Klee Lorentz spaces Lw, 1 and uniformly concave functions, Canad. Math. Bull. 39 (1996), 266-274. [48] S.J. Dilworth and S.J. Montgomery-Smith, The distribution of vector-valued Rademacher series, Ann. Probab. 21 (1993), 2046-2052. [49] S.J. Dilworth and D.A. Trautman, On two function spaces which are similar to L O, Proc. Amer. Math. Soc. 108 (1990), 451-456. [50] D. van Dulst and V. de Valk, (KK)-properties, normal structure andfixed points ofnonexpansive mappings in Orlicz sequence spaces, Canad. J. Math. 38 (1986), 728-750. [51] T. Figiel, W.B. Johnson and L. Tzafriri, On Banach lattices and spaces having local unconditional structure with applications to Lorentz function spaces, J. Approx. Theory 13 (1975), 297-312. [52] A. Garcfa del Amo and EL. Hern~ndez, Embeddings offunction spaces into LP + Lq, Banach Spaces (Merano, Venezuela, 1992), Contemp. Math. 144, W.B. Johnson and B.-L. Lin, eds, Amer. Math. Soc., Providence, RI (1993), 107-113. [53] J. Garcia-Cuerva, J.L. Torrea and K.S. Kazarian, On the Fourier type of Banach lattices, Interaction Between Functional Analysis, Harmonic Analysis, and Probability (Columbia, MO, 1994), Lecture Notes in Pure and Appl. Math. 175, Dekker, New York (1996), 169-179. [54] D.J.H. Garling, A class of reflexive symmetric BK-spaces, Canad. J. Math. 21 (1969), 602-608. [55] D.J.H. Garling, Stable Banach spaces, random measures and Orlicz function spaces, Probability Measures on Groups (Proceedings Oberwolfach), Lecture Notes in Math. 928, Springer-Verlag, New York (1981), 121-175. [56] R. Grz~lewicz and H. Hudzik, Smooth points of Orlicz spaces equipped with Luxemburg norm, Math. Nachr. 155 (1992), 31-45. [57] S. Guerre-Delabribre, Classical Sequences in Banach Spaces, Marcel Dekker, New York (1992). [58] I. Halperin, Uniform convexity in function spaces, Duke Math. J. 21 (1954), 195-204. [59] G.H. Hardy, Note on a theorem of Hilbert, Math. Z. 6 (1920), 314-317. [60] F.L. Hermindez and V. Peirats, Orlicz function spaces without complemented copies of s Israel J. Math. 56 (1986), 355-360. [61] EL. Hern~indez and B. Rodriguez-Salinas, On g.P-complemented copies in Orlicz spaces, Israel J. Math. 62 (1988), 37-55.
530
S.J. Dilworth
[62] EL. Hermindez and B. Rodriguez-Salinas, On g.P-complemented copies in Orlicz spaces, H, Israel J. Math. 68 (1989), 27-55. [63] EL. Hern~indez and B. Rodriguez-Salinas, Remarks on the Orliczfunction spaces L~ (0, c~), Math. Nachr. 156 (1992), 225-232. [64] EL. Hern~indez and B. Rodriguez-Salinas, Lattice-embedding L p into Orlicz spaces, Israel J. Math. 90 (1995), 167-188. [65] EL. Hern~indez and C. Ruiz, Universal classes of Orlicz function spaces, Pacific J. Math. 155 (1992), 87-98. [66] EL. Hern~indez and C. Ruiz, On embeddings offunction spaces into LP + Lq, Banach Spaces (Merano, Venezuela, 1992), Contemp. Math. 144, W.B. Johnson and B.L. Lin, eds, Amer. Math. Soc., Providence, RI (1993), 53-67. [67] W. Hoeffding, Probability inequalities for sums of bounded random variables, J. Amer. Statist. Assoc. 58 (1963), 13-30. [68] H. Hudzik, A. Kamifiska and M. Mastyto, On geometric properties of Orlicz-Lorentz spaces, Canad. Math. Bull. 40 (1997), 316-329. [69] R.A. Hunt, On L(p,q)spaces, Enseigne. Math. 12 (1966), 249-274. [70] W.B. Johnson, B. Maurey, G. Schechtman and L. Tzafriri, Symmetric structures in Banach spaces, Mem. Amer. Math. Soc. 217 (1979). [71] W.B. Johnson and G. Schechtman, Martingale inequalities in rearrangement invariant function spaces, Israel J. Math. 64 (1988), 267-275. [72] W.B. Johnson and G. Schechtman, Sums of independent random variables in rearrangement invariant function spaces, Ann. Probab. 17 (1989), 789-808. [73] M.I. Kadets and A. Petczyfiski, Bases, lacunary sequences and complemented subspaces in the spaces Lp, Studia Math. 21 (1962), 161-176. [74] N.J. Kalton, Basic sequences in F-spaces and their applications, Proc. Edinburgh Math. Soc. 19 (1974), 151-167. [75] N.J. Kalton, Orlicz sequence spaces without local convexity, Math. Proc. Cambridge Philos. Soc. 81 (1977), 253-277. [76] N.J. Kalton, Linear operators on Lp forO < p < 1, Trans. Amer. Math. Soc. 259 (1980), 319-355. [77] N.J. Kalton, Banach spaces embedding into L O, Israel J. Math. 52 (1985), 305-319. [78] A. Kamifiska, On uniformly convex Orlicz spaces, Indag. Math. 44 (1982), 27-36. [79] A. Kamifiska, Some remarks on Orlicz-Lorentz spaces, Math. Nachr. 147 (1990), 29-38. [80] A. Kamifiska, Extreme points in Orlicz-Lorentz spaces, Arch. Math. 55 (1990), 173-180. [81] A. Kamifiska, Uniform convexity of generalized Orlicz spaces, Arch. Math. 56 (1991), 181-188. [82] M. Kanter, Stable laws and embeddings of Lp-spaces, Amer. Math. Monthly 80 (1973), 403-407. [83] J.L. Krivine and B. Maurey, Espaces de Banach stables, Israel J. Math. 39 (1981), 273-295. [84] S. Kwapiefi and C. Schtitt, Some combinatorial and probabilistic inequalities and their applications to Banach space theory, Studia Math. 82 (1985), 91-106. [85] S. Kwapiefi and C. Schtitt, Some combinatorial and probabilistic inequalities and their applications to Banach space theory, II, Studia Math. 95 (1989), 141-154. [86] T. Landes, Normal structure and weakly normal structure of Orlicz sequence spaces, Trans. Amer. Math. Soc. 285 (1981), 523-533. [87] M. Ledoux and M. Talagrand, Probability in Banach Spaces, Springer-Verlag, Berlin (1991). [88] D.H. Leung, Embedding s into Lp,oc complementably, Bull. London Math. Soc. 23 (1991), 583-586. [89] D.H. Leung, s has a complemented subspace isomorphic to s Rocky Mountain J. Math. 22 (1992), 943-952. [90] D.H. Leung, Isomorphism ofcertain weak L p spaces, Studia Math. 104 (1993), 151-160. [91] D.H. Leung, Isomorphic classification of atomic weak L p spaces, Interaction Between Functional Analysis, Harmonic Analysis, and Probability (Columbia, MO, 1994), Lecture Notes in Pure and Appl. Math. 175, Dekker, New York (1996), 315-330. [92] D.H. Leung, Purely non-atomic weak LP spaces, Studia Math. 122 (1997), 55-66. [93] M. Levy, L'espace d'interpolation rdel (A O, A1)O, p contient s C. R. Acad. Sci. Paris S6r. I Math. 289 (1979), A675-A677. [94] J. Lindenstrauss and L. Tzafriri, On Orlicz sequence spaces, Israel J. Math. 10 (1971), 379-390.
Special Banach lattices and their applications
531
[95] J. Lindenstrauss and L. Tzafriri, On Orlicz sequence spaces, II, Israel J. Math. 11 (1972), 355-379. [96] J. Lindenstrauss and L. Tzafriri, On Orlicz sequence spaces, III, Israel J. Math. 14 (1973), 368-389. [97] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces, Lecture Notes in Math. 338, Springer-Verlag, Heidelberg (1973). [98] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces, I. Sequence Spaces, Springer-Verlag, Berlin (1977). [99] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces, II. Function Spaces, Springer-Verlag, Berlin (1979). [100] G.G. Lorentz, Some new functional spaces, Ann. Math. 51 (1950), 37-55. [101] G.G. Lorentz, On the theory ofspaces A, Pacific J. Math. 1 (1951), 411-429. [102] G.G. Lorentz, Relations between function spaces, Proc. Amer. Math. Soc. 12 (1961), 127-132. [103] H.E Lotz, Weak, convergence in the dual of weak LP, Preprint. [104] L. Maligranda, Orlicz spaces and interpolation, Seminfirios de Matem~itica 5, Universidade Estadual de Campinas (1989). [105] J. Marcinkiewicz and A. Zygmund, Quelques th~oremes sur les fonctions ind~pendantes, Studia Math. 7 (1938), 104-120. [106] M.B. Marcus and G. Pisier, Characterisations of almost surely continuous p-stable random Fourier series and strongly stationary processes, Acta Math. 152 (1984), 245-301. [107] A. Medzhitov and E Sukochev, The property (H) in Orlicz spaces, Bull. Polish Acad. Sci. Math. 40 (1992), 5-11. [ 108] V.D. Milman and G. Schechtman, Asymptotic Theory of Finite Dimensional Normed Spaces, Lecture Notes in Math. 1200, Springer-Verlag, Berlin (1986). [ 109] S.J. Montgomery-Smith, Boyd indices of Orlicz-Lorentz spaces, Interaction Between Functional Analysis, Harmonic Analysis, and Probability (Columbia, MO, 1994), Lecture Notes in Pure and Appl. Math. 175, Marcel Dekker, New York (1996), 321-334. [110] J. Musielak, Orlicz Spaces and Modular Spaces, Lecture Notes in Math. 1034, Springer-Verlag, Berlin (1983). [111] I.P. Natanson, Theory of Functions of a Real Variable, Vol. 1, Ungar, New York (1961). [112] N.J. Nielsen, On the Orliczfunction spaces L M (0, e~), Israel J. Math. 20 (1975), 237-259. [113] E.M. Nikishin, Resonance theorems and superlinear operators, Uspehi Mat. Nauk 25 (1970), 129-191; Engl. transl.: Russian Math. Surv. 25 (1970), 125-187. [ 114] I. Novikov and E. Semenov, Haar Series and Linear Operators, Kluwer Acad. Publ., Dordrecht (1997). [115] S.Ya. Novikov, Cotype and type of Lorentzfunction spaces, Math. Notes 32 (1982), 586-590. [116] S.Ya. Novikov, E.M. Semenov and E.V. Tokarev, On the structure of subspaces of the spaces A p(#), Amer. Math. Soc. Transl. 136 (1987), 121-127. [117] R. O'Neil, Convolution operators and L ( p , q ) spaces, Duke Math. J. 30 (1963), 129-142. [118] R. O'Neil, Integral transforms and tensor products in Orlicz spaces and L(p, q) spaces, J. Analyse Math. 21 (1968), 1-176. [119] G. Pisier, Factorisation of operators through L p, ~ or L p, 1 and non-commutative generalisations, Math. Ann. 276 (1986), 105-136. [120] B. Randrianantoanina, Injective isometries in Orlicz spaces, Function Spaces (Edwardsville, IL, 1998), Contemp. Math., K. Jarosz, ed., Amer. Math. Soc., Providence, RI (1999), 269-287. [121] Y. Raynaud, Deux nouveaux examples d'espaces de Banach stables, C. R. Acad. Sci. Paris S6r. I Math. 292 (1981), 715-717. [122] Y. Raynaud, Sur les sous-espaces de LP (Lq), S6minaire d'Analyse Fonctionelle 1984/1985, Publ. Math. Univ. Paris VII, 26, Univ. Paris VII, Paris (1986), 49-71. [123] Y. Raynaud, Sous-espaces s et ggomdtrie des espaces LP(Lq) et L 4~, C. R. Acad. Sci. Paris S6r. I Math. 301 (1987), 299-302. [124] Y. Raynaud, Finie reprdsentabilitg de s dans les espaces d'Orlicz, C. R. Acad. Sci. Paris S6r. I Math. 304 (1987), 331-334. [125] Y. Raynaud, Almost isometric methods in some isomorphic embedding problems, Banach Space Theory (Iowa City, IA, 1987), Contemp. Math. 85, Amer. Math. Soc., Providence, RI (1989), 427-445. [126] Y. Raynaud, On Lorentz-Sharpley spaces, Interpolation Spaces and Related Topics (Haifa, 1990), Israel Math. Conf. Proc. 5, Bar-Ilan Univ., Ramat Gan (1992), 207-228.
532
S.J. Dilworth
[127] Y. Raynaud, A note on symmetric basic sequences in L p ( L q ) , Math. Proc. Cambridge Philos. Soc. 112 (1992), 183-194. [128] Y. Raynaud and C. Schtitt, Some results on symmetric subspaces of L 1, Studia Math. 89 (1988), 27-35. [129] S. Reisner, A factorization theorem in Banach lattices and its application to Lorentz spaces, Ann. Inst. Fourier (Grenoble) 31 (1981), 239-255. [130] V.A. Rodin and E.M. Semenov, Rademacher series in symmetric spaces, Anal. Math. 1 (1975), 207-222. [ 131 ] H.E Rosenthal, On the subspaces of L p (p > 2) spanned by sequences of independent random variables, Israel J. Math. 8 (1970), 273-303. [132] C. Schtitt, Lorentz spaces that are isomorphic to subspaces of L | , Trans. Amer. Math. Soc. 314 (1989), 583-595. [133] C. Schiitt, On the embedding of 2-concave Orlicz spaces into L l, Studia Math. 113 (1995), 73-80. [134] A.A. Sedaev, The H-property in symmetric spaces, Teor. Funkcii Funkcional. Anal. i Prilozen. 11 (1970), 67-80 (Russian). [135] E.M. Stein and G. Weiss, Introduction to Fourier Analysis on Euclidean Spaces, Princeton University Press, Princeton ( 1971). [ 136] E Wojtaszczyk, Banach Spaces for Analysts, Cambridge University Press, Cambridge (1991).
CHAPTER
13
Some Aspects of the Invariant Subspace Problem
R Enflo and V. Lomonosov Department of Mathematics, Kent State University, Kent, OH 44242, USA E-mail: [email protected]
Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. Invariant subspaces of algebras containing compact operators . . . . . . . . . . . . . . . . . . . . . . . . 2. Generalizations of Burnside's theorem in Banach spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Generalizations of Burnside's theorem in Hilbert spaces . . . . . . . . . . . . . . . . . . . . . . . . . . .
535 535 538 543
Sketch of the proof . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Extremal vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Definitions and some basic properties of extremal vectors . . . . . . . . . . . . . . . . . . . . . . .
543 545 546
4.2. Invariant subspaces and a two sequences theorem
...........................
547
5. Operators without invariant subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.0. Introduction and history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Outline of the original considerations and construction . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Remarks on some other constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. More examples of transitive operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
549 549 550 555 556
5.4. A transitive operator on el . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
556 558
H A N D B O O K OF THE G E O M E T R Y OF B A N A C H SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 533
This Page Intentionally Left Blank
Some aspects of the invariant subspace problem
535
Introduction In this survey we present some circles of ideas and results which we feel constitute progress on the theory of invariant subspaces. We are not trying to give a complete survey of the theory as it stands today. There are many important techniques and directions - especially for operators on Hilbert s p a c e - that will not be discussed here. For some of these we refer the reader to [1,6] where also more references can be found. Behind the techniques of the first four sections in this survey is a common general idea: to find on the boundary of a convex set, points which are hard to move into the set by operators from a given s e t - and the connection between such points and invariant subspaces. In Section 1, this idea is carried out in the context of locally convex spaces. There we give results on invariant subspaces of algebras containing a compact operator. In Section 2 we consider first general Banach spaces. We present a generalization of the classical Burnside theorem for operator algebras. We also give connections to invariant subspaces of the so called two sequences theorem. So far this two sequences theorem has been proved only for Hilbert space. It should be an interesting problem to decide for which Banach spaces it is true. We discuss this further in Section 2. In Section 3 we present strengthenings of these results to operator algebras on Hilbert space and an invariant subspace theorem for essentially self-adjoint operators on real Hilbert space. The techniques in Section 3, developed by A. Simoni6 [37] - use the same underlying ideas as those in the first two chapters. However Simonic develops them in an impressive way that, for now, seems to work only in Hilbert space. In Section 4 we present techniques of extremal vectors and their connection to invariant subspaces. This techniques have, so far, been developed only in Hilbert space. However, they are, in essence, Banach space techniques, and also here, it should be interesting to find out for which Banach spaces they work. The techniques present a more constructive way to find the points which are hard to move. This leads, in particular to a strengthening of the two sequences theorem in Section 2 for some special cases. In Section 5 we present counterexamples to the invariant subspace problem. We also present results where counterexamples are used to support or disprove other conjectures.
1. Invariant subspaces of algebras containing compact operators Let E be a complex locally convex space, L(E) the algebra of continuous linear operators on E and R a sub-algebra of L(E). R is said to be transitive if E does not contain nontrivial closed subspaces that are invariant under each operator in R. A subspace M in E is said to be hyper-invariant for a subset S in L(E) if it is invariant under each operator which commutes with each operator from S. In 1935 von Neumann proved that each compact operator acting on a Hilbert space is non-transitive. In 1954 N. Aronszajn and K.T. Smith [4] generalized this statement to Banach spaces. In 1966 A.R. Bernstein and A. Robinson [10], using non-standard analysis, proved that a polynomially compact operator on a Hilbert space is non-transitive. E Halmos [21] converted their proof into one that uses only classical concepts. In 1968 W. Arveson and J. Feldman [5] proved that an algebra of operators on a Hilbert space with
536
P. Enflo and V. Lomonosov
one quasi-nilpotent generator is non-transitive if it contains a compact operator. In 1973 C. Pearcy and N. Salinas [27] proved Corollary 2.2 of this paper for the case of an algebra of operators with one generator. In 1973 the second author gave a new approach to invariant subspace theorems [23]. The most general result proved by this method was the following THEOREM 1.1 ([22]). Let R be a sub-algebra of L(E). The algebra R is dense in L ( E ) with respect to the weak operator topology if and only if it is transitive and its closure in this topology contains a nonzero compact operator. PROOF OF THEOREM 1.1. LEMMA 1.2. The algebra R is transitive if and only if for each nonzero vector x E E its orbit Rx = UAER A x is dense in E and if and only if for each nonzero functional f E E* its orbit UAER A* f is dense in E* in the w*-topology. Proof of the lemma is obvious. LEMMA 1.3. If a transitive algebra R E L ( E ) contains a nonzero compact operator K, then this algebra contains a compact operator with a nonzero fixed vector. PROOF. From the definition of compact operator it follows that there exists an open convex neighborhood of zero U in E such that the closure K (U) of the set K (U) is compact. Let us pick a vector y E E such that K y ~ O. Put V -- U + y. It is clear that the vector y can be chosen in such a way that the origin is not contained in the set K ( V ) , that is ( - K y ) q~ K ( U ) . From Lemma 1.2 it follows that for any element x ~ K ( V ) there exists an operator A 6 R such that A x ~ V. Therefore the pre-images A - I ( v ) of the open set V generate a covering of the compact K (V). We can choose from that covering a finite sub-covering {Ui }, i = 1 . . . . . n. By construction there exist operators A 1. . . . . An in R such that Ai (Ui) Q V. Since a compact topological space is normal, there exists a partition of the identity corresponding to the covering {Ui }, therefore there exist continuous functions f l . . . . . fn in K ( V ) such that 3~ ~> 0, Y~in_=l j~ ~ 1, f/(x) - - 0 if x q~ Ui. Consider a map 4~" K ( V ) --+ E defined by q~(x) -- y~in 1 fi(x)Aix. The right hand side of this equality is a convex combination of elements from the convex set V, therefore 4~ (x) 6 V if x E K (V). The composition K 4~ of the maps K and 4~ is a continuous map of the convex compact K (V) into itself. By the Schauder-Tikhonov fixed point theorem the map K4~ has a fixed point, therefore there exists an element xo ~ K ( V ) such that K ~ ( x o ) = xo and Y~in=l K 3 ~ ( x o ) A i x o - xo. The operator K1 -- ~Qinl ~ ( x o ) K A i is compact, is contained in R and has the fixed vector x0. Recall that x0 7~ 0 because K (V) does not contain the zero vector. [3 C O R O L L A R Y 1.4. Let A be a linear continuous operator on a complex locally convex space E which is different from a scalar operator and commutes with a nonzero compact operator in E. Then the operator A has a nontrivial closed hyper-invariant subspace, that is a subspace invariant for all operators commuting with A.
Some aspects of the invariant subspace problem
537
This seems to be the first invariant subspace result for general locally convex spaces. PROOF. Let R be the set of all operators from L ( E ) commuting with A. Assume that R is transitive. Then according to Lemma 1.3 there exists a compact operator K E R and an element x0 6 E such that K xo = xo ~: O. Let E1 be the linear subspace of E consisting of all fixed vectors of the operator K. It is clear that E1 is a finite dimensional subspace invariant for the operator A. Therefore A has an eigenvector in E with an eigenvalue )~. Let E2 be the subspace consisting of all eigenvectors of the operator A corresponding to the eigenvalue )~. This subspace is not empty and does not coincide with E because the operator A is not scalar. It is clear that this subspace is invariant for R, so R is not transitive, ff] COROLLARY 1.5 (Schur's lemma). If a transitive algebra of linear continuous operators in a complex locally convex space contains a nonzero compact operator, then the commurant of this algebra consists of scalar operators. LEMMA 1.6. If a transitive algebra R C L ( E ) contains a nonzero finite dimensional operator K, then this algebra contains a nonzero one dimensional operator F(e ~, e) such that F(e ~, e)x -- (e',x)e, where e ~ ~ E*, e, x E E. PROOF. Put E1 -- K (E). From Lemma 1.1 it follows that vectors of form K A x , where A 6 R, generate a dense set in E1 so that operators of form KA restricted to E1 generate a transitive algebra of operators in El. According to the classical Burnside theorem this algebra coincides with L (El). Therefore there exists an operator A 6 R such that K Ax = (e', x)e for any x c El, where e 6 E l , e' 6 E*. [] LEMMA 1.7. If a transitive algebra R C L ( E ) contains a nonzero finite dimensional operator, then the closure of the algebra R in the weak operator topology coincides with L(E). PROOF. Let F(et, e) be an operator which exists by Lemma 1.6. Since AF(et, e ) = F(e', Ae) and F(e ~, e)A = F(A*e ~, e), where A* is an operator dual to A, from Lemma 1.2 it follows that the closure of the algebra in the weak operator topology contains any one dimensional operator and therefore any finite dimensional operator. Finite dimensional operators are weakly dense in L ( E ) , so the lemma is proved. [-q LEMMA 1.8. Let K be a compact operator in E, p ( K ) its spectral radius. If p ( K ) < 1, then the sequence {K n } of powers of the operator K converges to zero in the strong operator topology. PROOF. Let U be an absolutely convex neighborhood of zero in E which the operator K maps into a relatively compact set and let p be the Minkowsky functional corresponding to this neighborhood. Let E u be the quotient space E~ ker(p). The norm in E u is generated by the functional p. Let the Banach space E u be the completion of the space E u . The canonical map 1 7 : E --+ E u is continuous. Since the kernel of the operator K contains
538
P. Enflo and V. Lomonosov
the kernel of the map H in the space E u , the operator K 1 is well defined by the formula Kl x = H K H - 1x. The operator K1 is extended continuously to an operator K2 acting in the space Eu. It is easy to check that the operators K1 and K2 are compact and K, KI and K2 have the same nonzero eigenvalues. Therefore the spectral radii of these operators coincide. From Gelfand's spectral radius formula it follows that A
lim p ( K ~ / T x ) - -
n - + oc~
lim p ( K ~ H x ) =
n-+oo
lim
n - + cx3
p(Knx)=O
(1)
for each element x E E. For this reason Knx ~ U, Kn+lx ~ K ( U ) for sufficiently large n and the set {Knx} is relatively compact. Let S be the set of the condensation points for the sequence {Knx}. It is clear that K ( S ) -- S and that the set S is contained in the kernel of /7 since l i m n ~ p ( K n x ) = 0. This implies that S E ker(K) and consequently S = {0}. 7q LEMMA 1.9. If a transitive algebra R contains a nonzero compact operator, then it contains a sequence which converges in the strong operator topology to a nonzero finite dimensional operator PROOF. By Lemma 1.3 the algebra R contains a compact operator K with an eigenvalue
)~l -- 1. Let )~l . . . . . )~n be all eigenvalues of K which are no less then 1 by modulus, let/Zl . . . . . /Zn be the maximal dimensions of Jordan cells in the corresponding spectral invariant subspaces and let M be the linear span of those subspaces. By the Leray-Riesz spectral decomposition Theorem [15] the space E is the direct sum of M and another subspace N, which is also invariant under K, and the spectral radius of the restriction KIN is strictly less than 1. Let Q be the projection operator onto M parallel to N. Put P()~) = (~. - 1) # 1 - 1 1 - - I i n _ 2 ( ) ~ )~i) lzi. It is clear that the operator P ( K ) has a nonzero restriction P(K)]M and P ( K ) K n x -- K n P ( K ) X = P ( K ) x for each x 6 M and for all nonnegative integers n. For each x E E put x = x l + x2 for x l 6 M, x2 E N. Then -
-
P ( K ) K n x : P ( K ) K n x l -+- P ( K ) K n x 2 = P ( K ) x l + P ( K ) K n x 2 . The second term in this formula converges to zero by Lemma 1.8. So the sequence P (K) K n converges in the strong operator topology to the finite dimensional nonzero operator P ( K ) Q . It is clear that P ( K ) K n E R, which finishes the proof. D Combining Lemmas 1.7 and 1.9 completes the proof of Theorem 1.1.
2. Generalizations of Burnside's theorem in Banach spaces In 1991 Theorem 1.1 was generalized for operator algebras in the following way [25]. In Theorem 2.1 below, let B be a Banach space, L ( B ) the bounded linear operators on B and K (B) the compact operators on B. Let IIIA III denote the essential norm of A, that is the distance from A to the space of compact operators.
Some aspects of the invariant subspace problem
539
THEOREM 2.1. Let R be a weakly closed sub-algebra of L(B), R ~ L(B). Then there exists x E B** and y ~ B*, x ~ 0 and y ~ O, such that for every A E R ](x, A'y)[ ~< IIIAIII.
(1)
For a given subset S C L ( B ) and a vector x 6 B put Sx UA6S Ax and S' = U A*. We say that S is transitive if it does not have a nontrivial invariant subspace and S is essentially transitive if the conclusion of Theorem 2.1 is false for S. We say that a subalgebra R C L (B) has the Pearcy-Salinas (PS) property if there exists net {Ac~} C R and a nonzero operator A 6 L (B) such that: =
lim(x , A ~*y ) - - ( x , A * y)
(2)
Ol
for every vector x E B** and every functional y 6 B* and lim IIIAc~ III - 0.
(3)
ot
Of course, every bounded operator is a weak limit of finite dimensional ones, so it is the assumption {A~ } C R that makes the condition nontrivial. COROLLARY 2.2. Let R be a weakly closed proper sub-algebra of L ( B ) with the (PS) property. Then the algebra R ~ is non-transitive. PROOF. Let x and y be as in Theorem 2.1. Then for every pair of operators T1, T2 in R we have
I(x,
T~A*TOy)I -
liml(x,
T2A~TOy)I <~limlIIT2A~TIlI[
~< lim IIT2 II IIZ~ IIIIIAc~ IIIo/
It is easy to see that one of the three subspaces
A ker(T*), TcR
R'y,
Span
(T~A*T;y)
Tj , T 2 c R
is a nontrivial invariant subspace for the algebra R f. From Corollary 2.2 we get the following strengthening of the Pearcy-Salinas theorem [27].
COROLLARY 2.3. Let T be a bounded operator on a Banach space and let R be its commutant. Assume that there is a nonzero operator A and net {Ac~} of operators that commute with T such that (2) and (3) hold. Then the algebra R I is non-transitive.
P. Enflo and V. Lomonosov
540
The commutant of T is a proper sub-algebra of L(B), so Corollary 2.3 is an immediate consequence of Corollary 2.2. In Hilbert space Theorem 2.1 has the following equivalent formulation. THEOREM 2.4. Let R be as in Theorem 2.1 and let B be a Hilbert space. Then there exist 11)
113
two bounded nets (x~) in B and (y~) in B such that x~ --+ x # O, y~ --+ y ~ O, and for every A ~ R (Axe, y~) --+ O. PROOF OF THEOREMS 2.1 A N D 2.4. Let B be a Banach space, Q a compact Hausdorff space, R a sub-algebra of L ( B ) , WQ the Banach space of weak*-continuous functions Q ~ B* with sup-norm, SQ the subspace of WQ consisting of the strongly continuous functions and NR the algebra of norm-continuous functions Q --+ R t. Let C = C (Q) be the algebra of all complex-valued continuous functions on Q. Every function f 6 C (Q) is a linear operator on the space WQ. We let f * denote its adjoint. DEFINITION. Let M be a subspace of WQ, invariant under the algebra C. We say that 0 6 M* is a point functional if there exists a point q 6 Q such that
f*O = f (q)O
(4)
for every function f E C. If E is a subspace in a Banach space let UE denote the unit ball of E and E • denote the annihilator of E in the dual space. We say that a functional 0 6 M* is a 6 (q, x) functional or simply 6-functional if there exists a point q E Q and an element x 6 B** such that if h 6 M then
O(h)-(x,h(q)).
(5)
LEMMA 2.5. Let 0 ~ M* be a point functional and assume M D SQ. Then the restriction, X = OIsQ is a 8-functional. Let H be a Hilbert space with orthogonal basis {en}, LIM(an) a Banach limit in the space l ~ , Q the compactification of the integers by {cx~}. Then O(h) = LIM(h(n), en) is an example of a point functional that is not 8-functional. Let Q x U8 be the topological product of the compact Q and the ball U8 with norm topology. The functions s ~ WQ and f 6 C (Q) define complex valued functions L (s) and Ll ( f ) on Q x U8 bythe formulas L(s)(q, x) = (x, s(q)), Ll ( f ) ( q , x) = f (q). LEMMA 2.6. L(s), L1 ( f ) are continuous functions on the set Q • UB. LEMMA 2.7. Let r ~ L(M)*, f ~ C(Q) and (L1 f)*cp = gcP. Then f*(L*qo) = g(L*qg).
541
Some aspects of the invariant subspace problem
PROOF. F o r a given function h E WQ we have (h, f * (L* go)) -- ( f h, L'go) -- ( L ( f h), go) -(L1 ( f ) L ( h ) , go) -- (L(h), L1 ( f ) * go) -- (L(h), Fgo) -- (h, FL*go). [] LEMMA 2.8. Let M, N be subspaces of WQ, invariant for C(Q). Let M ~ N and let 0 E M* be an extreme point of UN• Then 0 is a point-functional. The proof of this lemma is similar to de Branges famous proof of the Stone-Weierstrass Theorem [ 13]. LEMMA 2.9. Let T E K ( B ) and h E WQ. Then T*h ESQ. Let R be an algebra in L ( B ) and let h E WQ, IIh II - 1. By applying the algebra N R to h we get subspaces
N-
NR(h),
M-
Span(N, SQ).
then M D N and M and N are invariant for the algebra C (Q). Let 99 be a point functional, q9 E M*. Then, by L e m m a 2.5, golso is a 6(q, x) functional. LEMMA 2.10. If the functional go is in N • then for every operator A E R, (6)
I(x, A*h(q))l ~ 21gollllAIIIIhl.
Let OfF(T) be the subset of the spectrum of an operator T E L ( B ) consisting of the isolated points of finite multiplicity. LEMMA 2.1 1. Let the subspace L C B have finite co-dimension. Assume ~ > 0 and
Ilrxll < or. Ilxll
sup-
~L
(7)
Assume y E o-(T) and lyl > o~. Then y E CrF(T). LEMMA 2.1 2. If R is an essentially transitive sub-algebra, then R contains an operator T with thefollowingproperties" 1 E OfF(T) and ifF E cr(T)\CrF(T) then lyl < 1/2. LEMMA 2.13. Let R be a uniformly closed essentially transitive sub-algebra of L(B). Then R contains a nonzero finite dimensional projection. PROOF. Let T be an operator as in L e m m a 2.12. Since {9/" (T - F I ) - I E R} is a component of the resolvent set of T it follows that if (T - )/I)-1 exists and [VI ~> 1/2 then ( T - y i ) - I E R. By L e m m a 2.12 there exists a circle cr C C such that 1 is the only point of or(T) inside c~ and for every point y E o~, (T - y I) -1 E R. By the Riesz theorem P - -27ri 1 for (T -
V 1) -1 dy is a non-zero finite dimensional projection and P E R ~. If P -- P~ then P1 is finite dimensional projection in R. []
P. Enflo and V. Lomonosov
542
Now to finish the proof of Theorem 2.1 we need the following well-known fact [29]. LEMMA 2.14. I f a transitive algebra R contains a non-zero finite dimensional operator, then R is weakly dense in L ( B ). Essential transitivity implies transitivity, so by combining Lemmas 2.13 and 2.14 we finish the proof of Theorem 2.1. Of course Lemma 2.13 gives the possibility of obtaining different results on density. For example one of them is THEOREM 2.15. I f H is a Hilbert space and R is a uniformly closed essentially transitive sub-algebra of L ( H ) then R contains all compact operators in L ( H ) . Now we will prove the equivalence of Theorem 2.1 and Theorem 2.4 in Hilbert space. We first prove that Theorem 2.4 implies Theorem 1 if B is Hilbert space. Assume that Theorem 2.4 holds. Let C = lirl~ Ix~ I" lY~ I. If A 6 R then 0 = lim(Ax,~, y~) -- lim(A(x,~ - x), (y~ - y)) - (Ax, y) + lim(Ax, yot) 13l
lY
+ lim(Ax~, y), (y
where lim(Ax, y~) = lim(Ax~, y) = (Ax, y). ly
o/
Thus (Ax, y) = lim(A(x - x~), (y~ - y)). If K is a compact operator, then
l(Ax,
Y)I ~< li-~ I((A- g)(x -x=),
(y~ -
Y))I + l~ml(g(x -x~),
(y~ -
Y))I
I I A - KII4C. Now let z = y / ( 4 C ) . Then I(Ax, z)] ~< i~cf IIA - gll = IIIAIII, so Theorem 2.1 follows from Theorem 2.4. To prove that Theorem 2.1 implies Theorem 2.4 we use the following result of Glimm [20]. B is assumed to be Hilbert space. LEMMA 2.16. Let qJ ~ L(B)* and q/ ~ K (B) • Then there is a pair ofbounded nets (x~) and (y~) such that
Some aspects of the invariant subspace problem
543
(1) w-limaXa = 0 , w-lima Ya = 0 . (2) For every A ~ L ( B ) , q-/(A) = lirna(Axa, Ya). Now let q-'l (A) -- ( A x , y) where x, y are as in Theorem 2.1 and A 6 R. Theorem 2.1 gives that qq can be extended to Span(R, K ( B ) ) by putting q-tl ( T ) - - 0 for T E K ( B ) . Now let q-' be a Hahn-Banach extension of the qq to all of L (B). We now use L e m m a 2.5 and we get Theorem 2.4 with xa = x 4- Xa and Ya = Y - Ya. 7] Finally we will mention that in the case of a non-reflexive Banach space Theorem 2.1 gives invariant subspace corollaries only in the dual space B*. Since the first author in 1976 [16,17] showed that there are counter examples to the invariant subspace problem in general Banach spaces, this may be a sign that the following result is true: CONJECTURE 2.17 ([25]). I f A is a bounded linear operator in a Banach space then A* has a non-trivial invariant subspace. The known counterexamples do not contradict this conjecture. By Corollary 2.2, this would be true if the following is true: If A is a bounded linear operator in a Banach space then there exists an algebra with the (PS) property which contains the operator A. S.W. Brown [11] independently obtained similar results in the case of a commutative algebra.
3. Generalizations of Burnside's theorem in Hilbert spaces Given an operator A on a Hilbert space we define Re A - - A + A2 * and I m A A-A* With 2 " the normalizing conditions Ixl = lYl = 1 inequality (1) from Section 2 has the form -
I(x, A *y)[ <~ c IIIA III.
-
( 1)
In 1996 A. Simonic found an exact form of inequality (1) for operator algebras in a complex Hilbert space by proving the following THEOREM 3.1. Let R be as in Theorem 2.1 and let H be a Hilbert space. Then there exist two nonzero vectors f , g E H such that f o r every A ~ R IRe(Af, g)l <~ 111Re Alil(f, g).
(2)
Sketch o f the p r o o f
Let H be a real or complex Hilbert space. Fix a unit vector f0 6 H and choose a positive r 6 (0, 1). The set S is defined as follows: S-
{ f E H" ]if0 - fll ~< r}.
544
P. Enflo and V. Lomonosov
LEMMA 3.2. Let W be a subset of S and let A be a bounded linear operator on H. Suppose that Re(A f, f0 - f ) ~> 3 ~> 0,
for all f ~ W. Then there exists a positive number lZ >~O, such that for any e ~ (0, #):
lifo - ( z + cA)f I[ < lifo -
f[[,
for all f ~ W. DEFINITION. For a fixed ball S = {f 6 H: lifo - fl[ ~< r} around the unit vector fo 6 H, the polar hyper-plane Ps, of S with respect to the origin, is defined by the following set:
P s - - f E H: (f, fo)-----1--r 2. LEMMA 3.3. The function Ao : S w-> S, defined by 1
A o ( f ) = r2 + (f, fo) f ' maps the set S weak-continuously into itself Furthermore, the set of all fixed points for Ao is equal to Ps (3 S. DEFINITION. For every operator A in L ( H ) define a real function A A " S ~ ~ as follows: 1
A A ( f ) ---- r2(1 _ r2 ) Re(Af, fo - f ) . LEMMA 3.4. Suppose f is a vector in the polar hyper-plane Rs C3S, satisfying the following strict inequality for some A E L ( H ) : AA ( f ) > IIIRe
AIIIAI(f).
Then there exists a positive number ~ > O, together with a weak neighborhood W of f , such that for every h ~ W U S: AA(h) > IIIReAIIIAI(h)I +S. PROPOSITION 3.5. Let A ~ F~(TY) be a convex subset of the bounded linear operators acting on a real or complex Hilbert space H. Fix a unit vector fo E H and choose a positive number r 6 (0, 1). Suppose that for every vector g _J_fo and I[gl[ <<,1 there exists an operator A ~ A satisfying the following strict inequality:
ReA
f0+
/ 1- r z g
,fo-
r
g > > ][[Re A [[[(1 - [[g [[2).
Some aspects of the invariant subspace problem
545
Then A contains an operator Ao with an eigenvector in the set S={f6Hlllfo-fll~
4. Extremal vectors Extremal vectors were introduced in Ansari and Enflo [2] and Enflo [19]. For results on special operators see also Spalsbury [38]. Extremal vectors provide a tool to find hyperinvariant subspaces for compact and normal operators on Hilbert space in a unified, more constructive way. They also provide a tool to strengthen Theorem 2.4 in some special cases. For most of the proofs we refer the reader to [2]. Since Theorem 4.8 is presented here for the first time, we give a complete proof. We feel that the extremal vector technique for obtaining invariant subspace results is still in its beginning. The origin of this is to study best approximate solutions of the equations Tny = xo for the case when T has dense range and x0 is not in the range of T. More precisely, one considers, for e > 0 and for each n, the Yn of minimal norm such that
II
- xo 11
It turns out that T n Yn in many cases converges to a vector which is hyper-noncyclic for T, that is non-cyclic for all operators commuting with T.
546
P. Enflo and V. Lomonosov
4.1. Definitions and some basic properties o f extremal vectors H will denote a Hilbert space over the field of complex numbers, R ( T ) will represent the range of T. Backward minimal vectors. Let T ' H ~ H be a bounded operator with dense range. Let x0 6 H and e > 0 with e < [Ix0ll. There is a unique vector Yn,xo such that T Y n , x o - x0l] ~< e and
]lYn,.0II- inf{llyl[-
[[Zny
-
x0]] ~ e}.
The points yen,xo are called backward minimal vectors. Forward minimal vectors. Let T" H --+ H be an injective bounded operator. Let xo 6 H and e > 0 with e < Ilxoll. There is a unique vector V.,x E o such that IIV n~, x o -- XOII ~< ~ and
I[Tn Un,xo ~ II =
inf{ I1Tn v 11" 11v - x0 l[ ~<e } 9
The points T n Un,xo e a r e called forward minimal vectors. W h e n there is no ambiguity, we will drop some of the superscripts and subscripts from Yne,xo and Vne,x0. Sometimes the vectors T n Yn and Vn will also be referred to as backward, respectively, forward minimal vectors. It is clear that [[T n Yn,xo e - xo 11= e and [IU n - - XO 11 -e. The minimality gives easily Orthogonality relations.
If rn I Yn then Tnrn I T n Yn - x o . And we get THEOREM 4.1 (Orthogonality equations). There exist 6~ < 0 such that ye = ~ T* (Tye - xo) a n d (possibly another) 6~ < 0 such that (rE - xo) -- 6~ T * T v e .
REMARK 1. From the orthogonality equations we obtain Tnyn = - S n ( I - 8 n T n T * n ) - l T n T * n x o
and
Vn : (I - 6 n T * n T n ) - l x o .
Let An - - 6 n ( I - 6 n T n T * n ) - l T n T *n and Bn - - ( I 6nT*nTn) -1. Then An and Bn are positive operators. ][An][ < 1 and [[Bn[[ = 1 for all n. Let f n ( z ) - 1-8,.~ -~,.z . Then fn (z)
Some aspects of the invariant subspace problem
547
is analytic in a neighborhood of a ( T n T *n) and a ( A n ) = f n ( a ( T n T * n ) ) . The maximum -~n'llrnll2 . It follows that point of f ( a ( T n T * n ) ) is i---~.:ii~l~
[IAnll =
- a n . IIT n I12 1 -
~n"
< 1.
IITn II2
1 Then gn is analytic in a neighborhood of a ( T n T *n) and Now let g n ( z ) 1-~n.z a ( B n ) = g n ( a ( T n T * n ) ) . Since 0 E a ( T n T *n) it can be verified that the maximum point of a (T n T *n) is 1. Hence l]Bn l] = 1. The proofs of the following results can be found in [2]. .
THEOREM 4.2. Let xo E H and xo =/: O. The functions e w+ ye and e w-~ ve are analytic on (0, Ilxoll). THEOREM 4.3. For any b o u n d e d operator T with dense range, if xo q~ R(T), then Tyen-xo
Tyen
\
IIT y~, - xoll' IIT y~, II f o r some sequence 6n ~
/ --~0
O.
4.2. Invariant subspaces and a two sequences theorem LEMMA 4.4. I f T is quasi-nilpotent then IlYnk II/llYnk+l II ~ 0 as k --+ cx~f o r some subsequence nk. This follows easily from the spectral radius formula. In the following theorem, let K be a quasi-nilpotent, compact operator without eigenvector and let (nk) be as in L e m m a 4.4. We have THEOREM 4.5. For every weak accumulation point z o f K nk Ynk, K z is hyper-noncyclic f o r K and { T z [ K T = K T } is a hyper-invariant subspace f o r K. PROOF. By passing to a subsequence, we can assume K nk+l Ynk + 1 -'-+ YO weakly. We have T K n k + 1Ynk --+ T K z
in norm.
Put Tynk -- OtkYnk+l + rk rk 3_ Ynk+l. Then, obviously, Olk --+ 0 as k --+ cx~. We get T K nk + 1Ynk --- OtkK nk + 1Ynk + l -+- Knk + l rk 9
548
P. Enflo and V. Lomonosov
Scalar multiplying both sides by K nk + 1Ynk+ 1
--
XO
gives
(TKnk+lyn k, Knk+lyn~+l -- xo) = Otk(Knk+lynk+l, Knk+l ynk+l -- XO) + (Knk+l , Ynk+l Knk+lynk+l -- XO).
The first term on the fight-hand side --+ 0 since otk --+ 0. The second term is 0 because of the orthogonality relations 9 The left-hand side --+ ( T K z , Yo - xo) as k --+ cx~. Since the yf angle between Knyn and x o - Knyn is/> ~, Ilyo][2 ~< liminflIKnynll 2 ~ I[xoll2 - e 2. Thus Yo - xo # 0 and the theorem is proved. O We also have THEOREM 4.6. Assume that T is normal with dense range R ( T ) and that T has a cyclic vector. Assume xo ~ R ( T ) 9 Then f o r each e > 0 Tny rel , X 0 converges in norm to a hypernoncyclic vector f o r T. To prove the following two sequences theorem we use the following THEOREM 4.7. Let xo ~ H, xo q~ R ( T ) . I f f o r every 61 > 0 and all 62 with 0 < 62 < 61, there exists r > 0 with (T n Yne, Tn Yne -- xo) <~ --r f o r all n >~ 1 and all e, $2 ~ 6 ~ SI, then xo is hyper-noncyclic f o r T.
From this we derive the following THEOREM 4.8. Let R be a commutative algebra of operators on H which contains a O9 O) quasi-nilpotent operator. Then there exist two sequences Xn --+ x # 0 and Yn ~ Y # 0 such that f o r every bounded sequence An of operators in R, (Anxn, Yn) --+ O. PROOF. Let T E R be quasi-nilpotent. Assume xo ~ R ( T ) . Then by Theorem 4.7 either x0 is hyper-noncyclic or, for some e > 0 and some subsequence (n~) of the integers
(1)
(T n~ Yn~ , Tn~ Yn~ - xo) --+ O,
where (Yn~) are the backward minimal vectors with respect to x0 and e. We write Tn~yn~ = ( 1 - e2)x0 + y E V / 1 - e2s0 + V / 1 - y 2 E v / 1 - e2sv,
where Ilsoll = I l s ~ l l - 1 and sv -~ O. (1) then gives ((1 - e2)xo + yew/1 - e2so + V/1 - y2eV/1 - e2s~, e x o - yv/1 - e2xo
-,/1-
o.
549
Some aspects of the invariant subspace problem Now, let Av be a bounded sequence of operators in R, let
xv = ( 1 - e2)xo + y e ~ / 1 - e2so + V / 1 - v2ev/1 - e2sv and let
yv -- exo - Y v / 1 (2)
Then xv --+ (1
-
g2)x0
-+-
e2so - ~ / 1 - y 2 v / 1 -
ge~/1
e2sv. O)
-
g2s0
--
X and yv --+ exo - y~/1
-
62s0
=
y. Moreover
I(A~x~, yv)l -- [(A~T~Y.~, T"~Y.~ xo)l- [(T"~A~y.~, T"~Y.~ -
-
I(z~(~y~
-
xo)l
+ rn~), T n ~ y n ~ - xo)]
<~ ]Oln~[I(T n~ yn~ , Tn~ yn~
_
xo>l+ [(r" v rn~ ,
Tnv
Yn~ - xo)] .
(2)
Since Oln~ is bounded, by (1) the first term on the right-hand side of (2) --+ 0 as v ~ oc. The second term on the right-hand side of (2) is 0, because of the orthogonality relations. This proves the theorem. []
5. Operators without invariant subspaces 5.0. Introduction and history In this chapter we will present examples of operators on Banach spaces with only trivial closed invariant subspaces. We will refer to them as transitive operators. In this section we will give a short history. The first construction of a transitive operator was given by the first author, who completed and circulated in preprint form a first manuscript in 1975. An outline of the construction was presented in Seminaire Maurey-Schwarz in the Spring of 1976 [16]. (See, also the survey article from 1982 by Radjavi and Rosenthal [30].) An improved manuscript of the same construction appeared in the Institute MittagLeffler report series in 1980 and was submitted to Acta Math. in 1981. This version with some improvements in the notation and presentation- appeared in Acta Math. in 1987. In 1984, B. Beauzamy [7] and C. Read [31] presented constructions of transitive operators on some Banach spaces using a similar approach and basic considerations but with the technical aspects worked out in different ways - a discussion and comparison will be given below. Beauzamy's construction gave the first example of an operator T on a Banach space B where all non-zero vectors are super-cyclic: for every y 6 B, y 5/=0,
{t~TKy l t~ 6 R, K 6 N } -- B. A slight modification of Read's first construction gave a transitive operator on s
550
P. Enflo and V. Lomonosov
Simplifications of the construction of a transitive operator on s were given by A.M. Davie (see Beauzamy [8]) and by C. Read [33], whose example we will present below. This example has recently been used by V. Troitsky to test several conjectures on invariant subspaces. Later, C. Read [34-36] constructed stronger examples as well as transitive operators on new spaces and we present some of his results below. But, as we have seen above, there is no example of a transitive operator which is the adjoint of another operator, a n d - in p a r t i c u l a r - no example of a transitive operator on a reflexive space. In Section 5.1, we outline the construction in [16-18]; in Section 5.2, we compare it to the constructions in [7] and [31]; in Section 5.3, we give the stronger examples of Read, and in Section 5.4, we discuss the recent work of Troitsky.
5.1. Outline of the original considerations and construction In this section we will outline the original considerations and construction of an operator with only trivial invariant subspaces on a Banach space. The Banach space will be constructed at the same time as the operator and will be non-reflexive. There are very serious difficulties in carrying out a similar construction in a reflexive Banach space. So we feel that the construction gives some support to the conjecture that every operator on a Hilbert space has a non-trivial invariant subspace. We now turn to the basic considerations behind this approach. It is clear that every operator with a cyclic vector on a Banach space can be represented as multiplication by x on the set of polynomials under some norm. So what we will do is to construct a norm on the space of polynomials and prove that multiplication by x under this norm has only trivial invariant subspaces. Our next basic consideration comes from the fact that one can have an operator with a dense set of cyclic vectors without having all vectors cyclic. In order to be able to make some limit procedure work, we will construct the operator so that it has the following uniformity property: Let 1 be a cyclic vector of norm 1 in B. Let (p j) be a sequence which is dense on the unit sphere of B. For every m there is a positive number Cj,m such that for every Pn with [[Pj -- Phil < 1/2 m+4 there is a polynomial s in T with [[s <~ Cj,m such that [[s 111 ~< 1/2 m. It is easily verified that such a T has only trivial invariant subspaces (1.1). It follows easily from the fact that the spectrum of an operator is non-empty that there is no operator for which Cj,m depends only on j or only on m. If we have the operator T represented as multiplication by x, then s will just be multiplication by the polynomial s From now on we shall identify B with the completion of the vector space of all polynomials with real (or complex) coefficients under a suitable norm l] 1]. This leads us to the next basic consideration. Assume that we have a norm II l[ on the space of polynomials. Assume that p is a polynomial of norm 1 and assume that [[s - 111 ~< e and 11s[lop <<,K. This gives that for every polynomial h we have the inequality I l h l l - KIIhpll ~ IIh~p- h. 111 ~ EIIhllop.
Some aspects of the invariant subspace problem
551
This implies that if
1
Ilhllop ~ ~
Ilhll,
then
Ilhll
Ilhpll ~ ~ .
(1.2)
In order that the operator also satisfies (1.1), it is of course necessary that the inequality Ilhpll ~> Ilhll/2K holds uniformly in p in every ball of size e / 1 6 on the unit sphere, at least if we put e = 1/2 m. This uniformity will lead us to the concept of "concentration at low degrees" for polynomials, defined below. The concept is used also in later constructions of transitive operators. In many norms a neighborhood of a polynomial of degree n consists of polynomials with some concentration up to degree n. There is a sense in which the inequality (1.2) is sufficient for p to be moved close to 1 by a polynomial with small operator norm. This is given by our Theorem 5.1.2 below. In order to describe this theorem, we have to tell something about the way that we construct the final norm. It should be pointed out that this sufficiency of (1.2) depends on the fact that the norm constructed is non-reflexive. We do not know whether anything similar can be done in a reflexive space. Consider all pairs (q, e) where q is an arbitrary polynomial whose coefficients have real and imaginary parts rational, and e is of form 2 -k. We enumerate all such pairs and call the sequence (qn, en). We also insist that for a fixed q, if n l, n2 . . . . are all the integers such that qn = q, then e n l > En2 > en3 . . . . . Also we assume deg qn <~ n. Our construction will be completely determined by a sequence of polynomials ~n and constants C~ > 2. el . . . . . ~k and C1 . . . . . Ck will determine a number ak+l inductively as explained below and we define a sequence of norms as in the following definitions. DEFINITION. For any polynomial p, consider all representations p -- ~ a i , ~ x i g.~l I . . . g,~nn and put [Plopn -- inf Z where cients.
lai,3l 2i (Clel l1 )~' 9. . (Cnl~ n l1 )~n
I1~ denotes the usual el norm equal to the sum of the absolute values of the coeffi-
REMARK. In the final norm the operator x will have norm ~< 2, and multiplication by s norm ~< Ck I~11. DEFINITION. For any p, consider all representations
p -- r + ~
Sk (g~kakqk -- 1). 1
Put Ipl ~ = inflr]l -+- Z~lSklopnek. Put Ipl ~ = IPll, and let a~ be determined inductively by the condition la~qk I~- 1 = 1. REMARK. [g~kakqk -- 1[ n < ek and clearly the operator norm of multiplication by g is [g[opn. We see that [[n is the maximal norm satisfying the following four properties:
P. Enflo and V. Lomonosov
552
(A1) II n
~lll, (A2) le~a~q~ - 11n ~ Ek, k = 1,2 . . . . . n, (A3) [gklop <. Ckls (A4) Ix lop <. 2. Observe that II n and I lopn are decreasing sequences of norms and hence converge to some pseudo-norms. We write II II = l i m l l nTHEOREM 5.1.1. Assume that (Cn) and (gin) are sequences which define norms l] m as above and assume that there are sequences o f positive numbers Dn /7 cx~ and Ln f cx~ so that the following holds: (I) [plm is c o n s t a n t f o r m >~ ( d e g p ) - 1. Inparticular [qn[m is c o n s t a n t f o r m ~ n - 1. (II) For any n, consider all k <. n such that 6 n - - 6 k , and ]akqk - anqn[ (n-l) < en/16. Let K be the least such k. Then [s [l = L K , Cn = DK. Then the resulting limit norm defines a space B f o r which multiplication by x has no invariant subspace. PROOF. Let q [Iq l] = 1. Let e ank qnk --+ q in Hence for k
be an element of B, which we recall is the closure of all polynomials and be a fixed negative power of 2. Choose increasing nk such that enk = e and B. We can even insist that [lankqnk - q l[ < e / 6 4 . > 1, E
lankqnk -- an,qn, I(nk-1) --]lankqnk SO by (II), Cnk <, maxm~
Dm
-
II < ~ ,
anlqnl
and I~n~l~ ~ maxn~
II~n~q - 111 ~ Ilenkan~qn~ -- 111 +
Ilen~(q
Lm
so
that [gnk Iopnk is bounded ~< A.
--ankqnk)[[
<, e 4- [g'nk[opnk Ilq - ank qnk [I ~< e + A IIq -- ankqnk II. Letting k tend to infinity, we see that 1 is within distance e of the space generated by q and hence, letting e ~ 0, we see that 1 is in that space and hence the space coincides with B.
D
We will now drop ak in our notation so when it is clear from the context we will denote akqk by qk and assume Iq~l ~-l = 1. DEFINITION. ord p = degree of lowest order term of the polynomial p. DEFINITION. Let f be a positive real valued function defined on (0, cx~). We say that ~. - ~ j > ~ o a j x n j is more lacunary than f if ord ~ = no > / f (0)
Some aspects of the invariant subspace problem
553
and nj ~> f (n j_ 1)
for every j.
Our next theorem shows that, under the assumption of an inequality similar to (1.2) we can satisfy condition (I) as soon as the polynomials en are lacunary enough. THEOREM 5.1.2. Let el . . . . . eN, C1 . . . . . CN be given with Ck > 2. Assume f o r a l l h and some B that IhlopN Ihl ~
~.
1 6N+l
==~ IhqN+l IN
Ihl ~ B
Then, given K > 4B/eN+l, there exists a lacunarityfunction f such that if (1) leu+lll ---- K, (2) the lacunarity Of eN+l ~ f , (3) CN+I > 2, then with this choice of eu+l and CN+I we have [glN+l
_
iglN
for all g with degg ~< N.
We now assume that we have two sequences D n / 7 ec and Ln / ec. Assume that ]in--1 is defined. We will then define [in according to the following rule: consider all k ~< n such that ek = en and [q~ - qn in-1 < en/16. Let K be the least such k. Then [en l1 = LK, Cn = DK. If this rule is fulfilled for all n ~< N, we say that [iN is defined in a compatible way from the sequences D~ and Ln. If for every N, [ [U is defined in a compatible way from the sequences Dn and Ln, then obviously condition (II) of Theorem 5.1.1 is fulfilled. Our next theorem combined with Theorem 5.1.2 will now enable us to get also the condition (I) of Theorem 5.1.1 fulfilled. We first make A growth function F is a function that for every n and every 3n-tuple D1 . . . . . Dn, L1 . . . . . Ln, el . . . . . en gives a positive number F(D1 . . . . . Dn, L1 . . . . . Ln, e l . . . . . e n ) , and every (3n + 2)-tuple D1 . . . . . D n + l , L1 . . . . . L n + l , el . . . . . en gives a lacunarity function f and a positive number 6. We say that the sequence {Dn, Ln, en, Cn } grows faster than F if (1) e~ and Ck are defined in a compatible way from the sequences Dn and Ln for every k. (2) For every n, Dn+l and Ln+l are > F(D1 . . . . , Dn, L1 . . . . . Ln, e l , . . . , en). (3) For every n the lacunarity of en+l /> f and the moduli of the coefficients of en+l are ~< 3 where f and 3 are given by the growth function applied to Dj, L1 . . . . , Dn+l, L1 . . . . . Ln+l, el . . . . . en. DEFINITION.
We will, by slight abuse of language, say that a number depends only on []m thus meaning that it is determined by D1 . . . . . Dm, L1 . . . . . Lm, el . . . . . em, C1 . . . . . Cm. (Obviously different such sequences could give the same ]]m.) We now have
554
P Enflo and V. Lomonosov
THEOREM 5.1.3. There is a growth function F such that if {Dn, Ln, gn, Cn} grows faster than f , then f o r every n there exists Bn depending only on II n-l, such that f o r all N ~ n ~n
Iq - qnl N <
and
16
IhlopN
1
<
Ihl ~
imply
en
Ihq] N >1 Ihl~ Bn
We now combine this theorem with Theorem 5.1.2 to give also (I) of Theorem 5.1.1. The main difficulty in the construction is to prove Theorem 5.1.3. Completion of the construction assuming Theorem 5.1.3.
For every N ~> 0, choose L N + I and Dn+l > max{F(D1 . . . . . Dn,L1 . . . . . Ln, g.1 . . . . . g.n),4Bu+l/eU+l}.
(A)
Now we assume that for every r ~< N we have defined L1 . . . . . Lr, D1 . . . . . Dr, el . . . . . ~r, C1 . . . . . Cr according to (A) and the growth function F.
(A')
Assuming this we will choose ~N+I by the following considerations (B)-(E)" Take the smallest n ~< N 4- 1 such that IqN+l -- qnl N < en/16 and en -- eN+I. Then by Theorem 5.1.3 IhlopN
<
1
EN+I
Ihl ~
==> IhqN+ll N ~ Ihl~ Bn
By the compatibility assumption and (A) we now choose Is247 l1 -- Ln > 4Bn/en,
(B)
Dn > 2.
(C)
CN+I
--
By Theorem 5.1.2 by choosing e N+I lacunary enough we then get [pl N _ ]plN+l
for degree p ~< N.
(D)
We choose ~N+I more lacunary and with smaller moduli of the coefficients than what is given by F(D1, . . ., DN+I, L1 . . . . . LN+I, e l . . . . . ~ N ) .
(E)
By choosing the sequence {Dn, Ln, s Cn } according to (A)-(E), we thus get the following: {Dn, Ln, s Cn } is compatible by (B) and (C) and it grows faster than F by (A), (A I) and (E) so it satisfies the assumptions of Theorem 5.1.3. By (B) and (C) it also satisfies (II) of Theorem 5.1.1 and by (D) it satisfies (I) of Theorem 5.1.1. Thus for the limit norm multiplication by x has only trivial invariant subspaces. So in fact Theorems 5.1.1, 5.1.2 and 5.1.3 give
Some aspects of the invariant subspace problem
555
THEOREM 5.1.4. There is a growth function F such that if {Dn, Ln, en, Cn } grows faster than F, then limll n is a norm for which multiplication by x has only trivial invariant subspaces. The proof of Theorem 5.1.3 uses Theorem 5.1.5, also proved in [ 17,18]. Let A and B be polynomials in many variables. Let Pm A consist of those terms in A of total degree ~< m and Pn B consist of those terms in B of total degree ~< n. We say that a polynomial C has concentration d up to degree m if lPm C ll ~> d lCll where Ill denotes the el-norm, that is the sum of the moduli of the coefficients. We have the following THEOREM 5.1.5. Let A and B be polynomials in many variables with concentrations dl and d2 up to degrees m and n respectively. There exist a constant K = K ( d l , d2, m, n) depending on dl, d2, m and n but not on the number of variables such that IABI~ >1 KIAI~IBI~.
After this result, there are quite a few papers dealing with concentration at low degrees and products of polynomials in many variables. For more results and more references see [3,9,41 ].
5.2. Remarks on some other constructions
In Beauzamy [7] and Read [31 ] one also constructs a norm on the polynomials, such that multiplication by x on the completion of this space is a transitive operator. As part of the definition of the norm one declares certain vectors small and certain operators small. The purpose of having certain operators small is, of course, to pass from a dense set of cyclic vectors to having all vectors cyclic. In [7] (A1)-(A4) above are fulfilled. However, in [7] the e n ' S a r e monomials, e, - - O l n xr(n) where r(n) is rapidly increasing. This makes all non-zero vectors supercyclic. In particular, OlX i
--
1 is small for certain i ls and otis.
(2.1)
In the definition of the norm, Read [31] prescribes (A4). (A2) and (A3) are replaced by "hybrid" bounds, that is bounds on some projections in the space and combinations of multiplication by x and these projections. From that one derives that (2.1) holds also in the example in [31 ], and one constructs bounded en'S with arguments like Lemma 5.2.1 below where the idea above of concentration at low degrees is important. In Section 5.1, [III is defined to be the maximal norm satisfying (A1)-(A4). However, any norm satisfying (A1)-(A4) would work: maximality is not essential. Similarly in [31] minimality is used in the definition of the norm but is in [32] replaced by another condition to get a transitive operator on el. When, in all these examples, certain elements and operators are declared small, it is important that the norm does not collapse to 0. So to make few elements and operators small one uses both above, in [7] and [31 ], and in later examples, rapidly increasing sequences.
556
P. Enflo and V. Lomonosov
In Davie's simplification of Read [32] the following concentration at low degrees result is used. LEMMA 5.2.1. Let r >~ O, 6, 6, M > O. There exists K > O, such that f o r every m, 0 <~ m <. n, every polynomial g ofdegree <. n with IPmg[1 ~ ~ and Igll <~ M there exists q o f degree <<.n such that IPn(gq)
-
and
Zmll < 6
Iqll ~< K.
5.3. More examples o f transitive operators The first examples of transitive operators were strengthened in different ways by Read [3436] and in this section we will give some of the main results of these papers. We say that a vector x is hypercyclic for an operator T on a Banach space B if x has a dense orbit, that is {Tnx I n >~ 0} -- B. The following theorem was proved in [34]. THEOREM 5 . 3 . 1 . If B is s 1 E3 W where W is a separable Banach space, then there is an operator on B f o r which every non-zero vector is hypercyclic. This result strengthens earlier results in two ways: the class of spaces is very big and the operator has a property even stronger than having all non-zero vectors supercyclic. In [35] Read gave the first examples of transitive operators on spaces not containing s To describe this class we first make some definitions. DEFINITION. Let J be the James space, the Banach space of all sequences 4~ - (ai)~=l in Co such that
l[4~ll =
sup Pl
]api+l _ a p i
[2
< 00.
i=1
Put J ~ - - s ~]~iC~=l Ji where the infinite direct sum is the s and each Ji -- J. Now, let C be the class of all Banach spaces B of the form B = Y @ W where Y is Co or J ~ and W is any separable Banach space. Then we have
THEOREM 5.3.2. I f B is in C then there is a transitive operator on B. Recently Read [36] has modified his earlier construction to get a transitive quasinilpotent operator on s
5.4. A transitive operator on s 1 In this section we will describe the transitive operator T on s given in Read [33]. We will also present the recent result by Troitsky [40], showing that there are three operators
Some aspects o f the invariant subspace problem
557
S1, 82 and K (not multiples of the identity) such that T commutes with S1, S1 commutes with $2, $2 commutes with K which is compact. This shows that L o m o n o s o v ' s invariant subspace result on operators, which commute with an operator which commutes with a compact operator, cannot be extended to longer chains. We first define the operator. Let d -- (al, bl, a2, b2 . . . . ) be a strictly increasing sequence of positive integers. Also let ao -- 1, vo -- 0, and Vn = n(an + bn) for n ~> 1. Then there is a unique sequence (ei)~= 0 C F with the following properties: (0) fo -- eo; (A) if integers r, n, and i satisfy 0 < r ~< n, i 6 [0, 13n-4] _qt_ran, we have fi -- a n - r ( e i ei-ran);
(B) if integers r, n, and i satisfy 1 <<,r < n, i ~ (ran + Vn-r, (r + 1)an), (respectively, 1 <~ n, i, ~ ( v n - 1 , an)), then fi -- 2(h-i)/x/-6-~ei, where h -- (r + 1)an (respectively, h - lan); (C) if integers r, n, and i satisfy 1 <~ r <<,n, i ~ Jr(an + bn), nan + r bn], then j~ = ei -- b n e i - b n ;
(D) if integers r, n, and i satisfy 0 <<,r < n, i ~ (nan + rbn, (r + 1)(an + bn)), then fi -- 2 ( h - i ) / ~ e i , where h = (r + 1)bn. i
Indeed, since f/ -- ~ j = 0 )~ijej for each i ~> 0 and ~,ii is always non-zero, this linear relation is invertible. Further, lin{ei" i -- 1 . . . . . n} -- lin{j~" i -- 1 . . . . . n}
for every n ~> 0.
(*)
In particular all ei are linearly independent and also span F. Define T" F --+ F to be the unique linear map such that T e i -- ei+l. Provided d increases sufficiently rapidly, that is an ~ G ( n , ao, bo, a, b . . . . . a n - l , bn-1), bn ~ H ( n , ao, bo, al, bl . . . . . a n - l , bn-1, an),
where G and H are some real-valued functions, this is a continuous transitive operator on s We can assume that all ai and bi, i ~> 1, are even integers. Now we put S1 - - T 2 s o , of course, T and $1 commute. We define S2ei--
ei
i f i is even,
0
otherwise.
Then, by checking the different possible cases we get Szj~ -
{j~ 0
i f i is even, otherwise.
This gives S 1 S z e i -- T 2 S2ei --
e i+2
if i is even,
0
otherwise.
So $2 commutes with $1. Finally define K via K f o -- fo and K f / - - 0 for i ~> 1. Then K is a rank one operator and obviously K $2 - $2 K.
558
P. Enflo and V. Lomonosov
Troitsky also proved (private communication) that T is not the adjoint of another operator, so it does not disprove Conjecture 2.17 above. He also proved [39] that if in the matrix representation of T on (fi) the coefficients are replaced by their moduli, then the new operator ITI has invariant subspaces.
References [1] Y.A. Abramovich, C.D. Aliprantis and O. Burkinshaw, The invariant subspace problem: Some recent advances, Rend. Inst. Mat. Univ. Trieste XXIX (Supplemento) (1998), 3-79. [2] S. Ansari and E Enflo, Extremal vectors and invariant subspaces, Trans. Amer. Math. Soc. 350 (1998), 539-558. [3] R. Aron, B. Beauzamy and E Enflo, Polynomials on many variables: Real vs. complex norms, J. Approx. Theory 74 (2) (1993), 181-198. [4] N. Aronszajn and K.T. Smith, Invariant subspaces of completely continuous operators, Ann. of Math. (2) 60 (1954), 345-350. [5] W.B. Arveson and J. Feldman, A note on invariant subspaces, Michigan Math. J. 15 (1968), 61-64. [6] W.B. Arveson, Ten Lectures on Operator Algebras, CBMS Regional Conference Series in Mathematics 55, Amer. Math. Soc., Providence, RI (1984). [7] B. Beauzamy, Un operateur sans sans-espace invariant non-trivial: simplification de l'example de P. Enflo, Integral Equations and Operator Theory 8 (1985), 314-384. [8] B. Beauzamy, Introduction to Operator Theory and Invariant Subspaces, North-Holland, Amsterdam (1988). [9] B. Beauzamy, E. Bombieri, P. Enflo and H. Montgomery, Products of polynomials in many variables, J. Number Theory 36 (1990), 219-245. [10] A.R. Bernstein and A. Robinson, Solution of an invariant subspace problem of K.T. Smith and P.R. Halmos, Pacific J. Math. 16 (1966), 421-431. [ 11] S.W. Brown, Lomonosov's theorem and essentially normal operators, New Zealand J. Math. 23 (1994), 11-18. [12] S.R. Caradus, W.E. Pfaffenberger and B. Yood, Calkin Algebras and Algebras of Operators on Banach Spaces, Marcel Dekker, New York (1974). [13] L. de Branges, The Stone-Weierstrass theorem, Proc. Amer. Math. Soc. 10 (1959), 822-824. [14] J. Diestel and J.J. Uhl, Vector Measures, Amer. Math. Soc. Math. Surveys 15 (1977). [ 15] R.E. Edwards, Functional Analysis, Holt, Rinehart and Winstone (1965). [ 16] P. Enflo, On the invariant subspace problem for Banach spaces, Seminaire Maurey-Schwarz (1975-1976). [17] P. Enflo, On the invariant subspace problem for Banach spaces, Institute Mittag-Leffler, Report no 9 (1980). [18] P. Enflo, On the invariant subspace problem for Banach spaces, Acta Math. 158 (1987), 213-313. [19] P. Enflo, Extremal vectors for a class of linear operators, Functional Analysis and Economic Theory, Springer-Verlag (1998), 61-64. [20] J. Glimm, A Stone-Weierstrass theorem for C*-algebras, Ann. Math. 72 (1960), 216-244. [21] P.R. Halmos, Invariant subspaces ofpolynomially compact operators, Pacific J. Math. 16 (1966), 433-437. [22] G.L. Litvinov and V.I. Lomonosov, Density theorems in locally convex spaces and applications, Selecta Math. Soviet. 8 (4) (1989), 323-339. [23] VT Lomonosov, Invariant subspaces of the family of operators that commute with a completely continuous operator, Functional. Anal. i Prilozhen. 7 (3) (1973), 55-56. [24] V.I. Lomonosov, Some problems in the theory of invariant subspaces, Kharkov University Dissertation (1974). [25] V.I. Lomonosov, An extension of Burnside's theorem to infinite dimensional spaces, Israel J. Math. 75 (1991), 329-339. [26] V.I. Lomonosov, On real invariant subspaces of bounded operators with compact imaginary part, Proc. Amer. Math. Soc. 115 (3) (1992). [27] C. Pearcy and N. Salinas, An invariant subspace theorem, Michigan Math. J. 20 (1973), 21-31.
Some aspects o f the invariant subspace problem
559
[28] C. Pearcy and A. Shields, A survey of the Lomonosov technique in the theory of invariant subspaces, Amer. Math. Soc. Math. Surveys 13 (1974), 219-230. [291 H. Radjavi and H.P. Rosenthal, Invariant Subspaces, Springer-Verlag (1973). [301 H. Radjavi and H.P. Rosenthal, The invariant subspace problem, Math. Intell. 4 (1) (1982), 33-37. [31] C. Read, A solution to the invariant subspace problem, Bull. London Math. Soc. 16 (1984), 337401. [321 C. Read, A solution to the invariant subspace problem on the space ll, Bull. London Math. Soc. 17 (1985), 305-317. [331 C. Read, A short proof concerning the invariant subspace problem, J. London Math. Soc. 34 (1986), 335348. [341 C. Read, The invariant subspace problem for a class of Banach spaces, 2: Hypercyclic operators, Israel J. Math. 63 (1) (1988), 1-40. 135] C. Read, The invariant subspace problem on some Banach spaces with separable dual, Proc. London Math. Soc. 58 (1989), 583-607. [361 C. Read, Quasi-nilpotent operators and the invariant subspace problem, J. London Math. Soc., to appear. [371 A. Simoni~, An extension of Lomonosov's techniques to non-compact operators, Trans. Amer. Math. Soc. 348 (3) (1996), 975-995. [381 A. Spalsbury, Vectors of minimal norm, Proc. Amer. Math. Soc. 126 (1998), 2735-2745. [391 V. Troitsky, On the modulus of C.J. Read's operator, Positivity 2 (3)(1998), 257-264. [401 V. Troitsky, Lomonosov's Theorem cannot be extended to chains offour operators, Proc. Amer. Math. Soc., to appear. [41] R. Varga, Scientific Computation on Mathematical Problems and Conjectures, CBMS-NSF Regional Conference Series in Applied Math. 60, Soc. for Industrial and Applied Math., Philadelphia (1990). [42] B. Yood, Properties of linear transformations preserved under addition of a completely continuous transformation, Duke Math. J. 18 (1951), 599-612.
This Page Intentionally Left Blank
CHAPTER
14
Special Bases in Function Spaces
T.
Figiel
Institute of Mathematics, Polish Academy of Sciences, ul. Abrahama 18, 81-825 Sopot, Poland E-mail: t.figiel @impan, gda.pl
E
Wojtaszczyk
Institute of Mathematics, Warsaw University, Warsaw, Poland E-mail: przemekw @mimuw, edu.pl
Contents 1. Systems of functions
..............................................
1.1. Systems on [0, 1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. Wavelet bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3. Operations on orthogonal systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Positive results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
563 563 565 568 569
2.1. Unconditionality in L p . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Haar system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Franklin system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
569 573 574
2.4. Walsh system
575
...............................................
2.5. Meyer's wavelets 2.6. Polynomial bases
............................................. .............................................
2.7. Rational bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
576 577 579
3. Negative results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Non-explicit existence results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
580 581
5. Function spaces on compact smooth manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Preliminary definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. The main result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
583 583 585
5.3. Spaces on subsets of the manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
585
5.4. Spaces with boundary conditions 5.5. Decomposition of the manifold
.................................... .....................................
5.6. Decomposition of function spaces on the manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Bases with vector coefficients in spaces of vector-valued functions . . . . . . . . . . . . . . . . . . . . .
586 586 587 588
6.1. Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
588
6.2. Spaces of functions with values in a U M D space . . . . . . . . . . . . . . . . . . . . . . . . . . . .
590
H A N D B O O K OF THE G E O M E T R Y OF B A N A C H SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 561
562
T. Figiel and P. Wojtaszczyk
6.3. Equivalent X-bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Abstract We present several well known orthogonal systems of functions and discuss their properties as bases in selected classical function spaces. We show that in some function spaces there are no bases with additional special properties. We discuss non-explicit methods of constructing Schauder bases. On each compact, smooth manifold we construct a system of smooth functions which is a good basis in a wide range of spaces of Sobolev and Besov type. Finally we discuss systems of scalar valued functions which are bases (with vector coefficients) in some spaces of vector valued functions.
594 595
Special bases in function spaces
563
In this chapter the basis problem is treated from a perspective which is somewhat unusual in the Banach space theory. Given a set S, a linear space ,~' of scalar-valued functions on S and a sequence (fn)n~=l C X , the problem is to describe some (presumably interesting) classes of Banach spaces X, which arise as a completion of &" with respect to a norm I] 9il, in which (fn) forms a Schauder basis or an unconditional Schauder basis. For instance, let I = [0, 1] be the unit interval equipped with the Lebesgue measure. Then the classical Haar system (hn) on I forms a Schauder basis in every Lp(1) for 1 ~< p < cx~ and is unconditional in Lp(I) if 1 < p < cx~. It is also a Schauder basis in the closure of span{hn; n ~> 1 } in the II 9II~. On the other hand, the same Haar system fails miserably if we insist that the sequence should in addition form a Schauder basis in C (I). However, the classical Franklin system satisfies all those properties and in fact it turns out to form a basis in some other interesting function spaces as well. Knowing this, however, does not automatically provide the solution to the analogous question with a different choice of the set S. For instance, if d > 1 and S - Q - [0, 1]d is the unit d-cube with the Lebesgue measure, then to produce a sequence (fn) of functions on S that forms a Schauder basis in C (S) and in L p (Q) for 1 ~< p < which is unconditional for 1 < p < c~ requires some further work even though the Banach spaces Lp(Q) and Lp(I) are isometric for each p and the spaces C(Q) and C(I) are isomorphic. Thus our problem is not concerned with properties of any individual Banach space. It is rather a question pertaining to a family of Banach spaces which have a common dense linear subspace and may perhaps admit sort of simultaneous Schauder bases with certain additional interesting properties. This should give some idea why certain systems of functions have been introduced by various mathematicians and below we shall describe some of the relevant examples.
1. Systems of functions In this section we present the most important systems of functions. Some of them have been defined in other chapters of this Handbook and in particular in [32] but we repeat all the definitions here in order to present them in a unified and systematic fashion.
1.1. Systems on [0, 1] Such systems fall naturally into two categories. Bounded systems are such that the ratio between Ilfn 112 and Ilfn I1~ stays bounded. If this ratio is unbounded we say that the system is unbounded. 1.1.1. Bounded systems. Here we want to describe the most important uniformly bounded orthonormal systems on [0, 1].
Rademacher system.
It is defined as rn(t) -- signsin2nzrt for n = 1, 2 . . . . . It is an orthonormal system, but incomplete. In probabilistic terms rn (t) are independent, identically distributed functions.
T. Figiel and P. Wojtaszczyk
564
Walsh system. For an integer n >~ 1 we write n -- ~ - - 0 O(k) 2k where 69(k) = 0 or 1 are its dyadic digits. It is easy to see and well known that this representation is actually finite and unique. We define the Walsh system on [0, 1], Wn (t), n = 0, 1, 2 . . . . by wo(t) = 1 and for n ~> 1 we put Wn(t) -- 1-I~__0rk+l (t) (O(k). To make it well defined in every point we put 0 ~ = 1. The Walsh system is an orthonormal basis in L2[0, 1]. There is another way to look at the Walsh system. Let us consider the group { - 1 , 1 }r~ with the natural product measure, which is also a Haar measure. Then the system (g0n), where ~On(El, 82 . . . . ) = en, is equimeasurable with the Rademacher system, and the system 7z0 = 1 and for n ~> 1 7rn - I-I~_0 q9~ is equimeasurable with the Walsh system. Note that the system (Trn)n~=0 is the set of characters of the group { - 1 , 1 }rq.
Trigonometric system. The name actually denotes either of two closely related systems of functions. In the complex form we m e a n (eZrrint)n6Z or in the real form we mean the system 1 . . . . . x / 2 c o s 2 z r n t , f f 2 s i n 2 z r n t . . . . with n -- 1, 2 . . . . . The complex form of the trigonometric system can be identified with the set of all characters of the circle group qF. Either form of the trigonometric system is an orthonormal basis in L2 [0, 1]. 1.1.2. Unbounded systems. Those systems generally have a dyadic structure, i.e., they admit a natural ordering isomorphic to a dyadic tree. The dyadic rationals in [0, 1] are linearly ordered as follows. We let to -- 0, tl = 1 and for n > 1 we write n = 2 u + v with # >/0 and 1 ~< v ~< 2~ and we put tn = 2 - ~ - 1 (2v - 1).
(1)
Faber-Schauder system.
We define functions Sn (t) for n = 0, 1, 2 . . . . as follows: so(t) = 1, sl (t) = t and for n > 1 we define Sn (t) as a unique continuous, piecewise linear function on [0, 1] with nodes at points to, tl . . . . . tn such that Sn (tn) = 1 and Sn (tj) -- 0 for j < n. Note that for n > 1 with n -- 2 u + v we have supp Sn = [ t n - 2 - ~ - I , tn + 2-1z-l]. It is easily seen that span{sj}y_ 0 is exactly the space of all continuous, piecewise linear functions on [0, 1] with nodes at points {tj}N=0 . This implies that the Faber-Schauder system is linearly dense in C[0, 1]. It is also clear that if for f ~ C[0, 1] we define Pn ( f ) to be the piecewise-linear, continuous function on [0, 1] with nodes at to . . . . . tn interpolating f at those nodes, i.e., P n ( f ) ( t j ) = f ( t j ) for j - - 0 , 1 . . . . ,n, then P n ( s j ) = sj if j = 0 , 1 . . . . . n and Pn (s j) = 0 if j > n. Thus Pn is a partial sum projection of f with respect to the F a b e r - S c h a u d e r system. Since l[ Pn [l~ = 1 we infer that the Faber-Schauder system is a monotone basis in C[0, 1].
Franklin system.
The Franklin system (fn)n~__o is the G r a m - S c h m i d t orthonormalization of the Faber-Schauder system in the above order. Every continuous piecewise linear func?/ n tion f with nodes at points to, tl . . . . . tn can be written as f - ~ j - 0 a j s j - )--~j=0 bj f j . From this we see that the Franklin system is linearly dense in C[0, 1], so it is an orthonormal basis in L z[0, 1].
Haar system. In our context the easiest way to define the Haar system is to put hn (t) = IlSn~ll2 1Sn~(t) for n = 1, 2 . . . . . Since s n exists except at three points this definition holds
Special bases in function spaces
565
a.e. Clearly, for n = 2, 3 . . . . we have supphn = suppSn = [tn -- 2 - u - 1 , tn + 2 - u - l ] and hn(t) equals 2 u/2 on (tn - 2 - u - l , tn) and equals - 2 u/2 on (tn, tn + 2 - U - l ) . It follows easily from the above that two different Haar functions either have disjoint supports or the support of one (that with the higher number) is contained in an interval where the other is constant. Since f l h n ( t ) d t - 0 for n ~> 2, we infer that the Haar system is orthonormal. It can be seen from our discussion of the F a b e r - S c h a u d e r system (or directly from the definition of the Haar system) that span{hj }~=l equals the space of all step functions on [0, 1] whose discontinuities are only at points t2 . . . . . tn. This immediately shows that the Haar system is linearly dense in L p [0, 1] for 1 ~< p < cxz and that for the orthogonal (partial sum) projection we have the formula /7
P/7(f)--Z(f,
/7
hj)hj-~lljl-l
j=l
j=l
~
f(t)dtllj
(2)
J
in which Ij = [ o / j - l , o/j], where o/0, o/1 . . . . . o//7 is the increasing rearrangement of the set {to, tl . . . . . t/7}. Formula (2) immediately gives THEOREM 1. The Haar system is a monotone basic sequence in every rearrangement invariant function space X on [0, 1]. It is a monotone basis in X if and only if X is separable. 1 Note the following relation between the Walsh and Haar system: for each n -- 1 ' 2,
" " "
span{wj ~2n-| -- span{hj - - -J-"n }~_ 1" J j=0
(3)
Each of the above spaces equals the space of functions constant on all intervals (k2 -n, (k + 1)2 -n) with k = 0, 1 . . . . . 2 n - 1. This observation is the key to the proof that both those systems are linearly dense in L p[O, 1], 1 <<,p < cx~. REMARK. There are generalisations of some systems defined above where the exact dyadic structure is replaced by an "approximate" one (see, e.g., [27]).
1.2. Wavelet bases Recently a new class of orthonormal systems on R and on R d was discovered. We say that a function q/ e L2(R) is a wavelet if the system q J j k ( x ) " - - 2 J / Z t I / ( 2 J x - k) with j, k e Z is an orthonormal basis in L2(•). The system (q/jk)jkcZ is called a wavelet basis. 1 All monotone bases in Lp(#) with 1 ~
Aj+l,2k U Aj+l,2k+l = Ajk and for each j all sets Ajk are disjoint. We put f jk = #(Aj+l,2k)- 11aj+l,Zk_ # ( A j + l , 2 k + l ) -1 1aj+l,Zk+l "
566
T. Figiel and P. Wojtaszczyk
To obtain an analogous orthonormal basis in L2(]R d) with good properties we need 2 d - 1 functions. We say that the system q/s, s = 1 . . . . . 2 d - 1, is a wavelet set on IKd if the system q/js (x) "- 2Jd/Zq/s (2Jx -- k) with j E Z, k ~ Z ' / a n d s = 1 . . . . . 2 d - 1 is an orthonormal basis in L 2 (]l~d). The detailed description how to construct wavelets with particular properties is beyond the scope of this survey. It can be found in [50,19] or [73]. We will describe only some examples.
Haar wavelet.
It is defined as h(t) = 1[0,1/2 ) - 1[1/2,1). It is easy to check that it is a wavelet on R. To generate an analogous wavelet set on IKd let us put h = h 1 and define h ~ = l[0,1]. Then the Haar wavelet set is defined as the collection of 2 d - 1 functions ha(tl . . . . . td) -- hal (tl) . . . . . h ad (td)
(4)
where a ~ E "-- {(al . . . . . act)" as -- 0, 1 and al + . . . + ad > 0}. Haar wavelets defined in this way are not continuous. This can be greatly improved by the following construction.
Meyer's wavelets.
Let us fix a real valued function 69 (~) on ~ satisfying the following
conditions
0 ~< o (~) ~< , /1~ ,
(5)
O ( ~ ) -- O ( - ~ ) ,
o(~)
=
O(~) -
(6)
, / 1~
for I~[ < 27r,
(7)
0
for I~1 > 4zr,
(8)
f o r 0 ~< ~ ~< 27r.
(9t
1
(~92(~) § (~92(~ -- 27r) = ~-y
Let q/0 (t) be the function whose Fourier transform equals 69 (~). First let us observe that for a function f 6 L2(R) the following conditions are equivalent: (a) the system { f (x - n)}ncZ is orthonormal, (b) ~ n ~ Z I f ( ~ + 2krr)l 2 : 1/2zr a.e. This observation follows from the following calculation which uses Plancherel's Theorem:
Z n
an f (x
-
=
n)
2 --
f(~)
d~
~[f(~
§
an ein~ ~
n
E a n e in~ n
d~.
keZ
Since (5)-(9) immediately give ~ 169(~ + 2kzr)l 2 - 1/2zr we see that {q/~ - n)}nzZ is orthonormal in L2(•). Let V0 denote s p a n { q / ~ n)}nzZ. If f 6 V0, then f = ~ n a n ~ O ( 9- n), so f ( ~ ) = (y~nzzanein~)o(~). Thus V0 can be characterized as the space of all f 6 L2(R) such that f ( ~ ) - m ( ~ ) O ( ~ )
for some 2re-periodic, locally L2
Special bases in function spaces
567
function m(~) on R. From this and (7) and (8) we infer that the space V1 := { f E L2(R): f ( 2 . ) E V0} contains the space V0. Let m(~) be a 2Jr-periodic function on • which on [-Jr, 7r] equals x/T~69(2~) and let Z,(~) = ei~/Zm(~/2 + 7r)O(~/2). Clearly, supp,~ C [ - 8zr, - 2 7 r ] tO [~Jr, 3s--Tr].Let (P 1(t) be the function whose Fourier transform equals ,~ (~). Looking at the Fourier transform one can check that the system {q/1 (x - k)}kcz is an orthonormal basis in the space V1 O V0. Note that for any f 6 V1 O V0 with [If ][2 - - 1 the system {2J/Zf(2Jx)}j~z is orthonormal. From this it easily follows that t/-/1 is a wavelet on R. Elementary properties of the Fourier transform yield that q/1 is a real-valued C ~ function satisfying !P 1( _ 89_ x) - ~ 1( _ 89+ x). The above construction leaves a lot of possibilities to choose the function 69. Observe that if 69 is infinitely differentiable (and one can easily construct it in such a way) then also ,~, is C ~ . Since it is also compactly supported we infer that in this case both q/0 1 and q/1 are from the Schwartz class. Another extremum is to put 69(~) -- ~-l[-~r,~r](~) which gives ,~(~) -- ~ 1
ei~/2[l[_2rr _:r] + l[:r,2jr]]. This gives the Shannon wavelet
S ( x ) - (rc(x + 1 ) ) - I [sin27r (x + l ) _ sinn'(x + 1)].
(10)
The way to construct a corresponding wavelet set is analogous to (4). We put q/a(tl . . . . . td) -- q/a, (tl) . . . . . q/a~ (td),
(11)
where a 6 E. This is Meyer's wavelet set on iRd. If 69 is infinitely differentiable then all ~ a ' s are from the Schwartz class. This procedure to pass from one dimensional wavelets to several dimensional systems is standard in wavelet theory (cf. [73, 5.1]). Note that it involves an extra function 4) (in our examples q/0 and 1[0,1] respectively) which is called a scaling function. Its main property is that the system (4)(x - n)),~z is an orthonormal basis in the space span{ tllj,k } j <0,kcZ. The other important possibilities are:
Daubechies wavelets when for a given n = 1, 2 . . . . the wavelet qJ 1 and the scaling function q/0 can be chosen n-times continuously differentiable and with compact support. The wavelet set on iRd is given by procedure (11) and consists of n-times continuously differentiable functions with compact support.
Spline wavelets when for a given n = 1, 2 . . . . the wavelet q/1 and the function q/0 can be chosen in such a way that (a) ~0 and O 1 have (n - 1) continuous derivatives on ItS, (b) for each interval [k, k + 1] C R functions !P~ [ [k, k + 1] and ~pl ][k, k + 1] are polynomials of degree at most n, (c) functions q/0 and q/1 have exponential decay on IR, i.e., ]~ s (t)] ~< Cq Itl for some constants C and q < 1 for s = 0, 1. Properties (a) and (b) mean that !P ~ and q:l are cardinal splines of order n. The wavelet set on R d is once more given by procedure (11). Since a wavelet basis has no natural enumeration (it is indexed by two or three indices) we have to make precise in what order we consider it when we speak of it being a basis
568
T. Figiel and P. Wojtaszczyk
in some function space. When we say that the wavelet basis (q/j~)j,kez on R is a basis in a function space X we basically mean the lexicographic order. More precisely we require that projections P s f "-- Y~j<<.s~ k c z ( f , qQ~)qQk are uniformly bounded and for each f 6 X we have l i m s ~ Ps f - f and l i m s ~ _ ~ Ps f - O, and moreover we require that projections s --
Is, k<~r
are uniformly bounded and for each f E X and r 6 Z we have l i m r ~ _ ~ limr~
Qj f - - 0 r
and
Qj f - Pj f - Pj-1 f . For wavelet bases on ]Rd, (tffjak with a E E, j E Z and
k ~ Z d) we require that projections
a c E j <~s k6Zd
are uniformly bounded and for each f E X we have l i m s ~ Ps f - f and l i m s ~ _ ~ Ps f - - 0 . Moreover we require that for each a E E and j E Z and r E Z d the projections ~l,r Qj f := Z~Td,k<~ r (f, q/jak)qJj.a~ are uniformly bounded and converge appropriately (the inequality k ~< r is understood coordinatewise). Practically the order is not very important since good wavelet bases are unconditional bases. Even if this is not the case then usually for each j 6 Z the system (q/~k)k~Z is an unconditional basic sequence.
1.3. Operations on orthogonal systems
Tensoring. If we have two orthogonal systems (orthogonal bases) (~0n)n= i c~ 0 in LZ(~Qi, #i) with i -- 1 ' 2 then the system q91 | 992 (o91 092) "-q91 (0)1) 9 q 92 (0)2) where n, m - 0 , 1 , ' ' " n m , m is an orthogonal system (orthogonal basis) in L2(~Q1 x S-22, #1 x #2). Also if each (qg/)n~=0 is a basic sequence (basis) in Lp(12i, #i), 1 <<.p <<.cx~, then the system q91 @ ~02 is a basic sequence (basis) in Lp(S-21 x $22, #1 x #2). We need to order this system along the squares, i.e., ( n l , m l ) ~< (nz, m2) when m a x ( n l , m l ) < max(n2, m2) or max(nl, ml) -- max(n2, m2) and nl ~< n2 or max(nl, ml) = max(n2, m2) -- nl -- n2 and m l ~> m2. This fact is true for many other natural norms. For Lp with 1 ~< p < c~ we additionally have that if (~0n)n= i 0 are unconditional basic sequences (unconditional bases) in Lp(S-2i' #i) then (~01 | 992 m )c~ n,m--O is an unconditional basic sequence (unconditional basis) in Lp(S21 x ,(-22, #1 x #2). This follows from Khintchine inequality and (in greater generality) is due to McCarthy [48]. The above procedure to generate bases in several variables from bases in one variable is quite natural. Note, however, that it is different from the procedure we applied for wavelet bases in (11). The main difference between the two procedures is the homogeneity of the resulting two-dimensional basis. The most graphic example is when we look at the Haar wavelet basis on R. The procedure described here gives a basis in L2(R 2) whose supports
Special bases in function spaces
569
are all dyadic rectangles while the procedure (11) gives a basis whose supports are dyadic squares. The arguments from [34] show that those two bases are not equivalent in Lp(R 2) ifp-~2.
Systems of analytic functions. If we have an orthonormal basis (fn)n~__o of real-valued functions in L2[0, 7r] with f0 -- 1/~/-~-, we can use it to construct a complex orthonormal basis in Hz(D) (for definitions see [74]). We proceed as follows. First we define a system (gn)n~=O in Lz[-zr, 7r] by gn(t) "- -~2fn(ltl). It is an orthonormal basis in the subspace of all even functions in L2[-:rt', 7r]. Now we consider (~'n (t))~__l where-denotes the trigonometric conjugation operator;
,y
gn(t) -- p.v. ~
gn(r) cot(t
-
r) dr.
7/"
It follows from well known properties of the conjugation- (cf. [37]) that (gn)n~ 1 is an orthonormal basis in the space of all odd functions in L2[-Tr, ~] and that functions Fn(t) := gn(t) + i~'n(t) are (after identifying [-Tr, zr] with •) in Hz(T), the space of boundary values of analytic H2 (D) functions. Actually one checks easily that the system
{1} {, /is a complex orthonormal basis in H2 (~). This construction was applied to the Haar system by Billard in [3] to construct a basis in H1 (~'). S.V. Bo6kariov applied it to the Franklin system and proved in [4] that the resulting basis is an unconditional basis in Hp (q~), 1 < p < co, and a basis in the disc algebra. He improved this last result in [6] by showing that (fn)n~_-o is a basis in C.
2. Positive results
2.1. Unconditionality in L p In this section we present the sketch of a very general argument that many natural orthonormal systems with a dyadic structure are unconditional bases in L p spaces. What is important is an appropriate decay condition. Let us fix an even, bounded function qg(x) on R, decreasing on (0, oe) such that f o qg(x) ln(x + 1)dx < co. THEOREM 2. Let tp and ~* be functions on R such that the system (tlIjk)j,kEZ is an unconditional basis in L2(IR) with biorthogonal functionals tPj*k. Assume also that IV (x)l, Ig'*(x)l ~< ~o(x) for x E IiL Then (tPjl~)j,k~7~ is an unconditional basis in L p ( ~ ) for 1 < p
T. Figiel and P. Wojtaszczyk
570
Lp(IR) for 1 < p < oc, cf. [71]. The same argument gives also the following: Let (gn)n~=O be an orthonormal system on [0, 1] such that Ign(t)l <~ 2n/2q)(2n(t - t~)) f o r t ~ [0, 1] and n - O , 1 . . . . where points tn are defined in (1). Then the system (gn)n~=O is an unconditional basis in Lp[O, 1] f o r 1 < p < oc. This covers in particular the Haar system and the Franklin system on [0, 1]. The same argument can be applied to polynomial bases discussed in Theorem 11. Before we start with the proof let us explain the structure of unconditional bases in a Hilbert space. Using the Khintchine inequality we easily get the following result originally observed in [38]: THEOREM 3. I f {Xn}neA is an unconditional basis in a Hilbert space H, then the system {Xn /IlXn II Inca is equivalent to the orthonormal basis in H. PROOF OF THEOREM 2 (Sketch). We have to analyse the following operators defined on L2(R):
Qj = I d - Pj with j 6 Z and i < j keZ
Ur
= Z ej~(f, qJj*k)~Pj~, jk
where e (6jk) is a double sequence of signs. The point is to show that operators U+ are of weak type (1; 1) uniformly in e. Then the Marcinkiewicz interpolation theorem and duality give the theorem. For a given f 6 L1 (R) f-) L2(~) and )~ > 0 the CalderonZygmund decomposition (cf. [73, p. 147]) gives a family of disjoint dyadic intervals/rl with Ilrll - 2 -r such that )~ < Ilrl1-1 flrl Ifl ~< 2)~ and If(x)l ~< )~ for x E F " - R \ Ur! lrl. Then we write frl = f " llrl and f = g + h where g = f . 1F + Erl Pr(frl). Thus h = =
~ r l f r l - Pr(frl). We will prove the following two estimates.
E]
FACT 1. I f f E L I ( R ) f ) L2(~) and supp f C Irl then f o r x 6 R such that 12rx - ll > 2 we have [Ue Qr f (x)l <~ C I I f lllZrrl(2rx -- l) where rl(u) -- ~j>~o 2Jqg( 88 FACT 2. II ~rl Pr(frl)ll2 ~< C~/)~llflllPROOF OF FACT 1. To prove Fact 1 note first that by a change of variables it suffices to consider r = 0 and I -- 0. Then
IU Qof(z)l
j>~Olkl<~2J j>~Olkl>2J
If Ixl ~> 2 and Ik[ ~< 2 j then qg(2Jx - k) <~ ~0(2J-lx) so we estimate the first sum as
k) d, Cllfll,
Z j/~O
where we use the fact that ~ k e z r
keZ
2J o(zJ-'x), j/~O
- k) is a bounded function in z.
571
Special bases in f u n c t i o n spaces
To deal with the second sum we use supp f C [0, 1] and estimate as follows
Z Z 2jllfll' sup ~o(2Jt
- k)~o(2Jx - k)
tE[0,1]
j>/O lkl>2J
<~llflil~2J(~j~p(k-2J)~~ j/>0
+ ~_. ~~176
k>2'
9
k<-2J
Observe that one has min{qg(a), q)(b)} <~ qg( 89 + b)). Thus the two preceding sums over k can be estimated by 1)) and respectively. Moreover, if Ixl ~> 2
C9)( 89
Cq)( 89
then 2 j ix - I I ~> 89j [x[. This completes the proof of Fact 1.
[]
Let us put rl(u) -- O(t) for u 6 R, where t -- max{lul, 1 }. Note that 17 is integrable on R, because f2 ~ O(t) dt -- Z
2 j f2 ec q) ( ~ ) 2 jt
dt -- ~I Z L ~
j/>0
j-1 q)(u) du ~< C.
IIPrflloo <~C2rllflli. Using
PROOF OF FACT 2. To prove Fact 2 we first show that change of variables we see that it suffices to consider only r - 0 but then
[Pof(x)l ~
Z2J~
j<<.o
(12)
j/>0
kcT~
the
Llf(t)l (2Jt- )dt (2Jx- )
j~<0 ~< CIIflll ~ 2 2 ~< CIIfll~. j~
[Pr(fr~)(x) I <<. CIIfrzlI~ 2 r r / ( 2 r x - l), and hence
2 r,1
2
~< 2 Z 2 r rl
rl r',l'
2r'llf"zll'llL'"lllfo(2rt-l)r#(2r"-l') dr"
Z rI>lr, F
Since I/r/I - 2 -r, from the definition of get 2 r' [[ [[1 ~< 2k, so
fr'l' I
4X
2rf ~7(2rt-l)tl(2r't-l')dt.
ll/r, llJ rl
frl and the Calderon-Zygmund decomposition we
rr >/r,l I
T. Figiel and P. Wojtaszczyk
572
This leads directly to the required estimate I ~< CXllflll, because we prove below that there is a C depending only on r/such that sup
Z
r,1 r'>r,l'
2 r f rI (2rt -- 1)rI (2r't -- 1') dt ~ C. d
Using the substitution 2rt - 1 = u, we can deduce the latter estimate from the following lemma. Fq 4. There is C < ~ depending only on rl such that, if mutually disjoint dyadic intervals with IIj,kl = 2 - j ~ 1, then LEMMA
Z (j,k)~S
F
(Ijk)(j,k)~S
is any family of
r l ( u ) r l ( 2 J u - k ) d u ~ C.
PROOF OF THE LEMMA. Clearly, it will suffice to produce two integrable functions on R, say 4~0, 4)1 such that, for any dyadic interval Ijk with j ~> 0, if Jjk -- [k2-J - 1, k 2 - J + 1], then
lo - fjj ~ n(u)n(2au - k)du <<.fj~ I, "--
rl(u)rl(2Ju-k)du \Jjk
~bo(u) du,
<~ flj q~l(u)du. k
Observe that if lul ~ 1, then 2~(2u) ~< O(u), since r/(u) - 2r/(2u) = ~0(u/4) >~ 0. Recall that r/is even, nonnegative, integrable and non-increasing on [ - 1 , ~ ) . Now, if u ~ Jjk, then r/(u) ~< r/(Ikl2-J - 1), hence Io ~< 2 - j r/(Ikl2 - j - 1) s
r/(u)du,
while if u 6 R \ Jjk, then 2 j r/(2 j u - k) <~ rl(k2 - j - u), hence 2 j ll <<.f• rl(u)rl(k2 - j - u) du - (rl * r/) (k2-J). It is easy to verify that the convolution rl * r/is an even, integrable function, non-increasing on [0, cx~). We put 4~1(x) = ( n , r/)((Ix I - 1)+), where, as usual, x+ = max{0, x}. Then 4~1 is also even and integrable. Thus the upper estimate for 11 will follow as soon we check that, for x ~ Ijk one has 4)1 (x) ~> (r/ 9 r/)(k2-J). This is obvious if Ixl ~< 1, while if x ~ Jjk and Ixl > 1, then I x l - 1 ~< Ikl2-J and hence 4~1(x)/> (7 * rl)(Ikl2-J). A suitable function 4~0 can be defined by the formula ~b0(x) = cn((Ixl- 2)+), where c = f• r/(u)du. Indeed, if x ~ Jjk and Ixl/> 2, then 0 ~< Ixl - 2 ~< Ikl2-J - 1, hence 4~0(x) >~ co(Ikl2-J - 1). The latter estimate is obvious if Ixl < 2. Consequently, we get fljk C~O(X)dx >~2 - J c o ( I k l 2 - J - 1) ~> I0. This completes the proof of Lemma 4 and thus of estimate Fact 2. D
573
Special bases in function spaces
To complete the proof of Theorem 2 we still need to show the weak (1; 1) estimate for operators U~. This follows from Facts 1 and 2. Clearly
[{x. ]g,f(x)[ >xll <, x. lUng(x) I >-~
+
x"
lUsh(x)] > ~
.
(13)
Since If(x)l ~< )~ for x 6 F we get fF If(x)l 2 dx ~< )~llflll, so from Fact 1 we infer that Ilgl12 ~< C~/)~llfll~. Using this we get
{
x Is
/
g(x)l>5
"J
x Is g(x)l
4
4
C
~< ~_~[[U,g[[2_ ~_~11gl[2 ~< ~-[[fl[1.
(14)
Letus denote A -- ~ r l 10./rZ. Then IAI ~< ~ Ilflll. On the other hand,
f, \A
rl
\A
IU~Q~(fr,>l<<.~f, IU~Qr(frl)'. rl
\Irl
so from Fact 2 we get
]~
IU~hl <~C ~ \A
Ilfrzll ~< CIIflll
rl
and
x: IS~h(x)l > -~
~< IAI-+10
x E R \ A" lUsh(x)[ > 2C
C
~< ~ Ilflll + - ~ Ilfll~ - ~-Ilfll~. From (13) using (14) and (15) we see that operators U~ are uniformly of weak type (1; 1). This completes our sketch of proof of Theorem 2. The above argument for wavelets is presented in [71 ] following ideas of [29]. The extension to biorthogonal wavelets is presented here for the first time. Further extensions to multidimensional biorthogonal wavelets associated with arbitrary dilation are given in [60]. It is a very interesting fact that each wavelet bases in L p (R) (and also any basis to which Theorem 2 applies) is equivalent to the Haar basis. This follows directly from results of [28]. By interpolation we infer that Theorem 2 holds when Lp (R) is replaced by any separable rearrangement invariant function space X on R whose Boyd indices (cf. [32]) satisfy 1 <
Px <, qx < ec.
2.2. Haar system The fact that the Haar system is an unconditional basis in Lp[0, 1] is due to PaleyMarcinkiewicz and is classical 9 The Haar system is not only one of the bases in a rear-
T. Figiel and P. Wojtaszczyk
574
rangement invariant function space on [0, 1] but in a precise sense it is the best. This is expressed in the following two theorems. THEOREM 5. Let X be a separable rearrangement invariant space on [0, 1]. The following conditions are equivalent: (1) The Haar system is an unconditional basis in X. (2) The Boyd indices of X satisfy 1 < p x <. qx < oo. (3) X has an unconditional basis. Even when the Haar system is only a basis in a rearrangement invariant space X (e.g., when X -- L 1[0, 1]), it still has certain universality property. THEOREM 6. Let X be a separable rearrangement invariant space on [0, 1]. Suppose that X is isomorphic to a subspace of a Banach space Y with a basis (yn)nec__1. Then there exists a block basis of the basis (Yn) which is (1 + e) equivalent to the Haar system in X. PROOF. Since X is a rearrangement invariant space, only the joint distribution of the Haar /oo 2J--1 system matters. More precisely, suppose we have a family of subsets {Ajkjj=Ok=O C [0, 1] such that for each j, k one has Ajk -- Aj+l,2k U Aj+l,2k+l, and I A j k l - 2 - j . Then the system of functions fjk "--2J/2(1Aj+l,2k -- 1Aj+l,2k+l) is isometrically equivalent to the Haar system in X. Now if X C Y and (yn)~CC=ois a basis in Y, then we inductively construct sets Ajk as above in such a way that fjk is (almost) a block basic sequence. This is done using the following corollary of the classical Liapounoff convexity theorem (cf. [40]):
for every finite subset V C X* and every measurable subset A C [0, 1] there exists a subset B C A such that IBI = IA \ BI and x*(1B) = x*(1A\B) for x* ~ g. F1 It follows from this theorem that neither L 110, 1] nor (y~n~176 L l+l/n[0, 1])2 embeds into a space with unconditional basis. The above theorems are of fundamental importance in Banach space theory and their detailed proofs and references to the original articles can be found in [43, 2.c].
2.3. Franklin system The fundamental feature of the Franklin system is its exponential decay. If we write fn for an individual Franklin function and Kn (x, y) for the kernel of the partial sum projection, i.e., Kn(x, y)"-- Y~'~=0 f k ( x ) f k ( y ) , SO that
Pn f (x) -
s k=O
then we have:
( f , fl~) f k ( x ) -
fo1Kn(x, y) f (y) dy
575
Special bases in function spaces
THEOREM 7 (Ciesielski). and tn as in (1)
There exist constants C and 0 < q < 1 such that f o r every n ~ 0
(16)
I K , ( x , y)] <~ C(n 4- 1)q nlx-yl,
(17) (18)
]f~(x)] ~ C(n 4- 1)3/2q "lx-"l.
From (16) above we immediately conclude that the Franklin system (fn)n~=O is a Schauder basis in C[0, 1] and in L 110, 1] so by interpolation it is a Schauder basis with uniformly bounded basis constant in all L p[0, 1] for 1 <~ p < cxz. A minor modification of the proof of Theorem 2 gives that the Franklin system is an unconditional basis in L p [0, 1] with 1 < p < e~, a result proved by S.V. Bo6kariov [4]. The smoothness of the Franklin system gives that it is an unconditional basis in Lip ~ [0, 1] for 0 < ot < 1 and that f E Lip ~ iff (f, fn ) -- o((n + 1)-~-1/2). So as a basis of Lip ~ this basis is equivalent to appropriately scaled unit vector basis in co. This also gives a characterisation (and isomorphism with s of Lip~[O, 1]. Namely f ~ Lip~ iff (f, fn) -- O((n + 1)-~-1/2). Also the Franklin system is an unconditional basis in the atomic space Hp[O, 1] for p > 89cf. [68]. The above ranges are optimal what reflects the fact that we have (18) but nothing like this makes sense for higher derivatives. The Franklin system is also an unconditional basis for some range of Besov spaces. For definitions and general information about Besov spaces the reader may consult [22] or [53]. The Franklin system was introduced in [26] and the fundamental Theorem 7 was proved in [13]. Its basic properties and many of the results mentioned here are explained in [36] and [72]. For connections with Hardy spaces for p < 1 and the role of smoothness see [68] or [65]. First results on spline bases in Besov spaces are in [62].
2.4. Walsh system THEOREM 8. The Walsh system is a basis in Lp[0, 1]for 1 < p < e~ and it is not a basis in Ll[0, 1]. PROOF. For f sider the series unconditionally k l < k2 < ... <
- - 2 k+l
Pk,(f) /----1
<<.Cllfllp.
(19)
p
Now for N - 2 k 4- s with s < 2 k we have N
N
Eli, n--0
1
E Lp[O, 1] with 1 < p < e~ put P k ( f ) -- ~_.~n=2S (f, Wn)Wn and con~ - - 1 P k ( f ) . From (3) and Theorem 5 we see that this series converges in L p[O, 1]. This implies that for each increasing sequence of integers ks we have
p
n----0
p
576
T. Figiel and P. Wojtaszczyk
But from the definition of Walsh functions we see that for 2 k ~< n < 2 ~+l we have l/An .rk -Wn_2k and for 0 ~< n < 2 k we have Wn 9rk -- Wn+2k. A moment's reflection gives that we can write 2~+s
Z
( f " rk, Wn. rk)wn, rk -- ~
n=0
Pk, ( f " rk)
(21)
l
for some sequence kl < k2 < ... < klz. From (19), (20) and (21) we get the claim. The fact that the Walsh system is not a basis in L 1[0, 1] follows easily from Theorem 19 below, but can be also seen more directly by estimation of the Dirichlet kernel of the Walsh system, see [63], Theorem 9 of Chapter 1. [-1 THEOREM 9 ([67]). The trigonometric system and the Walsh system are equivalent bases in Lp[O, 1] only for p = 2. The order of the Walsh system is crucial in the proof of this theorem. We conjecture, however, that no permutation of the Walsh system is equivalent to the trigonometric system in L p [0, 1] when p -r 2. This is the case for p = 1. In the trigonometric case we know that for any finite subset A C Z the norm in L1 (qr) of the projection PA f -- ff'~ncA i(rl) eint is at least c In [A[ (cf. [49]). On the other hand we know from (2) and (3) that for the Walsh system there are finite subsets with cardinalities tending to (x~ for which the corresponding projections are uniformly bounded. One of the most important applications of Walsh system to geometry of Banach spaces is the paper [ 11 ] where uncountable family (A~)~
~ 2 nj/2.
(22)
2 Starting from one dimensional wavelets and using formula (11) one gets corresponding system !//;k (x) with a 6 E, j 6 Z and k 6 Z d. It follows from the one dimensional argument that for each multiindex ot E 1%d1 the system 0~ Ox ~ q/~k is a Riesz basis in L2 (Rd).
(23)
Special bases in function spaces
577
The rule of thumb is that those systems are best possible bases in most function spaces. Here are some examples: (a) Let S C I~Id be a smoothness, then the system (6'jak) is a basis in Cos (IRd) (for definition of smoothness and the space Cos (IRd) cf. [57]). This follows immediately from the fact that the orthogonal projection onto span{ q/j.ak" j ~< 0, a E E, k 6 Z d } can be written as PF -- ~,k~ZJ (f, ~0 (x -- k)) ~o (x - k). (b) Let S C l~ld be a smoothness, then the system (t/-ta)j k is an unconditional basis in Lps (iRd) for 1 < p < cx~ (for definition of smoothness and the space L ps (Rd) cf. [57]). From Theorem 2 and procedure (11) we infer that for each multiindex ol the system (23) is an unconditional basis in Lp(I~d), so we get the claim. (c) Meyer's wavelets are also unconditional bases in Besov spaces. For 1 < p < oo it follows from the above and the fact that such Besov spaces are interpolation spaces between appropriate Sobolev spaces (see Theorem 6.4.5 of [2]). This is also true for other values of p, cf. [50] or [73]. (d) Meyer's wavelets are unconditional bases in real-variable Hardy spaces Hp (IRd) for 0 < p < e~ (cf. [50]). 2.6. Polynomial bases The problem here is to find (natural) systems of polynomials which are (unconditional) bases in some important spaces. This problem was investigated for the scale of Lp-spaces in two settings: (1) We look for algebraic polynomials in Lp[-1, 1] or in C [ - 1 , 1]. (2) We look for trigonometric polynomials in Lp(7~) or in C(qP). Let us start with the case (ii). The following result is classical, see [43].
THEOREM 10. The trigonometric system (eint)n6Z, in its natural order given as O, 1, - 1 , 2 , - 2 , 3 . . . . . is a basis in Lp(7~) for 1 < p < cx~. It is an unconditional basis in Lp(TE) only for p - - 2. Exponentials in any order are not a basis in any of the spaces C (qP) and L I(T). This follows from the solution of the Littlewood conjecture [49] (see also Theorem 19). The following, relatively simple, system of trigonometric polynomials, cf. [69], has nice properties. First we define polynomials P0k (t) as follows:
P~
~
e il'
(24)
I/1~1
and for k = 1,2 . . . . we put
P~(t)- (-1) k Z
v/Fk(n)eint + Z
n~
v/Fk(n)eint'
(25)
n>O
where Fk (n) is defined as
Fk(n)--
2-k(2 k+l - I n [ )
for 2 k ~< In] ~ 2 k+l,
2 - k + l ( l n [ - 2 k-l)
for 2 k- 1 ~< In] ~< 2 k,
0
otherwise.
(26)
578
T. Figiel and P. Wojtaszczyk
Now for j = 0, 1 . . . . . 3 92 k- 1 _ 1 and k = 0, 1 . . . . we define P;(t)--P~
t-3.2
27rj ) k-1 "
(27)
We order this system lexicographically and we denote the resulting system by (Pn)n~__o. Then we have THEOREM 1 1. The system (Pn)n~_O has the following properties: (1) it is a complete orthogonal system, (2) deg Pn <<.4n, (3) it is a basis in L1 (qF) and in C (qF), (4) it is an unconditional basis in Lp(qF) for 1 < p < cx~. PROOF. Claim (i) follows from direct computation with Fourier coefficients and (ii) is easily inferred from the definitions. To see (iv) we prove the estimate Ie~ ~<
C2 j min(1, (2 j Itl) -3/2) and then we apply the argument of Theorem 2. To see (iii) we write the orthogonal projection onto span{P;: k ~< r and j = 0 . . . . . 3 . 2 k-1 - 1} as a combination of three Fourier multipliers and check that those multipliers are uniformly bounded. To fill in the remaining gaps we use the above pointwise estimate and conclude that
(P) ~3"2k-1-1 is uniformly in k equivalent (in L1) to the unit vector basis Jj=O
in
~ "2k-j 9
[-7
Analogous systems can be obtained using the periodisation of Meyer's wavelets (for details about the periodisation procedure for wavelets see [50,19,73]). More complicated constructions give systems having the same properties but with (ii) replaced by deg Pn <, ( 89+ e)n (for n big enough) with e > 0 any preassigned number cf. [44]. One cannot take e -- 0 in those estimates. Namely we have THEOREM 12. If a system of trigonometric polynomials (fn)n~176 is a basis in C(7s or in LI(qF) then there exists an e > 0 such that for n big enough we have degfn ~> ( 89+ e)n cf [61 ] and also [69]. If the system (fn)n~=O is an unconditional basis in Lp (~) for some p, 1 < p ~ 2 < c~, then lim sup n -1 deg fn > 89 cf [70]. REMARK. In [70] an unconditional basis in Lp(T), 1 < p < oo is constructed for which max{deg fj" j <, 2 n} - 2 n-1. This basis is actually the periodisation of the Shannon wavelet. Now let us consider the algebraic case. The situation in C [ - 1 , 1] is reduced to the trigonometric case by the classical substitution S ( f ) (0) - f (cos 0). Using this and slightly modifying the constructions in previous theorems we get THEOREM 13. If the system of algebraic polynomials (fn)n~176 is a basis in C [ - 1 , 1] (or in L l [ - 1 , 1]) then there exists an e > 0 such that deg fn >~ (1 + e)n for n big enough. Conversely, given an e > O, we can find a basis (fn)n~176 0 in C [ - 1, 1] (and another basis in L 1 [ - 1, 1]) consisting of algebraic polynomials and such that deg fn ~< (1 + e)n.
Special bases in function spaces
579
In the case of L p norm the above substitution introduces a weight which creates a problem in the translation of the results. Nevertheless it is known that for each p there exists a system of Jacobi polynomials which is a basis in L p[ 1, 1]. Let (fn)n=0 c~ oc be the system of algebraic polynomials with deg fn~ = n for n = 0, 1, 2 , . . . that is orthonormal on [ - 1 , 1] with respect to the measure (1 - x2) ~ dx. This definition requires oe > - 1 . A special case of a theorem of Muckenhoupt [52] can be formulated as follows: -
THEOREM 14. Let - 1 < ol < oo and let 1 < p < oo. The system (fn~)~_o is a basis in L p [ - 1 , 1] if and only if 1
l+~ < p 4
when - l < c~ <~
3+2~ < p < ~
4
whenoe>
1
(28)
2' 1
2"
(29)
For unconditional bases the results are weaker and less explicit. The best published result seems to be the following theorem of Canturija [ 18]. T H E O R E M 15. For any p, 1 < p < oc and any e > 0 there exists an orthogonal system o f algebraic polynomials (fn)n~=o which is an unconditional basis in L p [ - 1, 1] and satisfies
deg fn < n j+e f o r n big enough.
2.7. Rational bases The question of existence of polynomial bases considered in previous section is a particular case of a general (open ended) question of existence of bases whose elements belong to some analytically defined family. A natural question of this type which has been around for some time is: does there exist a basis in C[0, 1] consisting of rational functions of uniformly bounded degree. A very general approach to this and other similar questions has been proposed recently by E Petrushev [58]. It is based on the following perturbation result: Let us take a Meyer's wavelet ~P E L2(]1{) O ,5' and take a function r such that [tP(J)(t) - q:,(J)(t) [ <~ e(a + Itl) -M
(30)
for j = 0, 1. . . . . k and all t 6 IR, where integers k, M and a number e > 0 are properly chosen. Then the system cI)j,k(x) := 2J/ZcI)(2Jx - k) is an unconditional basis in L2(N) with biorthogonal functionals having very good decay. This result leads to the fact that the system {(I)j,k}j,kcZ is a basis (unconditional basis) in many function spaces, just like wavelets. An analogous construction can be performed on the unit interval. It is possible to find a rational function q~(x) satisfying (30) and since translations and dilations do not change the degree of rational function we obtain THEOREM 16 ([58]). There exists a system o f rational functions o f uniformly bounded degree which is a basis in C[0, 1] and an unconditional basis in Lp[O, 1], 1 < p < oc. The above ideas can be also applied for functions on IRn (see [39]).
580
T. Figiel and P Wojtaszczyk
3. Negative results There are natural restrictions on properties of bases a given space may have. We will list several results in this direction. The following is true for classical spaces. THEOREM 17 ([35]). The space Lp[0, 1]for 1 < p < oo, p ~ 2, does not have a subsymmetric basis because every normalized unconditional basis contains a subsequence equivalent to the unit vector basis in g.p. THEOREM 18. I f the space Ll (lz) has a normalized unconditional basis (en)n~=l, then this basis is equivalent to the unit vector basis in ~ 5. In particular lZ has to be a purely atomic measure, so LI[0, 1] does not have an unconditional basis. I f the space C ( K ) has a normalized unconditional basis (en)n~__l, then this basis is equivalent to the unit vector basis in co. In particular C[0, 1] does not have unconditional basis (see [42] or [72]). REMARK. Actually spaces L 110, 1] and C[0, 1] do not embed into spaces with unconditional bases. As we know neither the Walsh system nor the trigonometric system is a basis in L 1 [0, 1]. This is connected with the fact that those systems are uniformly bounded. Note that the unit ball from L ~ is a weakly compact subset in L m[0, 1], and thus does not contain a subsequence equivalent in L 110, 1] to the unit vector basis from el. This shows that the following theorem implies in particular that there is no uniformly bounded, normalized basis in L 1[0, 1]. THEOREM 19 ([66]). Every normalized basis in the space L1 [0, 1] contains a subsequence equivalent to the unit vector basis in g.1. For every normalized basis (fn)n~=l in C[0, 1] there exists a sequence of integers nk such that the map (y~n~=l an fn) ~ (ank) is onto co. This in particular implies that there is neither p-besselian basis in C[0, 1] nor a philbertian basis in L m[0, 1]. Let us recall that a normalized basis (Xn)n~=l in a Banach space X is oo lan ip ) 1/p , - p-besselian (for given p E [1, cx~) ) if II Y~n~=l anXnll ~ C ( Y~n=l - p-hilbertian (for given p E (1, ~ ) ) i f II ~ n ~ l anxnll <~ C(~n~=m lanlP) 1/p. It was proved by J. Bourgain [8] that the disc algebra A(D) has no p-besselian basis. The proofs of Theorem 19 and the result of Bourgain are very complicated and local in nature. To give a flavor of those technical results let us formulate one fact which contains an estimate of the Lebesgue functions and hence yields a lower estimate for norms of the partial sum projections. N 1 PROPOSITION 20 ([66]). Let (S, 13, m) be a probability measure space and let (fi , gi)i= be a biorthogonal system of functions on S such that (1) Ilgill~ ~< l f o r i 1,2 . . . . . N, (2) fs I ~--~/N__,Si fi 12dm <<.C ~-~N_1 Isi 12.
Special bases in function spaces
581
Then there exists a constant C I > 0 (depending only on C but not on N) such that
=xZf
l~k~N
k
Zgi(t)fi(s) i=l
dm(t) dm(s) >~ C' lnN.
The above results about bases in C and L 1 should be compared with the following result of [30] and [31]: Every normalized basis in Lr[O, 1], 1 < r < c~, is p-besselian f o r some p and q-hilbertian f o r some q.
4. Non-explicit existence results It happens sometimes that one can prove the existence of a basis in a space without providing any explicit construction. This happens when we can establish that a space X is isomorphic to a space Y with a basis, using the decomposition method. Another way is to patch up the basis from various finite dimensional pieces. A concrete example of this approach is THEOREM 21 ([ 10]). The ball algebra A (~n) f o r n = 1, 2 , . . . has a basis. Another important result of this type is THEOREM 22 ([33]). Every separable s
1 <~ p <~ ~ has a basis.
For the definition and properties of s the reader should consult [1,41,33,11] and the references given there. Two general results of this type have been proved recently by W. Lusky. They are largely motivated by a general question whether a translation invariant subspace of C (G) or L1 (G), where G is a compact abelian group, has a basis. It should be pointed out that methods of Lusky's papers are far reaching generalisations of some arguments from [33]. In particular the original argument for Theorem 22 starts with showing that separable/2pspaces for 1 <~ p < cx~ satisfy condition (ii) of Theorem 23. THEOREM 23 (Lusky [46]). Let X be a separable Banach space. The following conditions are equivalent: (1) X has a basis; (2) there exist p, 1 <<.p <<.~ , and a sequence of finite rank operators Rn : X ~ X with n -- 1, 2 . . . . and such that (a) l i m n ~ Rn(x) -- x f o r all x E X, (b) Rn Rm = R m i n (n,m) f o r n :/: m, (c) operators Rn+l - Rn factor uniformly through g.p, 1 ~ p ~ cx~. THEOREM 24 (Lusky [47]). (a) Let X C C ( K ) have a basis and C ( K ) be separable. If C ( K ) / X is non-reflexive, then it has a basis. (b) Let X C L1 (lz) be non-reflexive and L1 (lZ) separable. If L1 ( # ) / X has a basis then X has a basis.
582
T. Figiel and P. Wojtaszczyk
We will present the proof of the simpler variant of Theorem 23, which is sufficient in many cases: Suppose that X satisfies (ii) of Theorem 23. Then X @ g.p (respectively X • co if p -- oo) has a basis. Let us denote Qn -- Rn+l - Rn for n -- 1, 2 . . . . and Q0 = R1. Let us fix a factorization Qn - Tn o Sn, n = 0, 1, 2 . . . . with []Tn 11 ~< ~. and []Sn 11 ~ ~ and T n ' X ~
kn g.p
kn
and S n ' g p --+ X. For x ~ Un~=l R n ( X ) we define an element j ( x ) = (Qn(x))n~=O ( ~--]m=o oo Q n ( X ) ) p . The closure of those elements in ( Y~m=o Q n ( X ) ) p we denote by Z. We define bounded operators
S"
ep
~
n=O
p
Z Qn(X) n--0
T"
and
Z On(X) n--0
p
~ p
ep n=0
p
by S(wn) -- (Sn(wn)) and T((xn)n~=O) = (Tn(Rn+2 - Rn-1) (Y]~k=O xk))n~__0~ . Note that the series Y]~k=o need not be convergent, but for the sequence (Xn)n~__0 that is eventually 0 this is not a problem. Then we have (Rn+2 - Rn-1) (Y~&=oXk)~176 __ (Rn+2 -- Rn-1)(,Lk=n_ZXk)'~"n+3 ~ and this easily implies that T is bounded. For the sequence (Wn)n~=0 E ( Y]~n=0 g kn p )p that is eventually 0 we have
TSTS(wn) = T
SnTn(Rn+2 - Rn-1)
Skwk k --O
= r
Qn
Sk wk k--0
--
k=0
Tn(Rn+2 -- Rn-1)
Qs s--0
--"
n --O
T n ( R n + 2 -- R n - 1 )
SkWk k=0
n--0
SktOk k =0
= TS((tOk)C~=o).
(31)
n =0
Thus T S is a projection in ()--]~n=0 ~ g kn p )p, so S is an isomorphism when restricted to Im T S. From (31) we see that I m S T S -- Z, so Z is isomorphic to Im T S, so by a classical result of Petczyfiski [56] Z_ is isomorphic to gp (or co if p - oo). We define operators ~'n (and also Qn = Rn+l - Rn) on Z by Rn((Qk(x))~=o) -- (Q~Rn(x))~= o. Those are uniformly bounded operators. For each n let us fix a finite dimensional subspace Zn C Z such that the Zn's are uniformly complemented in Z and uniformly isomorphic to gpn and such that j Rn+2 (X) C Zn. Let us put G - ( Y']'~n=0Zn)p ~ g p and define projections Pm " X @ G --+ X @ G by
Pm (X, (Zk)k~176
= (R~x + j-~(9.m-~Zm),ZJ . . . . . Zm-~, ( I - - Rm)Zm + jOmX, 0 . . . . ). (32) First note that for Zm - (Q~(xm))~=o we get j-l(Q_.m_lZm) - Qm_l(X m) so j - l Q m _ 1 are uniformly bounded operators. Also j Qm are uniformly bounded, so Pm are uniformly
Special bases in function spaces
583
bounded operators. A straightforward calculation using (32) and properties of operators Rm gives that for k > m one has Pk Pm = Pm and for j < m - 1 one has Pj Pm = Pj. Also p2 = Pm. It is clear that Pm's converge pointwise to identity on X (9 G. Thus the sequence of projections P2m gives an FDD on X (9 G. One also checks that operators P2m+2 -- P2m factor uniformly through gp. Thus (P2m+2 -- P2m)(X (9 G) is uniformly isomorphic to a uniformly complemented subspace of an gpm space, so there exist Fm's such that (P2m+2 -- P2m)(x (9 G) (9 Fm have uniformly bounded basis constants and such that
(t m) m=0
p
So we conclude that X (9 g p has a basis, being isomorphic to X (9 G (9 ( Z m = 0 Fm)p. The proof of Theorem 23 in full generality is more complicated because we have to find Zm's as complemented subspaces of ker Rm+| and take care of various necessary technical details. Theorem 21 is a consequence of Theorem 23. We note first that A (1~d) ~ A (I~d) (9 co because it contains a (necessarily complemented) copy of co, e.g., generated by the functions en(z) = [2-1(1 + (z, ~'n))] kn where ~n E 0I~d and the kn's grow sufficiently fast. The 2n
--2n+l
operators Rn are defined by R n ( f ) - Y~k=0 wk + )__~k=Zn+l( 2 - k2-n)wk where wk are homogeneous polynomials of degree k and f -- }-~c=0 wk. One checks that they satisfy all requirements. 5. Function spaces on compact smooth manifolds
5.1. Preliminary definitions Let M be a compact d-dimensional C~ If .T is a function space on I~d then we want 9C(M) to be the function space on M consisting of those functions that locally belong to U. It is rather easy to define .T(M) if .T = C k, where k ~> 0. We shall describe a construction of U ( M ) which is good in the case of Sobolev spaces, i.e., . T - Wpk, and s , where k ~> 0, s > 0 and 1 <~ p q ~< oc. We shall present Besov spaces, i.e., . T - B p,q results only for those two classes, but our notation reminds of a wider scope of these considerations. For more details we refer to [ 15]. One can show that the space Wpk (M) is isomorphic (as a Banach space) to Wpk ([0, 1]d), so the essential feature of the theorem to be presented is the existence of a sequence of functions that is a simultaneous basis for a wide class of function spaces on M. Let us fix a finite C cc partition of unity on M, say (~i)icI, such that for each i there is an open set Ui C M and a diffeomorphism qti :Ui --+ IRd such that (the closure of) the support of ~ i is contained in Ui. (If M is a manifold with a non-empty boundary OM, then near OM one uses the closed half-space IRd+ instead of 1Rd here and in the sequel.) Clearly, the family (Ui, tI-li)icI is a finite atlas for M. With this choice, for a given function space .T = .T(R d) one may define a corresponding function space on M in the following way. A function f defined on M is said to belong to .T(M) provided that there exist functions gi E a~(R d) such that gi vanishes off a compact
T. Figiel and P. Wojtaszczyk
584
subset of tlxi(Ui) and f = Y~i f i , where filui --gi o tlxi and filM\Ui = 0 . (A natural choice is f / = ~ i f for each i.) One may define a suitable norm on ~-(M) by the formula
Ilgi 117,
Ilfll-- i n f ~ i
where the infimum is extended over all such representations of f . (Let us mention that if is merely known to be the completion of a function space 9 on IRd with respect to a norm, then the construction of 9t-(M) can be carried out analogously. First we define the normed space 9 (M) and then we let 5C(M) be its completion.) Here we are only interested in such spaces 9c for which space ~ ' ( M ) and its norm topology do not depend upon the choice of the partition of unity (~Pi) and the atlas (Ui, Oi). Then the space 5C(M) is a linear topological space, which is isomorphic to a Banach space (provided that ~(IK d) is a Banach space) but it has no intrinsically defined norm. Even if OM is empty, we shall need certain spaces of functions which in some sense vanish near boundary. In this context there is some difference between Sobolev spaces and Besov spaces. If U = C k or U - Wpk where k ~> 0 and 1 ~< p < ec, we let ~ ( M ) denote the closure in 5C(M) of the subspace of smooth functions that vanish in a neighborhood of the boundary. o
Thus if 3 M -- 0 then U ( M ) -- 3t'(M). s (M) where s > 0, 1 ~ p q ~< ec satisfies for each integer m > s The Besov s p a c e B p,q the relation s Bp,q (M) -- (W 0 (M), % m (M))s/m, q . o
The related space
s (M) is defined as the interpolation space Bp,q
o
o
o
s Bp,q(M)-(W 0 (M), Wpm(M)) s/m,q' if 0 < s < m, 1 ~< p, q ~< c~ (choosing another m leads to the same space up to equivalence o
but its range is not closed in of norms). There is a natural map B sp,q (M) ---> Bp,q(M), s Bp,qS(M) for some values of s The result for Besov spaces makes use of the sequence space bPp,q which is defined for any real p and 1 ~ p, q ~< co. A sequence a - (an)n~=l is in bPp,q provided that
IlallbPp,q <
oo,
where
q) l/q I[allbPp,q =
2mP m=0
Jan[p \ n=2m
with the usual modifications if p or q equals cx~. We shall fix a positive smooth measure/z on M, i.e., on each set Ui one has d # - hi dx, where hi is a positive C ~ function and dx is the measure transported from the Lebesgue measure on ~i (Ui) by means of q/i. The measure # will be used when dealing with duality and biorthogonal systems.
Special bases in function spaces
585
5.2. The main result The following result was proved in [16,17,14]. THEOREM 25. Let m be a fixed positive integer Then there is a sequence (fn)n~=l of elements of C m (M) with the following properties: (1) (fn)n~__l is a Schauder basis for each of the spaces Ck(M) and Wkp(M) for O<.k<.mandl <.p<~, (2) (fn) n~= l is an unconditional Schauder basis for each of the spaces Wkp(M),
1 < p < cx~ and the multiplier (n (k-1)/d) is an isomorphism from Wkp(M) onto Wlp (M) for O <~1 <. k <. m and 1 < p < c~, (3) for each numerical sequence a with an = Ofor large n one has
C ' Ila IIb~
~anfn n=l
C Ila IIb~.~, s (M) Bp,q
where p = s / d - 1/p + 1/2 and C < oc depends only on m, s and M, o
(4) there is a (unique) sequence (gn)nOc__1 in C m (M) such that fm f i g j dlz -- 6i,j, (3) the sequence (gn)neC__lsatisfies the analogues of (i), (ii) and (iii) with respect to the
spaces ~k (M) and ~Vkp(M) and 13Sp,q (M). Observe that this theorem establishes an isomorphism of the scale of Besov spaces from 0 to m with a subset of the much simpler scale of the sequence spaces bPp,q. In particular, it gives the unconditionality of the basis (fn)~~ 1 for all p, q (if q = oo, then (fn)~__l s (m)). is unconditional in the closure of the smooth functions in Bp,oc Let us mention that the theorem is stated so that it can be readily extended to include all spaces with - m ~< k ~< m or - m < s < m, provided that one knows the needed facts about duality for Sobolev spaces and for Besov spaces. This is explained in [ 16]. In this section we can only outline the approach used to prove the theorem.
5.3. Spaces on subsets of the manifold Since it is rather easier to construct bases in function spaces on sets that have a product structure, the proof of the theorem is based on a decomposition of the manifold M into d-dimensional cubes. If U is an open subset of M, we let ~-(M, U) ___U ( M ) denote the subspace of those f 6 U ( M ) that vanish on U. We need to study some closed subsets C of M for which there is a reasonable concept of f ' ( C ) . Those sets will not be too wild. In particular, the set C will be the closure of its interior IntM (C), the subspace F ( M , IntM(C)) will be the range of a continuous linear projection V in U ( M ) and also # ( C \ IntM (C)) = 0. The notation 5c(C) will then be used to denote the quotient space U ( M ) / U ( M , IntM (C)). If V is a continuous linear projection of t - ( M ) onto t - ( M , IntM(C)) and S = I d - V is the complementary projection, then S induces an isomorphism of U ( C ) onto the subspace S ( U ( M ) ) and we
T. Figiel and P. Wojtaszczyk
586
have ~-(M) = S(.T(M)) G f ( M , IntM(C)). Our decomposition of f ( M ) makes use of some suitably chosen subsets Ci c M and corresponding projections Si. It works because if C ~ C C is another such closed set and V ~ is a projection corresponding to C f and V V t = V~V, then V ~ - V --- SV ~ is a linear projection whose range is in a natural way isomorphic to the space f ( C , IntM(C~)) (here we extend slightly an earlier definition). In this way the space 9C(M) can be decomposed as a finite direct sum of such summands. The latter summands can be further classified using the concept of f ( Q ) z spaces.
5.4. Spaces with boundary conditions If C C 1Rd is a compact parallelepiped, then the previous construction can be applied in order to define f ( C ) . We need some related spaces with various "boundary conditions". Let Q = [0, 1]d and let Z c Q be the union of a family of (d - 1)-dimensional faces of Q. The function space f ( Q ) z is defined in the following way. We let ai = - 1 if {x ~ Q" Xi = 0} C Z and ai -- 0 otherwise. We let bi = 2 if {x 6 Q" xi = 1} c Z and bi -- 1 otherwise. Consider the parallelepiped Q - [al, bl] x . . . x [ad, bd] and the space f ( Q ) . If f is a function on Q such that there is g E f ( Q ) that vanishes in the interior of Q \ Q and whose restriction to Q equals f , then we say that f ~ f ( Q ) z and we let
IlfllT(a)z = IlgllT(0)Using this definition of Wkp(Q)z and BSp,q(Q)z, it can be easily verified that the classical interpolation formula (Wkp, W~ )O,q = Bp,q s , where s -- (1 - 0)k + 01, can be extended to the spaces with subscript Z. Given f , there are 89 + 2)(d + 1) various types of spaces of the form ~ ( Q ) z . Since the sets Q \ z are products of intervals, the construction of a basis in ~ ( Q ) z is facilitated. On the other hand, it turns out that the space U ( M ) is in a rather nice way isomorphic to the direct sum of finitely many subspaces which are isomorphic to ~ ( Q ) z . The isomorphism will be essentially the same operator for a wide range of .T"s.
5.5. Decomposition of the manifold A subset Q~ c M is said to be a d-cube in M if there is a diffeomorphism q~" R d --+ M such that 4~ ([0, 1]4) _ Q,. We restate Theorem 3.3 in [16] as the following lemma. LEMMA 26. There exists a covering of M by a sequence of d-cubes Q 1 , . . . , QN such that for each i, if Ci - [_jij---1 Q j and ~ i is a diffeomorphism such that qsi ([0, 1] d) = Qi,
then Ci has a Lipschitz boundary and the set Zi - cI)[-1 (Qi f-) Ci-1) is the union of a family of (d - 1)-dimensional faces of [0, 1]d. Moreover, if M is a submanifold of a d-manifold M I without boundary, then the set Ci (-J M' \ M has a Lipschitz boundary and the set Z? = ~[-1 (Qi 71 (OM U Ci-1)) is the union of a family of (d - 1)-dimensional faces of[O, 1] d.
Special bases in function spaces
587
In the general case the construction of the decomposition of M into d-cubes uses some basic facts from Morse theory. In concrete cases much simpler constructions are possible, for instance the Euclidean ball in IRd can be decomposed into 2d + 1 d-cubes, while its boundary, the sphere S d- 1 C IRd can be decomposed into 2d (d - 1)-cubes. LEMMA 27. Let Co . . . . . CN be the sets constructed in Lemma 26 and let m ~ 1. Then there exist commuting linear projections Vo . . . . . VN in C m (M) such that, f o r each i = 0 . . . . . N, one has (1) V i ( C m ( M ) ) -- c m ( m , IntM(Ci)), (2) the operator Vi extends to a bounded linear operator in C k (M) and in Wkp (M) f o r each 0 <<.k ~ m and 1 <. p < ec, (3) if f ~ C m ( M ) vanishes in a neighborhood of OM, then so does Vi f , hence the
operator Vi extends to a bounded linear operator in ~k (M) and I~kp (M) f o r each O <. k <~ m and l <~ p < ec, Moreover, (i), (ii), (iii) remain true if the operator Vi is replaced by Vi*, its adjoint operator in the sense of the Hilbert space L2 (lZ). This lemma is a consequence of Proposition 4.3 in [16].
5.6. Decomposition of function spaces on the manifold Let us define the operators R 1 . . . . . RN which map measurable functions defined on M into functions defined on the unit cube Q = [0, 1]d in IRd, according to the formula
R i f -- ( f l Q i ) o (4~i [Q). Observe that the operators V0 . . . . . VN satisfy the relation Vi Vj = g m a x { i , j } , and hence the differences Pi - Vi-1 - Vi are projections such that Pi Pj = 0 for i :/= j and y ~ U 1 Pi -- Id. Given a set Z, we let Z ~ denote the union of those (d - 1)-dimensional faces of Q that are not contained in Z. The following lemma is a version of L e m m a 4.14 in [16]. LEMMA 28. If ~ -- Wkp, where 1 <~ p <. ec and 0 <. k <. m, then f o r each i -- 1 . . . . . n the operator Ri induces (1) an isomorphism of Pi(.U(M)) onto U ( Q ) z / ,
(2) an isomorphism of Pi ( ~ ( m ) ) onto f ' ( Q ) z o , (3) an isomorphism of P*(.U(M)) onto U(Q)zo,, i
(4) an isomorphism of P i * ( ~ ( M ) ) onto f'(Q)z~Let us mention that the operators in U ( M ) constructed in L e m m a 27 are of a special form which guarantees that they are continuous operators in the space L o ( M ) of measurable functions and so are their adjoint operators. Thus we can speak of one well defined operator which acts as an isomorphism in various function spaces. Combining the facts we have mentioned so far with the duality relations for Sobolev spaces, we arrive at a variant of Theorem 4.9 in [ 16].
588
T. Figiel and P Wojtaszczyk
PROPOSITION 29. For each integer m ~ 1 there is a linear operator which induces an isomorphism between the space f (M) and y~N 1G,T'(Q)zi where ~ denotes C K, W~ or Bp,qS for 1 <~ p, q <, c~, - m <~k <~ m and - m < s < m. Namely, the operator f
(Ri Pi f ) N 1 has this property. If OM is non-empty, then this operator induces also isomorphisms between Jr(M) and Y~N 1 G f ' ( Q ) z ? for the same f ' s as before. In this way the construction of Schauder bases with nice properties for Sobolev and Besov spaces on smooth compact manifolds has been reduced for each given 9t" to the analogous problem for a finite number of spaces f ( Q ) z , corresponding to all choices of Z. Once this is done, the bases with the same nice properties can be produced for 9t'(M), for any smooth compact d-dimensional manifold M. (The only restriction is that the number m must be specified in advance.) The main point is to construct appropriate spline systems on [0, 1] and then we use procedure 11. The proof that those systems satisfy the properties listed in the theorem is rather involved. Among other ideas it uses the properties of approximation by vector valued splines in the context of L p ( I d) -- Lp(I; Lp(Id-1)).
6. Bases with vector coefficients in spaces of vector-valued functions 6.1. Preliminaries It is well known that some function spaces have vector-valued analogues. If X is a Banach space and K is a compact Hausdorff space then the space C(K; X), defined in [32], is the X-valued analogue of the space C ( K ) . Similarly, the space L p ( # , X) is the X-valued analogue of L p (#). In a number of cases, for a given function space E and a Banach space X one can give a reasonable definition of a norm on the algebraic tensor product E | X so that the completion of E | X in that norm can be regarded as the X-valued analogue E (X) of the space E. If elements of E are (represented as) scalar functions on a set S, then the space E | X is in a natural way isomorphic to the linear space G of X-valued functions on S spanned by the functions of the form f x , where f E E, x E X. We shall consider only such cases where the construction of E ( X ) is regular in the following sense. If T :X ~ Y is a bounded linear operator, then the composition operator C T : E ( X ) --+ E ( Y ) , C T f = T o f , satisfies IICTII -- IITII and if T is a linear isometry of X into Y, then Cv is an isometry of E (X) into E (Y). The regularity assumption implies that, if F E E (X) and the range of F is contained in a finite dimensional subspace of X, then F 6 G. It is easy to see that the construction of C (K; X) and of L p (#, X) is regular in the sense we have just described. Let us introduce a natural extension of the concept of Schauder basis to this context. In the sequel we shall assume that dim X > 0 (the case X = 0 often requires a special treatment). DEFINITION 30. A sequence (fn)nC~=l in E is said to be a Schauder basis with vector coefficients for E (X), or simply an X-basis for E(X), provided that for each F E E ( X ) there is a unique sequence ( X n ) ~ 1 in X such that F -- Y~n Xn fn, the series being convergent in
589
Special bases in function spaces
the norm topology of E (X). The sequence (fn) is said to be an unconditional X-basis if in addition the latter series converges unconditionally in E (X) for each F 6 E (X). The X-basis constant bc((fn), E, X) and the unconditional X-basis constant ubc((fn), E, X) are defined as in the scalar-valued case. The classical argument can be used to show that bc((fn), E, X) (respectively ubc((fn), E, X)) is finite if and only if (fn) is an X-basis (respectively unconditional X-basis). The regularity condition implies that, if (fn) is a Y-basis in E (Y) and X is a Banach space finitely representable in Y (in particular, if X is a closed subspace of Y), then bc((fn), E, X) ~< bc((fn), E, Y),
ubc((fn), E, X) ~< ubc((fn), E, Y).
Given a Schauder basis (fn) in E, one may ask for which Banach spaces X is (fn) an X-basis or an unconditional X-basis. In this section we shall provide some partial answers. If ( f , ) is an X-basis for E(X), then there is a sequence of linear operators ~n "E(X)--+ X, defined for each F ~ E ( X ) by the formula F - ~_~n~ ~ n ( F ) f n . The 4~n'S are continuous linear operators, because bc((f/7), E, X) < ec. Let (4)n),~ 1 C E* be the sequence of biorthogonal functionals of the basis (fn). Then, for each n, the operator 4)/7 is a unique continuous extension of the operator 4~n | Idx : E | X --+ X that maps f | x to 49n( f ) x . Let Sn -- Z i % l ~bi | fi be the nth partial sum operator with respect to the basis (f/7) in E. The analogous partial sum for the X-basis ( f , ) in E ( X ) is nothing else than the unique continuous extension of the linear map S, N Idx : E N X ~ E | X. This is easy to verify for elements F e E (X) of the form F = f N x. Hence the operator S, N Idx equals the restriction to E | X of the continuous linear operator F P-> ~i%1 (J~i( f ) j ~ . This leads to the formula bc((fn), E, X) --
upll s,
| Idx "E | X --+ E |
xll
(33)
/7
Analogously, we obtain the formula for the unconditional basis constant which involves estimates for the operators Pa -- ~-~i~1 aiOi | fi ubc((fn), E, X) -- sup n
sup a61 n"
1[Pa | Idx "E | X --+ E | X I .
(34)
Ilall ~< 1
Thus in order to show that a given basis for E is an X-basis (or an unconditional X-basis), one needs to verify that there is C < oc so that
It
| IdxE
@ x --+ E @ XI] ~< C I I T ' E --+ Ell,
for all operators T which appear in the above formulae. There are two cases where this estimate holds for any T with C = 1. LEMMA 3 1. Let E = C(K) where K is a compactHausdorff space and let X be a Banach space. If T : E --+ E is a bounded linear operator, then
IT |
X ) - + C(K; X) u ~< IIT'C(K)---> C ( K ) II.
590
T. Figiel and P. Wojtaszczyk
LEMMA 32. Let E = Lp(#), where 1 ~ p < cx~. Let X be a Banach space isometric to a subspace of a quotient of an Lp space. If T : E --+ E is a bounded linear operator, then [IT | Idx" Lp (#, X) ~ Lp (#, X)1[ ~< 1[T" E --> E []. Observe that if p = 1 then X in Lemma 32 may be any Banach space. In particular, if either E is an L1 space or E is a C(K) space, then any Schauder basis (fn) in E is automatically an X-basis in E (X) for each Banach space X. Moreover, if K denotes the scalar field and dim X > 0, then bc((fn), E, X) -- bc((fn), E, K),
ubc((fn), E, X) -- ubc((fn), E, K).
If 1 < p < c~, then the situation is less simple. If X is C-isomorphic to a subspace of a quotient of some L p space, then all bases (fn) in E satisfy bc((fn), E, X) <~Cbc((fn), E, K),
ubc((fn), E, X) <~Cubc((fn), E, K).
In the general case not every basis for E is also an X-basis, but some bases are X-bases for each X. Let ,k be the Lebesgue measure on / = [0, 1] and let (fn) be a sequence in C(I) that is a basis for C(I) and L1 ()~). (For instance, the Franklin system has this property.) If X is a Banach space, then (fn) is an X-basis for C ( I ; X ) and for L I(A, X). Using an interpolation theorem, we obtain that for each 1 < p < oo one has bc((fn), L p, X) < oo, i.e., (fn) is an X-basis for L p (,k, X) for each 1 ~< p < e~. The situation is different for other natural bases in Lp()~), where p 6 (1, oo), and this leads us to the class of UMD spaces.
6.2. Spaces of functions with values in a UMD space The class of UMD spaces, introduced and studied by D. Burkholder and discussed in his article [12], is in our context the class of those X such that the Haar system (hn) is an unconditional X-basis in Lp()~, X) for some p 6 (1, oo). Let us remark that if there is any sequence (fn) in Lp()~) that is an unconditional X-basis for Lp()~, X), then so is the Haar system. Moreover, one has ubc((fn),Lp()~),X) >/ ubc((hn), Lp()O, X). This is a consequence of Theorem 6. On the other hand, the result of J. Bourgain [7] implies that the trigonometric system is an X-basis for L p ()~, X) if and only if X is a UMD space. The trigonometric system is an unconditional X-basis for Lp()~, X) if and only if p -- 2 and X is isomorphic to a Hilbert space [21 ]. The result is actually more general, in particular the same is true for the Walsh system. The proofs in [21 ] use methods of local theory of Banach spaces and operator ideals. The connection between operator ideals and X-bases is treated in [59]. If 1 ~< p < oo, then each finite rank operator T acting in Lp(#) can be represented by a measurable kernel K = K(s, t) so that T f ( s ) -- f K(s, t ) f ( t ) d # ( t ) . The kernel K is of the form y]~j c~j(s)~j(t) with ckj ~ Lp(#) and ~pj ~ Lp,(#). We shall state a result which allows one to estimate the norms of the operators that appear in (33) and (34) in terms of their kernels. We do this only in the case of Lp()~),
Special bases in function spaces
591
where 2. is the Lebesgue measure on I -- [0, 1], so as to make its statement easier. The square 12 is equipped with the product measure 2. | 2.. Let us recall the formula for the coefficients of an integrable function K on 12 with respect to the system of Haar wavelets on the square I 2 defined as in (4). If Q - J1 x J2 c I x I is a dyadic square, there are 22 - 1 -- 3 Haar functions whose support is Q. Let us relabel them as h~ 'j), h ~ 'j), h (j'~ so that we have the following formula for the Haar coefficients of K
(K, h~) -
f
K h ~ d2. Q 2. =
f
K(t,, te)h (a')(tl)h~2)(t2)dtl g, at2
where h (gl) _ h j, h 7) = Ih g l, and h j denotes the Haar function whose support is J. Let IJI denote the length of the dyadic interval J. Observe that every dyadic square J x J~ in the plane is of the form J x (J + n lJ l) for some n c Z. For each non-zero a ~ {0, 1}2 and n E Z and j ~ 1~, we define functionals (with values in [0, oo]) on the set of integrable functions on 12 by the formulae ]K[ a =sup{l(e,h~g+nlgl)•
(J + n l J ] ) x J c i2},
]Kln,j--sup{l(K, hj, Q h j ) I" JZC__J+nlJlC__I, ]JZl--2-JlJ]},
(35)
where (h j, @ h j ) ( x , y) -- h j , ( x ) h j ( y ) . If K is integrable on 12 we define the following two functions
K,(t)=f
K(t, t2)dt2,
K2(t)-f
K(t,,t)dtl.
Finally, let us recall a definition of the dyadic BMO space BMOd(I). A is said to have a bounded mean oscillation on dyadic intervals, if there for each dyadic interval J c_ I there is a constant c such that f j I f ( t ) denote by If]BMOO(I) the least C with this property. (This is defines space BMOd(I), because constant functions have oscillation 0.) THEOREM 33 ([25]).
(36) function f 6 L1 (2.) is C < ~ such that - cl dt ~< CIJI. We a seminorm on the
Let K be an integrable kernel on 12 such that
IKIcz- IKIr ''') + ~ ~(n)(IKIr ''') + IKIr ~ + IKI~''~ < ~ , n-C-0
where ~b(n) = log(1 + Inl). If the integral operatorwith kernel K is bounded in Lp(2.) for some p ~ (1, oo), then the functions K1, K2 defined by (36) belong to BMOd(I). Conversely, if X is a UMD space and 1 < p < oo, then there is C = C(X, p) < oo such that if K is an integrable function on I 2, then the norm of the integral operator TK : L p (2. , X) --+ L p (2., X) defined by K satisfies IITK II ~
C(IKIcz
+ IK11BMOd(I) + IK21BMOdCZ) + I(K, 1)l).
The scalar case of this theorem is a dyadic analogue of the David-Journ6 Theorem [20]. Let us mention that the methods used in [20] do not apply in the vector-valued case.
592
T. Figiel and P. Wojtaszczyk
In [25] part of the proof of Theorem 33 is fully presented only in the scalar case. It is indicated at the end of that paper that the extension to the X-valued case is possible: "It depends on obtaining a new intermediate estimate for P(g, f ) " . The estimate and its proof is due to Jean Bourgain (1987). It appears below as (37). Its proof makes use of the following lemma from [9]. LEMMA 34. Let X be a UMD Banach space and let 1 < p < oc. Then there is C < oc so that, for any probability space (Y2, .T, P), if ~1 ~ .T2 c ... c ~ n ~ .T is an increasing sequence of a-fields and E1 . . . . . En is a sequence of n independent Rademacher variables independent of .Tn then, for any n-tuple f| . . . . . fn ~ L p ( X ) of X-valued Bochner p-integrable random variables on Y2, one has
E~ ~
r
j I .Tj)
<~ CE~ ~-~ Ej f j
j
P
j"
. p
Since [9] does not contain the proof of this lemma we will give one here 9 PROOF. Let us consider an auxiliary increasing sequence of o--fields
and functions F, Fj defined by the formulae
F -- F(~, t) -- Z
~J f j ( t ) ,
Fj -- E ( F I~j) -- ~
j
r
I .T'j).
i<~j
Note that the differences Fj - Fj_ 1 form a martingale difference sequence and that they can be written in the form
j I f ' j ) + Y~ Ei (E(fi I .T'j) - E(j~ I .T'j-,)).
Fj - Fj-1 -- r
i<j
For brevity, we put gi,j --- E(fi I -fi'j) -- E(f/ I -)c'j-1). Now let 6j, j = 1 . . . . . n, be another independent sequence of Rademacher variables, independent of the r and the .T/'s. By the unconditionality of martingale differences, II ~# ~j (fj - f j - 1 ) l l p <~ CllFllp for each choice of •j E {-- 1, 1}. The same is true for the averages, hence
CIIFIIp >1 t ~ e ~ j ( ~ j ( F j -
Fj-1)
J
=
Ave 8
Z
8jEjE(fj If'j)
j
+ Y~ ~ j
i<j
8jr gi,j p
Now we do the substitution Ej ~ (~jEj. This amounts to a measure preserving transformation in the (E, 6)-product space, hence it preserves the value of our expression. We obtain the following Ave 8
Z j
r
I.Tj)
+Z Z j
i<j
Sj~iEigi,j
p
593
Special bases in function spaces
Since all the products (~i6j have mean value 0, from the convexity of the norm it follows that the latter expression is greater than or equal to II Ej ~jE(fj I ~j)[I p- This completes the proof of Lemma 34. [] Now let E j denote the conditional expectation operator with respect to the a-field f j generated by the dyadic subintervals of [0, 1] of length 2 - j and let A j -- Ej -- E j - 1 . Following [25] we let P(g, f) -- ~ j A j ( g ) E j _ I f , where g is an integrable function on [0, 1] and f is a bounded X-valued measurable function on [0, 1]. Then for each 1 < p < ec one has
IIP(g,
f)I1,
C(ilfli ollglip + ilfllpsupllAjgll ), J
(37)
where C < ec depends only on p and on the UMD constant of X. The proof of (37) makes use of the contraction principle for the Rademacher series with vector coefficients. Notice first that the series in the expression for P (g, f ) is a martingale difference sequence with respect to the sequence of ~r-fields ( f j ) . It follows that for any choice of ~j E {--1, 1} we have
HP(g , f)llp
<~C E 6 j ( E j _ l f ) A j g
.
j
Since
E j _ , f - E j f - Ajf,
P
we can write for any choice of 6j 6 {-1, 1}
Z 6j(Ej_, f)Ajg - E 6j(Ej f)Ajg - Z 6j(Aj f)Ajg. J
J
J
Now, taking the average with respect to all choices of (6j) and using the triangle inequality, we obtain the estimate
UP(g,f)l[p<~C(Ave E 6 j ( E j f ) A j g j
E3j(Ajf)Ajg
+Ave P
j
). P
Thanks to Lemma 34, we can estimate the first summand on the right-hand side as follows <~ Ave
Z 6 j E j ( f Ajg)
~< CAve p
J
E6jf(Ajg) J
p
C llf lloollg llp .
The last step involved a standard application of the Fubini theorem and the contraction principle for the Rademacher series. The second summand is even easier to estimate. Using now the contraction principle with respect to the Aj (g)'s rather than to f , we estimate that summand by ~< (sup IlAjglloo)A~e J
E6j(Ajf) J
<~ C ii f ]ip sup ]]Ajgl]~.
p
Combining those estimates we obtain inequality (37).
J
594
T. Figiel and P. Wojtaszczyk
In the final subsection we will present a result from [24]. Here we state the following result which is a direct consequence of Proposition 1 from [24] and of (37) (cf. Remark 1 in [24]). Actually Example 36 needs only Proposition 1 from [24]. THEOREM 35. Let K be an integrable kernel on 12 such that IKIfz,-
IKl~0l'l) n-riO
j/~O
where ~b(n) -- log(1 + Inl). If the integral operator with kernel K is bounded in LpO0 for some p ~ (1, cx~), then the function K2 defined by (36) belongs to BMOd(1). Conversely, if X is a UMD space and 1 < p < cx~, then there is C - C(X, p) < cx~ such that if K is an integrable function on 12, then the norm of the integral operator TK'LpO~, X) --+ L p 0~, X) defined by K satisfies IITK II ~ c ( I g l f z , + Ig2lBMOd(1) -q--[(g, 1)1). REMARK. The integrability assumption on K can be weakened. It is only important that the Haar coefficients of K are defined (perhaps by singular integrals) and so are K1 and K2. Analogous result can easily be deduced for L p o n the real line. REMARK. The above theorems can be extended to more general measure spaces, in particular R, R d and I d. Also, as indicated in [24], the analogous results are true with the Haar system replaced by other natural bases which have similar structure as the Haar system.
6.3. Equivalent X-bases Let us present some results of [24]. For the sake of simplicity, we shall do this only on the unit interval and on the real line. Recall that if (en) C E, (fn) C F are Schauder bases in Banach spaces E, F, then (en), (fn) are equivalent if and only if there is a linear isomorphism T : E --+ F such that Ten = fn for each n = 1, 2 . . . . . More generally, if X is a Banach space and (en) C E, (respectively (fn) C F) is an X-basis for E ( X ) (respectively for F ( X ) ) , we say that (en) and (fn) are equivalent X-bases if the operator T | Idx extends to a linear isomorphism of E (X) onto F (X). EXAMPLE 36. Let T be the operator that maps the Haar functions to the corresponding Franklin functions (normalized in L2). Then T is a unitary map in L2 so that T* -- T - l . Suppose we know that IIT'Lp --+ L p l [ - Cp < cx~, for 1 < p < cx~. Then we obtain [[T*'Lp --+ Lpl[ -- Cp, < c~ where p1 _ p / ( p _ 1). This proves that T is an isomorphism of Lp for 1 < p < cx~. Analogously, if we know that T is bounded in L p ( X ) for every U M D space X and each 1 < p < ~ , then passing to the dual operator we obtain that also T -1 is bounded in L p ( X ) (recall that the dual of a U M D space is UMD, too). This will prove that the Franklin system in L p ( X ) is equivalent to the Haar system in L p ( X ) , hence it is an unconditional X-basis.
Special bases in function spaces
595
The required estimate for IIT'Lp(X, X) ~ L p()~, X)I I can be obtained from Theon rem 35, which can be applied to each K(s, t) of the form ~ j = l hj | fj. To estimate (K, h j, | h j) in (35) we use Ciesielski's estimates (17), (18). This argument can be applied to some other natural orthogonal systems as well, like wavelets or spline systems. A simple one-sided estimate, like the one above, can be dualized so as to provide the estimate going in the opposite direction and hence the equivalence between two systems. It can also be adapted to the setting of biorthogonal systems. Let us mention that using rescaling it is easy to deduce from the above results analogous results about Lp(#) spaces, where # is not a probability measure. It is unknown for what function spaces E and Banach spaces X the Haar and Franklin systems are equivalent X-bases in E (X). Let us mention that for E = L 1 and X = IR those systems are not equivalent [64]. It is known (cf. [54]) that the Haar and Franklin system are equivalent in a rearrangement invariant space X on [0, 1 ] if and only if the Boyd indices satisfy 1 < p x <~ q x < oo.
References [1] D. Alspach and E. Odell, Lp spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 123-159. [2] J. Bergh and J. L6fstr6m, Interpolation Spaces, Springer-Verlag, Berlin (1976). [3] P. Billard, Bases dans H et bases de sous espaces de dimension fini dans A, Linear Operators and Approximation, ISNM, Vol. 20, EL. Butzer, J.-R Kahane and B. Sz.-Nagy, eds, Birkh~iuser, Basel (1972). [4] S.V. Bo6kafiov, Existence of a basis in the space of functions analytic in the disc and some properties of the Franklin system, Mat. Sbornik 95 (137) (1974), 3-18 (in Russian). [5] S.V. Bo6kariov, Construction using Fejdr kernels of interpolating dyadic basis in the space of continuous functions, Trudy Steklov. Inst. 172 (1985), 29-59 (in Russian). [6] S.V. Bo6kariov, The conjugate Franklin system is a basis in the space of continuous functions, Dokl. AN SSSR 285 (3) (1985), 521-526. [7] J. Bourgain, Some remarks on Banach spaces in which martingale difference sequences are unconditional, Ark. Mat. 21 (1983), 163-168. [8] J. Bourgain, On bases in the disc algebra, Trans. Amer. Math. Soc. 285 (1) (1984), 133-139. [9] J. Bourgain, Vector-valued singular integrals and the H I-BMO duality, Probability Theory and Harmonic Analysis, J.A. Chao and W. Woyczyfiski, eds, Marcel Dekker, New York (1986), 1-19. [10] J. Bourgain, Homogeneous polynomials on the ball and polynomial bases, Israel J. Math. 68 (3) (1989), 327-347. [11] J. Bourgain, H.R Rosenthal and G. Schechtman, An ordinal LP-indexfor Banach spaces, with application to complemented subspaces of LP, Ann. of Math. 114 (1981), 193-228. [12] D.L. Burkholder, Martingales and singular integrals in Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 233-269. [13] Z. Ciesielski, Properties of the orthonormal Franklin system, II, Studia Math. 27 (1966), 289-323 [14] Z. Ciesielski, Spline bases in classical function spaces on compact C ~c manifolds, Part III, Constructive Theory of Functions, B. Sendov, P. Petrushev, R. Maleev and S. Tashev, eds, Sofia (1984), 214-223. [ 15] Z. Ciesielski and T. Figiel, Spline approximation and Besov spaces on compact manifolds, Studia Math. 75 (1982), 13-36. [16] Z. Ciesielski and T. Figiel, Spline bases in classical function spaces on compact C ~ manifolds, Part I, Studia Math. 76 (1983), 1-58. [17] Z. Ciesielski and T. Figiel, Spline bases in classical function spaces on compact C ~c manifolds, Part II, Studia Math. 76 (1983), 95-136.
T. Figiel and P. Wojtaszczyk
596
[18] Z.A. (~anturija, On unconditional polynomial bases in the space Lp, Studia Math. 71 (1981), 163-168. [19] I. Daubechies, Ten Lectures on Wavelets, SIAM, Philadelphia (1992). [20] G. David and J.L. Journ6, A boundedness criterion for generalized Calderon-Zygmund operators, Ann. of Math. 120 (1984), 371-397. [21] M. Defant and M. Junge, Unconditional orthonormal systems, Math. Nachr. 158 (1992), 233-240. [22] R.A. DeVore and G.G. Lorentz, Constructive Approximation, Springer-Verlag, Berlin (1993). [23] L.E. Dor and E. Odell, Monotone bases in Lp, Pacific J. Math. 60 (2) (1975), 51-61. [24] T. Figiel, On equivalence of some bases to the Haar system in spaces of vector-valued functions, Bull. Pol. Acad. Sci. 36 (1987), 119-131. [25] T. Figiel, Singular integral operators: a martingale approach, Geometry of Banach Spaces, Proc. Int. Conf. Strobl, Austria 1989, London Math. Soc. Lecture Notes 158, P.EX. Mtiller and W. Schachermayer, eds, Cambridge University Press, Cambridge (1990), 95-110. [26] Ph. Franklin, A set of continuous orthogonalfunctions, Math. Annalen 100 (1928), 522-529. [27] G. Gevorkian and A. Kamont, On general Franklin systems, Dissertationes Math. (Rozprawy Mat.) 374 (1998). [28] G.G. Gevorkian and B. Wolnik, The shift operator on the wavelet, Preprint. [29] G. Gripenberg, Wavelet bases in Lp(R), Studia Math. 106 (2) (1993), 175-187. [30] V.I. Gurarii and P.I. Gurarii, Bases in uniformly convex and uniformly smooth Banach spaces, Izv. Akad. Nauk SSSR, Ser. Math. 35 (1971), 210-215. [31] R.C. James, Super-reflexive spaces with bases, Pacific J. Math. 41 (1972), 409-419. [32] W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84.
[33] W.B. Johnson, H.E Rosenthal and M. Zippin, On bases, finite dimensional decompositions and weaker structures in Banach spaces, Israel J. Math. 9 (4) (1971), 488-506. [34] N.J. Kalton, C. Leranoz and E Wojtaszczyk, Uniqueness of unconditional bases in quasi-Banach spaces with applications to Hardy spaces, Israel J. Math. 72 (3) (1990), 299-311. [35] M.I. Kadets and A. Pe~czyfiski, Bases, lacunary sequences and complemented subspaces in the spaces Lp, Studia Math. 21 (1962), 161-176. [36] B.S. Kashin and A.A. Saakyan, Orthogonal Series, Nauka, Moscow (1984) (in Russian). [37] Y. Katznelson, An Introduction to Harmonic Analysis, Wiley, New York (1968). [38] G. K6the, Das Triigheitsgesetz der quadratischen Formen in Hilbertschen Raum, Math. Z. 41 (1936), 137152. [39] G. Kyriazis and E Petrushev, New bases for Triebel-Lizorkin and Besov spaces, Preprint. [40] J. Lindenstrauss, A short proof of Liapounoff's convexity theorem, J. Math. and Mech. 15 (1966), 971-972. [41] J. Lindenstrauss and A. Petczyfiski, Absolutely summing operators in 12p-spaces and their applications, Studia Math. 29 (1968), 275-326. [42] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces, L Sequence Spaces, Springer-Verlag, Berlin (1977). [43] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces, II. Function Spaces, Springer-Verlag, Berlin (1979). [44] R.A. Lorentz and A. Sahakian, Orthonormal trigonometric Schauder bases of optimal degree for C(K), J. Fourier Anal. Appl. 1 (1) (1994), 103-112. [45] W. Lusky, On Banach spaces with the commuting bounded approximation property, Arch. Math. 58 (1992), 568-574. [46] W. Lusky, On Banach spaces with bases, J. Funct. Anal. 138 (2) (1996), 410-425. [47] W. Lusky, Three space properties and basis extensions, Preprint. [48] Ch.A. McCarthy, Cp, Israel J. Math. 5 (1967), 24c-271. [49] O.C. McGehee, L. Pigno and B. Smith, Hardy's inequality and the L 1 norm of exponential sums, Ann. of Math. 113 (3) (1981), 613-618. [50] Y. Meyer, Wavelets and Operators, Cambridge University Press, Cambridge (1992). [51 ] E. Michael and A. Pdczyfiski, Separable Banach spaces which admit s ~ approximation, Israel J. Math. 4 (1966), 189-198. [52] B. Muckenhoupt, Mean convergence of Jacobi series, Proc. Amer. Math. Soc. 23 (1969), 306-310.
Special bases in function spaces
597
[53] S.M. Nikol'skii, Approximation of Functions of Several Variables and Imbedding Theorems, SpringerVerlag, Berlin (1975). [54] I. Novikov, Criterion of equivalence of Haar and Franklin systems in symmetric spaces, Mat. Zametki 52 (3) (1992), 96-101. [55] I. Novikov and E. Semenov, Haar Series and Linear Operators, Kluwer Academic Publishers, Dordrecht (1997). [56] A. Petczyfiski, Projections in certain Banach spaces, Studia Math. 19 (1960), 209-228. [57] A. Petczyfiski, Sobolev spaces, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [58] E Petrushev, Bases consisting of rational functions of uniformly bounded degrees or more general functions, J. Funct. Anal. (to appear). [59] A. Pietsch and J. Wenzel, Orthonormal Systems and Banach Space Geometry, Encyclopedia of Math. Appl., Vol. 70, Cambridge University Press, Cambridge (1998). [60] W. Pompe, Unconditional biorthogonal wavelet bases in L p(Rd), Preprint. [61] A.A. Privalov, On the growth of degrees of polynomial basis and approximation of trigonometric projectors, Mat. Zametki 42 (2) (1987), 207-214 (in Russian). [62] S. Ropela, Spline bases in Besov spaces, Bull. Acad. Pol. Sci. Serie Math., Astr., Phys. 24 (5) (1976), 319-325. [63] F. Schipp, W.R. Wade and E Simon, Walsh Series; An Introduction to Dyadic Harmonic Analysis, Akad6miai Kiad6, Budapest (1990). [64] P. Sj61in, The Haar and Franklin systems are not equivalent bases in L 1, Bull. Acad. Pol. Sci. Serie Math. XXV (11) (1977), 1099-1100. [65] J.O. Str6mberg, A modified Franklin system and higher order spline systems on R n as unconditional bases for Hardy spaces, Conference on Harmonic Analysis in Honor of Antoni Zygmund, W. Beckner, A.E Calderon, R. Fefferman and EW. Jones, eds, Wadsworth, Belmont, CA (1983), 475-494. [66] S.J. Szarek, Bases and biorthogonal systems in the spaces C and L1, Arkiv Mat. 17 (2) (1979), 255-271. [67] Wo-Sang Young, A note on Walsh-Fourier series, Proc. Amer. Math. Soc. 59 (1976), 305-310. [68] E Wojtaszczyk, Hp-spaces, p <<,1 and spline systems, Studia Math. 77 (1984), 289-320. [69] E Wojtaszczyk and K. Woiniakowski, Orthonormal polynomial bases in function spaces, Israel J. Math. 75 (1991), 167-191. [70] E Wojtaszczyk, On unconditional polynomial bases in L p and Bergman spaces, Constr. Approx. 13 (1997), 1-15. [71] P. Wojtaszczyk, Wavelets as unconditional bases in L p (R), J. Fourier Anal. Appl. 5.1 (1999), 73-85. [72] P. Wojtaszczyk, Banach Spaces for Analysts, Cambridge Studies in Advanced Mathematics 25, Cambridge University Press, Cambridge (1991). [73] E Wojtaszczyk, A Mathematical Introduction to Wavelets, London Math. Soc. Student Texts 37, Cambridge University Press, Cambridge (1997). [74] P. Wojtaszczyk, Spaces of analytic functions with integral norm, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published).
This Page Intentionally Left Blank
CHAPTER
15
Infinite Dimensional Convexity V.E Fonf Department of Mathematics, Ben Gurion, University of the Negev, Beer-Sheva, Israel E-maih fonf@ indigo.bgu.ac.il
J. Lindenstrauss Institute of Mathematics, The Hebrew University of Jerusalem, Jerusalem, Israel E-maih joram @math.huji.ac, il
R.R. Phelps Department of Mathematics, University of Washington, Seattle, WA, USA E-maih phelps@math,washington.edu
D e d i c a t e d to the memory o f Yaki Sternfeld
Contents 1. Introduction and classical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Integral representation theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Compact simplices and related sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Exposed points and the Radon-Nikod3)m property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Boundaries and support points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Convex polytopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7. Miscellaneous topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
H A N D B O O K OF THE G E O M E T R Y OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 599
601 603 613 628 639 650 660 665
This Page Intentionally Left Blank
Infinite dimensional convexity
601
1. Introduction and classical results The study of convex sets in infinite dimensional spaces lies at the heart of the geometry of Banach spaces. For instance, the unit ball completely determines the metric properties of a Banach space, while its weak*-compact convex dual unit ball plays a ubiquitous role. In a single chapter we can describe only a portion of the vast amount of material concerning infinite dimensional convex sets, with the choice of topics being greatly influenced by the interests of the authors. We start by emphasizing compact convex sets in locally convex spaces and then turn to closed convex bounded (CCB) sets in Banach spaces and their relationship to the structure of the space. More explicitly, following this Introduction we have the sections 2. Integral representation theorems (for elements of compact convex sets, as initiated by Choquet). 3. Compact simplices (the compact convex sets where integral representations are unique) and related sets. 4. Exposed points and the Radon-Nikod3~m Property (RNP) (for Banach spaces and their convex subsets, also in the setting of complex convexity). 5. Boundaries and support points (including James' theorem and its relatives). 6. Convex polytopes (in infinite dimensional spaces). 7. Miscellaneous topics: (a) Stable convex sets. (b) Sets with dense extreme points. (c) Convex tiling. (d) Proximinal sets and related sets. When presenting a specific result, we will generally refer to the original paper in which it appeared, as well as to some books or expository papers which include the result as part of a wider treatment. We have sometimes included proofs, especially if they are particularly short or illuminating or previously unpublished. We recall some of the basic definitions: A map T from one linear space X to another is said to be affine provided it "preserves affine combinations", that is, T()~x + (1 - )~)y) = )~Tx + (1 - )~)Ty whenever x, y E X and )~ is real. In particular, affine images of convex sets are convex. For a subset A of a linear space X, the set co A (the convex hull of A) denotes the smallest convex set containing A. The smallest closed convex set containing A, cl co A (the closed convex hull of A) is simply the closure of co A. The linear span of A is denoted by span A; it is the smallest linear space in X containing A. A subset A C X is said to be infinite dimensional if span A is infinite dimensional. The closure of span A is denoted by[a]. The Hahn-Banach extension theorem has a geometric version, the separation theorem, which is arguably the most fundamental result in the study of infinite dimensional convex sets. It can be formulated a follows: If K and C are nonempty convex subsets of a locally convex space X (over the real numbers) and if K is disjoint from the nonempty interior of C, then there exists a nonzero continuous linear functional f E X* such that sup f (C) <~ i n f f ( K ) . If C has empty interior, a corollary to this still produces a functional f with the same properties, provided K is compact and convex, C is closed and convex, and the
602
V.P. Fonf et al.
two sets are disjoint. By a locally convex space we always mean a space which is also Hausdorff. DEFINITION 1.1. A point x in a convex set C is said to be an extreme point of C provided x--y=z whenevery, z~C andx=Ay+(1-A)z for s o m e 0 < A < 1. The set of all extreme points of C is denoted by ext C. A face F of a convex set K is a nonempty convex subset of K such that y, z E F whenever y, z 6 K and Ay + (1 - A)z 6 F for some 0
I f K is a compact convex subset of a locally convex
This theorem is frequently applied when K is the (weak* compact convex) unit ball of the dual X* of a real Banach space X or, more generally, when K is any norm-bounded weak* closed convex subset of X*. Consider, for instance, X = L1 (S2, r , / z ) for a o-finite measure #. Then X* -- L ~ (1-2, r , #) and the extreme points of the unit ball of X* are precisely the functions of modulus one (a.e. #). Similarly, if one considers K = { f L ~ " 0 ~< f ~< 1}, then K is weak* compact and convex and ext K = { f 6 K" f 2 = f}. This last fact is a key step in one proof [120] of Liapunov's Theorem, which states that i f # 1 , # 2 . . . . . # n are finite nonatomic measures on (1-2, r ) , then the range in R n o f the vector measure S --+ (lZl (S), #2(S) . . . . . # n ( S ) ) is compact and convex. There is a useful partial converse to Theorem 1.2. PROPOSITION 1.3 (Milman). I f K is a compact convex subset o f a locally convex space and if A C K is such that K = cl co A, then ext K is contained in the closure of A. In finite dimensions, one need not take the closure of the extreme points: THEOREM 1.4 (Minkowski-Carath6odory). I f K is an n-dimensional compact convex set, then K = co(ext K). Moreover, any point of K is a convex combination o f at most n + 1 extreme points o f K. Klee [ 101 ] has extended the foregoing theorems to certain noncompact sets, utilizing the following concepts. DEFINITION 1.5. An extreme ray of a convex set C is an open half-line p C C with the property that the open segment (x, y) is contained in p whenever (x, y) C C and (x, y)
Infinite dimensional convexity
603
intersects p. A set is said to be linearly closed if its intersection with each line is closed. We let exr C denote the union of the extreme rays of C. THEOREM 1.6. I f C is a locally compact closed convex subset o f a locally convex space, and if C contains no line, then C is the closed convex hull o f ext C U exr C. THEOREM 1.7. I f C is a linearly closed finite dimensional convex set which contains no line, then C is the convex hull o f ext C U exr C. The Krein-Milman theorem (or sometimes merely the existence of an extreme point) has found wide application, de Branges [39] has used it to prove the Stone-Weierstrass theorem. (See [152, Theorem 5.7] for a generalization.) Fixed-point theorems are fundamental to many parts of analysis. One that involves compact convex sets is the following. A proof may be found in practically any book on general functional analysis. THEOREM 1.8 (Schauder-Tichonoff fixed point theorem). I f K is a compact convex subset o f a locally convex space, then K has the fixed point property, that is, if f : K --+ K is continuous, then f has a f i x e d p o i n t in K. Klee [ 100] has shown that if K is a convex but not compact subset of a normed linear space then K fails the fixed point property. It should be noted that any infinite-dimensional compact convex subset o f a Banach space is homeomorphic to the Hilbert "cube" or "parallelotope", namely, to the set {X = (Xn) E/~2: IXnl ~ I / n , n = 1, 2 . . . . }. This was first shown by Keller [98] for infinitedimensional compact convex subsets of Hilbert space. Klee [100] observed that any compact (convex) set in a Banach space is affinely homeomorphic to a subset of Hilbert space. Hence Keller's theorem applies to all compact convex sets in a Banach space. Further material on fixed points can be found in [44] and more about infinite-dimensional topology in [ 10] and [ 165]. Though we consider in this survey only locally convex spaces, it is worthwhile to say a few words also on Hausdorff topological linear spaces which are not locally convex. In such spaces compact convex sets behave differently. It was proved by Roberts [ 145] that if 0 ~< p < 1, then the metric linear space Lp[0, 1] (with the obvious topology) contains compact convex sets which fail to have extreme points.
2. Integral representation theorems To illustrate what is meant by an "integral representation" of an element of a convex set, we reformulate Minkowski's Theorem 1.4: If K is a compact convex subset of a finite dimensional space X and x E K, then x is a finite convex combination of extreme points of K; that is, there exist extreme points x l . . . . , xk of K and positive numbers/s //~k with ~/14 1 such that x -- ~ lZiXi. This can be reformulated as an integral representation theorem as follows: For any point y of K let ey be the "point mass" at y, i.e., ey is the .
- -
.
.
.
.
V.P. Fonf et al.
604
Borel measure which equals 1 on any Borel subset of K which contains y, and equals 0 otherwise. Abbreviating exi by ei, let Ix = Y~ Ixiei; then Ix is a regular Borel measure on K, Ix >~ 0, and Ix(K) = 1. Furthermore, for any continuous linear functional f on X, we have f x f dIx - Y~Ixi f (xi) = f (Y~Ixixi) = f (x). The equality f ( x ) -- f x f dIx (for all f 6 X*) is what we mean when we say that Ix represents x. Since the continuous linear functionals f separate points of X, this is equivalent to saying that x = ~ #ixi. DEFINITION 2.1. Suppose that K is a nonempty compact subset of a locally convex space X, and that IX 6 P ( K ) , the set of all regular Borel probability measures on K. A point x in X is said to be represented by IX if f (x) = f/( f dIx for every continuous linear functional f on X. (We will sometimes write Ix ( f ) for f/( f dix, when no confusion can result.) (Other terminology: "x is the barycenter of IX", "x is the resultant of IX".) The restriction that X be locally convex is simply to insure the existence of sufficiently many functionals in X* to separate points; this guarantees that there is at most one point represented by IX. Note that any point x in K is trivially represented by ex; the interesting (and important) fact brought out by Minkowski's theorem above is that in the special case of a compact convex subset K of a finite dimensional space, each x in K may be represented by a probability measure which is supported by the extreme points of K. To obtain such a representation in infinite dimensional spaces, a necessary first step is to show (below) that every probability measure on K has a resultant. PROPOSITION 2.2. Suppose that A is a compact subset of a locally convex space X, and that the closed convex hull K of A is compact. I f # is a probability measure on A, then there exists a unique point x in K which is represented by Ix, and the function Ix --+ (resultant of ix) is an affine weak* continuous mapfrom P ( A ) into K. PROOF. We want to show that the compact convex set f ( x ) = fA f d # for each f in X*. For each f , let H i closed hyperplanes, and we want to show that (-]{Hi: compact, it suffices to show that for any finite set f/, . . . , is nonempty. To this end, define
T" K ~ R n
K contains a point x such that = {y: f ( y ) -- Ix(f)}; these are f ~ X*} A K # 0. Since K is fn in X*, the set (']in_l Hfi A K
by Ty = ( f l ( y ) , f2(y) . . . . . fn(Y));
then T is linear and continuous, so that TK is compact and convex. It suffices to show that p E TK, where p = ( # ( f l ) , # ( f 2 ) . . . . . # ( f n ) ) . If p q~ TK, by the separation theorem there exists a linear functional on R n which strictly separates p and TK; representing the functional by a = (al, a2 . . . . . an), this means that (a, p) > sup{(a, Ty): y ~ K}. If we define g in X* by g = ~ ai fi, then the last assertion becomes fa g dix > sup g ( K ) . Since A C K and # ( A ) = 1, this is impossible, and the first part of the proof is complete. It is straightforward to verify that the resultant map is affine, and an argument using the compactness of K shows that it is continuous. U]
Infinite dimensional convexity
605
A simple, but useful, characterization of the closed convex hull of a compact set can be given in terms of measures and their barycenters. PROPOSITION 2.3. Suppose that A is a compact subset of a locally convex space X. A point x in X is in the closed convex hull K of A if and only if there exists a probability measure It on A which represents x. PROOF. If # is a probability measure on A which represents x, then for each f in X*, f ( x ) = # ( f ) ~< sup f ( A ) <~ sup f ( K ) . Since g is closed and convex, it follows from the separation theorem that x is in K. Conversely, if x is in K, there exists a net in the convex hull of A which converges to x. Equivalently, there exist points y~ of the form o/ o/ yo~ -- Y-~i"~=li"~ -~i (i"~ > O, ~ ~'i -- 1, x i E A, c~ in some directed set) which converge to x. We may represent each y~ by the probability measure #~ -- ~ ~.~ exZ. By the Riesz representation theorem, the set of all probability measures on A may be identified with a weak*-compact convex subset of C(A)*, and hence there exists a subnet #/~ of #~ converging (in the weak* topology of C(A)*) to a probability measure # on A. In particular, each f in X* is (when restricted to A) in C ( A ) , so limf(y/~) = l i m f f d # ~ = f f d # . Since y~ converges to x, so does the subnet y~, and hence f ( x ) = lim f ( y ~ ) = fa f dlz for each f in X*, which completes the proof. R The above proposition makes it easy to reformulate the Krein-Milman theorem as an integral representation theorem. Recall the statement: If K is a compact convex subset of a locally convex space, then K is the closed convex hull of its extreme points. Our reformulation is the following: Every point of a compact convex subset K of a locally convex space is the barycenter of a probability measure on K which is supported by the closure of the extreme points of K. To prove the equivalence of these two assertions, suppose the former holds and that x is in K. Let A be the closure of the extreme points of K; then x is in the closed convex hull of A. By the preceding proposition, then, x is the barycenter of a probability measure # on A. If we extend # (in the obvious way) to K, we get the desired result. Conversely, suppose the second assertion is valid and that x is in K. Then (defining A as above), the same proposition shows that x is in the closed convex hull of A, hence in the closed convex hull of the extreme points of K. In applications of this integral form of the Krein-Milman theorem, it frequently happens that the extreme points of K are actually closed. This is the case, for instance, in the next three classical examples. That these classical results can be deduced from the KreinMilman theorem was pointed out by Choquet in [26] (see also [31, Vol. II]). (a) Completely monotonic functions. A real valued function f on (0, ec) is said to be completely monotonic if f has derivatives f(0) = f, f(1), f(2) . . . . of all orders and if ( - 1 ) n f(n) ~> 0 for n = 0, 1, 2 . . . . . Thus, f is nonnegative and nonincreasing, as is each of the functions ( - 1)n f(~). (Some examples: x -~ and e -~x (or ~> 0).) Bernstein proved a fundamental representation theorem for such functions. (See [ 171, Chapter 4] for several proofs and much related material.) We will state the theorem only for bounded functions; the extension to unbounded functions (with infinite representing measures) follows from this by classical arguments [171]. We denote the one-point compactification of [0, cx~) by
[o, ~ ] .
V.P Fonf et al.
606
If f is bounded and completely monotonic on (0, e~), then there exists a unique nonnegative Borel measure tz on [0, oc] such that/z([0, cx~]) = f ( 0 + ) and for each x > O, THEOREM 2.4 (Bernstein).
f (x) =
f0 ~
e -ax d#(oe).
The idea of the proof is to let K be the set of all bounded completely monotonic functions f such that f (0+) ~< 1. This is a compact convex subset of the space of all real-valued infinitely differentiable functions on (0, oe), with the topology of uniform convergence (of functions and all their derivatives) on compact subsets of (0, oc). The extreme points of K are the functions x --+ e -c~x, 0 ~ oe ~< oe, and the Krein-Milman theorem produces a measure on this set which yields the desired measure # on [0, oc]. Uniqueness comes from the Stone-Weierstrass theorem. (b) Infinitely divisible functions. An element f of the set K above is said to be infinitely divisible if for each n ~> 1 there exists fn in K such that f = (fn)n. Let C be the convex cone of all functions g of the form g = - log f , where f 6 K, f > 0, and f is infinitely divisible. Those g E C such that g (1) -- e-1 form a compact convex set K1, and the extreme points of this set are the functions g~(x) -- exp
{le X /
go(x) -- e -x
1 - e -~
and
'
0
goc(x)=e -l.
The Krein-Milman theorem (applied to K1 in the weak topology defined by the functionals "evaluation at a positive rational number"), followed by exponentiating, leads to the following result, called the L6vy-Feller-Khintchine formula. (See [56] for much background material on this topic and its role in probability theory.) THEOREM 2.5. To every infinitely divisible function f of K there corresponds a unique finite, nonnegative Borel measure # on [0, oe] such that f (x) -- exp -
f
~x ~11 --e e- -~ d/z(a)},
x>0.
(c) Positive definite functions. A complex valued function f on an Abelian group G is said to be positive-definite provided
g/ Z
~i~j f(ti
-- tj) >/0
i,j=l
whenever tl . . . . . tn are elements of G and )~l An are complex numbers. It is easily seen that if f is positive definite, then f ( 0 ) is real and If(t)l ~< f ( 0 ) for all t in G. If a function f is a character of G (i.e., a homomorphism of G into the group of all complex numbers of modulus 1), then f is positive definite. Suppose that G is locally compact and . . . . .
Infinite dimensional convexity
607
let K be the set of all continuous positive definite functions f on G such that Ilfll~ ~ 1. Then K is a weak*-closed subset of all those f in L ~ ( G ) which satisfy
f f g(s + t)g(s) f (t) ds dt >~0
(g E Ll (G)).
The nonzero extreme points of K are the (essentially) continuous characters X of G, which leads to the following theorem. THEOREM 2.6. To every continuous positive definite function f on G there exists a finite nonnegative measure # on the characters such that
f (t) - f •215
(t E G).
This is a generalization of a classical theorem of Bochner (where G is the real line and each character is of the form t --+ e ixt for some real x). Since the extreme points form a closed set, it can be proved by the Krein-Milman theorem; the Stone-Weierstrass theorem can be used to show that # is uniquely determined by f . This result has a close connection with group representations, since each continuous positive definite function on G defines, in a canonical way, a continuous unitary representation of G, and the characters correspond to the irreducible representations. The above integral representation essentially shows that every cyclic representation of G (and hence every representation of G) is a "direct integral" of irreducible representations. For further details, see [50] and [132]. Our next aim is to present Choquet's integral representation theorem for points in a compact convex set K by measures "supported" by ext K even if ext K is not closed. It is convenient to use a more intrinsic (but equivalent) formulation of "represented" by introducing the space of continuous affine functions on K. DEFINITION 2.7. For a compact convex subset K of a locally convex space X, define
A (K) to be the space of all continuous real-valued affine functions h on K, with supremum norm. It is clear that A(K) is uniformly closed subspace of C(K) and that it contains the subspace M of all functions of the form K 9 x --+ f ( x ) + r, where f 6 X* and r is a real number, so there are enough functions in A (K) to separate points of K. Note that K can be identified canonically with a subset K of the unit ball of A(K)* (~:(f) -- f ( k ) for f E A(K)) and that the unit ball of A(K)* is c o ( K U - K ) . The reformulation of "represents" is contained in the next result. PROPOSITION 2.8. The subspace M (above) is uniformly dense in A(K), and hence a point x E K is represented by a probability measure # on K if and only if h (x) = f h d# for all h E A (K). PROOF. Given h E A(K) and e > 0, consider the two disjoint compact convex subsets of X x R defined by K1 = {(x, r) 6 X x IR: x E K and r = h(x)} and K2 = {(x, r) E
608
V.P Fonf et al.
X x R: x 9 K and r = h(x) + e}. The separation theorem yields a hyperplane in X x JR, which is necessarily the graph of a continuous affine function g on all of X such that h(x) < g(x) < h(x) + e for x 9 K. Since the function g must have the form f + r, f 9 X*, r 9 JR, the density assertion has been proved. [3 One difficulty in finding measures supported by the extreme points of K stems from the fact that the set of extreme points need not be a Borel set [ 12]. This difficulty is avoided in case K is metrizable, as shown by the following observation. PROPOSITION 2.9. If K is a metrizable, compact convex subset of a topological vector space, then the extreme points of K form a G~ set.
PROOF. Suppose that the topology of K is given by the metric d, and for each n ~> 1 let Fn = {x 9 K" x = 2-1 (y + z), where y, z 9 K and d (y, z) ~> n-1 }. It is easily checked that each Fn is closed, and that a point x of K is not extreme if and only if it is in some Fn. Thus, the complement of the extreme points is an F,,. D THEOREM 2.10 ([27,29]). Suppose that K is a metrizable compact convex subset of a locally convex space X, and that xo is an element of K. Then there is a probability measure # on K which represents xo and is supported by the extreme points of K (i.e., #(ext K) -- 1). The proof (below) is due to Herv6 [86] and Bonsall [ 14]. It uses the concept of the upper envelope f of a real-valued function f on K. DEFINITION 2.11. If f is a real-valued function on K and x 9 K, let f ( x ) = inf{h(x)" h 9 A ( K ) and h ~> f}; this is called the upper envelope of f . The function f has the following useful properties, whose proofs are elementary: (a) f is concave, bounded whenever f is bounded, and upper semicontinuous (hence Borel measurable). (b) f ~< f and if f is concave and upper semicontinuous, then f = f . (c) If f, g are bounded, then f + g ~< f + ~ and I f - g[ ~< [If - g[], while f + g -f + g if g 9 A ( K ) . If r > 0, then r f = r f . PROOF OF THEOREM 2.10. Since K is metrizable, C ( K ) (and hence its subspace A ( K ) ) is separable. Thus, we can choose a sequence of functions {hn} in A ( K ) such that Ilhnll~ - 1 and the set {hn}neC__lis dense in the unit sphere of A(K); in particular, it separates points of K. Let f0 -- ~ 2-nh2; the series converges uniformly, hence is in C ( K ) and it is a strictly convex function in C(K). (Indeed, if x 7~ y, then hn(x) r hn(y) for some n, so hn is nonconstant and affine on the segment [x, y]. It follows that h 2 is strictly convex on [x, y] and therefore fo is strictly convex on [x, y].) Let B denote the subspace A ( K ) + R f o of C ( K ) generated by A ( K ) and f0. From property (c) above, it follows that the functional p defined on C ( K ) by p(g) = g,(xo) (g E C(K)) is subadditive and positivehomogeneous. Define a linear functional on B by h + rfo --+ h(xo) + r fo(xo) (h in A ( K ) , r real). We will show that this functional is dominated on B by the functional p, i.e., that
Infinite dimensional convexity
609
h(xo) + r fo(xo) <~ (h + rfo)(xo) for each h ~ A ( K ) , r ~ JR. If r >~ 0, then h + r fo -h + rfo, by (c), while if r < 0, then h + rfo is concave, and hence by (b) h + rfo = h + rfo ~ h + r fo. By the H a h n - B a n a c h theorem there exists a linear functional m on C ( K ) such that m(g) <~ g(xo) for g in C ( K ) , and m(h + rfo) -- h(xo) + rfo(xo) if h ~ A ( K ) , r ~ JR. If g E C ( K ) and g ~< 0, then 0 >~ g'(x0) ) m(g), i.e., m is nonpositive on nonpositive functions and hence is continuous. By the Riesz representation theorem, there exists a nonnegative regular Borel measure # on K such that m(g) -- lz(g) for g in C ( K ) . Since 1 6 A ( K ) , we see that 1 -- m(1) - #(1), so # is a probability measure. Furthermore, # ( f o ) -- m ( f o ) -- fo(xo). Now, fo ~< fo, so # ( f o ) ~< # ( f o ) . On the other hand, if h ~ A ( K ) and h ) fo, then h >~ fo, and consequently h(xo) = m(h) -- # ( h ) ~> # ( j ~ ) . It follows from the definition of fo that fo(xo) >1 # ( f o ) , and therefore # ( f o ) -- # ( f o ) . This last fact implies that # vanishes on the complement of X - {x" f o ( x ) - fo(x)}. We complete the proof by showing that A" is contained in the set of extreme points of K. Indeed, if x - 89 y + 89 z, where y and z are distinct points of K, then the strict convexity of fo implies that f o ( x ) < 89 fo(Y)+
89 fo(z) <~ I f o ( Y ) + I fo(z) <~ fo(x).
D
See Section 3 for an application of Choquet's theorem to ergodic measures (where the extreme points need not form a closed set). In most applications of Choquet's theorem the representing measure is unique and therefore they fit into the setting of Section 3. Here is a simple application to normed linear spaces (without any uniqueness assumption). PROPOSITION 2.12 ([ 142]). Let X be a normed linear space and suppose that x, Xn (n 1, 2, 3 . . . . ) are elements of X. Then the sequence {Xn } converges weakly to x if(and only if) {Xn } is bounded and lim f (Xn) - f (x) for each extreme point f ofthe unit ball Bx, of X*. PROOF. Suppose that f ( x n ) ~ f ( x ) for each f in ext B x , and that g is an arbitrary element of B x , ; we must show that g(xn) --+ g(x). Let Y denote the closed linear span of the sequence {xn } and x, so that Y is separable and hence the unit ball K of Y* is weak* compact, convex and metrizable. The restriction go of g to Y is in K and by Choquet's theorem there exists a probability measure/z on K which represents go and is supported by ext K. If f E ext K, then it has norm one and (by the Krein-Milman theorem) the weak* compact convex set of its H a h n - B a n a c h extensions to X has at least one extreme point. By regarding the points xn as functions on K, an application of the Lebesgue bounded convergence theorem shows that g (x) -- lim g (xn). [3 The central concept in proving existence (and uniqueness) of integral representations in the nonmetrizable case is the notion of a maximal measure, which is defined as follows" DEFINITION 2.13. Suppose that )~ and # are Borel probability measures on a compact convex subset K of a locally convex space. Define X >- # provided X ( f ) ~> # ( f ) for every convex continuous function f on K. This is a reflexive, transitive and proper ordering. We say that/z is a maximal measure if it is maximal in this ordering. (The fact that # >- X and X >- # imply that # = X comes from the Stone-Weierstrass theorem, which guarantees that the span of the differences of convex continuous functions is dense in C (K).)
V.P. Fonf et al.
610
It is evident that if )~ >- Ix then for every continuous affine f , ~ ( f ) = # ( f ) . It is not hard to verify that IX represents the point x 9 K if and only if IX >- ex and it is geometrically persuasive that if )~ >- Ix >- ex, then the support of )~ must be in some sense closer to ext K than the support of Ix. For compact metric K, this ordering can be characterized in terms of dilations. DEFINITION 2.14. A mapping T : K --+ P ( K ) is called a dilation if Tx >- ex for each x 9 K and the function x --+ Tx ( f ) is Borel measurable for each f 9 C (K). We can extend T to a map (still called a dilation) from P ( K ) into itself by
( T # ) ( f ) = fK T x ( f ) dix(x),
IX 9 P ( K ) , f
9 C(K).
It is readily verified that if )~ = TIx, then k >- Ix. The nontrivial converse is true if K is compact metric. THEOREM 2.15 ([25]). Suppose that K is a compact metric convex subset o f a locally convex space and that )~, IX are regular Borel probability measures on K. Then )~ >- tz iff there exists a dilation T such that )~ = T Ix. A proof of this theorem can be found in [ 136, Section 13] where the history of this result is also discussed. By an application of Zorn's lemma to the (nonempty) set of all probability measures which represent x it follows easily that there always exists a maximal measure representing x. The following characterization of maximal measures is central to both the existence and uniqueness theorems. PROPOSITION 2.16. A probability measure IX on K is maximal if and only if i x ( f ) = I x ( f ) f o r every convex continuous function f on K; equivalently, f o r every continuous function f on K. Equivalently, Ix is maximal if and only if Ix is supported by {x 9 K: f (x) = f ( x ) } f o r each such f . It follows from this and from the existence of a strictly convex continuous function on K in the metrizable case that (in the latter case) IX is maximal if and only if Ix(ext K) -- 1. In the general case it is easy to show that ext K = fEC(K)
It can be deduced from this fact and Proposition 2.16 without much difficulty that Ix(D) -- 0 for every compact set D for which D C E where E is a G~ subset of K which is disjoint from ext K. We thus get the Choquet-Bishop-de Leeuw theorem: THEOREM 2.17 ([12]). Suppose that K is a compact convex subset of a locally convex space X, and that xo is in K. Then there exists a probability measure tx on K which represents xo and which vanishes on every Baire subset of K as well as every G~ subset o f K which is disjoint from the set of extreme points of K.
Infinite dimensional convexity
611
The proofs of 2.16 and 2.17 can be found also in [2, Section 1.4], and [136, Section 4]. Here is an application of the Choquet-Bishop-de Leeuw theorem to Banach spaces, proved independently by Haydon [84] and Kadets and Fonf [96]. For a generalization, see Theorem 5.7. THEOREM 2.18. Let X be a real Banach space and let K be a weak* compact convex subset of X* such that ext K is norm separable. Then K is the norm closed convex hull of its extreme points (and hence is itself norm separable). PROOF. Let M = sup{llf[l: f ~ K}, suppose e > 0 and let {fi} be a norm dense subset of ext K. For each i, let Bi denote the intersection with K of the closed ball of radius e/3 centered at j~. Thus, each Bi is weak* compact and convex and U Bi D ext K. Let f be a point of K and let/z be a maximal probability measure on K with resultant r ( # ) = f . Since U Bi is a weak* F~ set, we have # ( U Bi) -- 1. Let n be a positive integer such that, if D = /7 U i = I Bi, then # ( D ) > 1 - e / 3 M . T h e n / z can be decomposed as # = )~#l + (1 - )~)#2, where )~ = # ( D ) and # l , #2 are probability measures on K defined by ~#l = #ID
and
(1 - )~)#2 = # I ( K \ D ) -
(If)~ = 1, let # 2 be an arbitrary probability measure on K.) Then f = r ( # ) = )~r(/zl) + (1 - ~,)r(/z2). Since r ( # 2 ) E K we have oe
II f - )~r (/zl
-
)ll r < 2 11
9M ~ .
3M
-o
3
Since #1 is a probability measure supported by uin=l Bi, the point r ( # l ) lies in the convex hull of U i n= I Bi, which is weak* compact. Hence r ( # l ) - - ~-~i=1 n i~i gi , where gi E Bi , i~i 0 and y~in=l )~i = 1. Let h = ~in=l )~i fi. This is a point of co(ext K) and IIr(#l) - h l l ~< e/3. Consequently, II f - h II ~< [] f - )~r (/z,
+
-
)11 r
+ IIr
- h I[
+
+
Thus, co(ext K) is norm dense in K. Recall that when we say that a probability measure # on a compact convex set K "represents" a point x of K, we mean that # ( f ) -- f (x) for each continuous affine function f on K. One way of extending the representation theorems is to show that this latter equality holds for a larger class of functions. PROPOSITION 2.19 ([30]). If K is a compact convex subset of a locally convex space X and if lz is a probability measure on K with resultant x, then l z ( f ) = f (x) for each affine function f of first Baire class on K. For a proof of this proposition see also [2, p. 16]. Using this proposition Mokobodzki and Rogalski [148] proved that every affinefunction of the first Baire class on a compact convex set K is, in fact, the pointwise limit of a sequence of continuous affine functions on K. In [30] it is shown also that Proposition 2.19 is no longer true for affine functions of the second Baire class (i.e., the pointwise limit of a sequence of functions of the first Baire
612
V.P. Fonf et al.
class). Indeed, if K is the set of probability measures on [0, 1] say, then the function q9 which assigns to every probability measure on [0, 1] the total mass of its atomic part, can be shown to be of the second Baire class. It is obviously also an affine function from K to [0, 1]. If x 6 K is the Lebesgue measure and # its representing measure on ext K then clearly
0- ~O(X)_-?6fextK ~0(y) d # ( y )
= 1.
A special case of Proposition 2.19 is the following
PROPOSITION 2.20. Suppose that the Banach space X is separable and does not contain a subspace isomorphic to el. Then for each functional f in the unit ball B x , there exists a probability measure lz on ext B x , such that F (f)
-- fext Bx,
F(g)d#(g)
for each F E X**.
Since B x , is convex, compact and metrizable in the weak* topology, Choquet's original theorem yields a probability measure # on ext B x , such that the displayed equality holds for those F which are weak* continuous. The hypothesis that X does not contain el guarantees that every element of X** is the pointwise limit of a sequence of such weak* continuous functionals [134], so the foregoing proposition (or Lebesgue's bounded convergence theorem) applies. In the next section we examine the case when the maximal representing measure of each point is unique. The next theorem shows that even if the maximal measure is nonunique it is always possible to select one such measure in a nice way. THEOREM 2.21 ([143,168]). Suppose that K is a metrizable compact convex set. Then there exists a Borel measurable map x --+ lZx from K into the probability measures on ext K such that, for each x E K, the measure #x represents x, and such that lZx is an extreme point of the set of all probability measures on ext K which represent x. The question of whether a continuous function on ext K can be extended to an affine continuous function on K can be considered as a "Dirichlet problem for compact convex sets". The solution presented below uses a modified version of the concept of the upper semicontinuous envelope of a real-valued function introduced earlier in the proof of Choquet's Theorem 2.10. Suppose that K is compact and convex and that f :ext K --+ R. For any x E K define f(x)-
inf{h(x): h E A ( K ) , hlextK ~> f } .
We define the related lower semicontinuous lower envelope f by f -
- (-f~-").
THEOREM 2.22 ([1]). Suppose that K is compact and convex and that f is a real-valued continuous function on ext K. Then f has an extension to an element of A ( K ) if and only if f f on c l e x t K and # ( f ) v ( f ) whenever lz and v are two maximal probability measures on K with the same barycenter.
Infinite dimensional convexity
613
See [2, Section II.4] for an exposition, where the simplified proof is attributed to J.-E. Bj6rk. The theory exposed in this section has been extended to the setting of weakly complete convex cones (not containing a line) in locally convex spaces. In this setting conical measures take the role which probability measures had in the compact case. For precise definitions of these concepts and a recent exposition of this theory we refer to [7]. A large part of this theory is a direct generalization of the compact case explained in this section. The book [7] contains several interesting applications of integral representation on cones (e.g., to potential theory and statistics). A generalization of the results in this section in a different direction will be presented in Section 4.
3. Compact simplices and related sets An n-dimensional convex set K in a linear space X is a simplex if it is the convex hull of n + 1 affinely independent points, or equivalently if it has n + 1 extreme points. There are several ways to define this notion of a finite-dimensional simplex so that the dimension n does not enter in the definition. One way to do this is to identify K with the set K x 1 in X @ R and to consider the cone P = {()~k,)~): k E K, )~ >~ 0} generated by this set. This cone defines in an obvious way a partial ordering on X G ~. It turns out (see Theorem 3.2 below) that a finite-dimensional convex set K is a simplex if and only if the partial ordering defined by P is a lattice ordering on P - P. Incidentally, the simple construction above shows that any convex set K is affinely equivalent to a base of convex cone P with vertex at the origin. (A base of a cone P is by definition a set of the form P n H with H a hyperplane not containing 0, the vertex of P.) DEFINITION 3.1. A convex set K in a linear space X is called a simplex if the cone P -{(,kk, ,k)" k E K, )~ ~> 0} defines a lattice order on P - P C X • ~ . Before continuing let us consider the following example. Let K be a compact convex set in a linear topological space and let P be the cone of all finite non-negative Baire measures on K. Clearly P defines a lattice order. Let Q be the subset of P consisting of maximal measures in the sense of Definition 2.13. The cone Q also defines a lattice order since from Proposition 2.16 it follows that if # and v are maximal measures so is v A #. The set Q1 of probability maximal measures on K is a base for Q and thus by the definition above it is a simplex. Note that Q1 is compact (in the w*-topology of measures) if and only if ext K is closed. The main result on compact simplices is the following. THEOREM 3.2 ([28,29,99]).
L e t K be a c o m p a c t convex set in a locally convex space X . The f o l l o w i n g three assertions are equivalent: (i) For each x E K there is a unique m a x i m a l probability measure IZx on K so that lZx ~ x.
614
V.P. Fonf et al.
(ii) K is a simplex. (iii) For every 0 < ~, ~ and every x, y E X such that ( a K + x) M ( ~ K + y) ~ 0 there is a 0 <. y anda z E X so that ( a K + x) M ( ~ K + y) = y K + z. There are many expositions of the proof of T h e o r e m 3.2 in the literature. We shall present here only two (of the simpler) implications in the theorem. (i) =~ (ii). Let Q1 be the maximal probability measures on K. If (i) holds, then the resultant map # ~ r ( # ) is a one-to-one affine map from Q1 onto K. Since Q1 is a simplex the same is true for K. (ii) =~ (iii). We assume that K is a base of a cone P, i.e., K = P A f - l ( 1 ) for some linear functional f , and that P induces a lattice order on P - P. Without loss of generality we m a y assume that y = 0 and fl = 1 in (iii). Let z E (x § otK) M K for some ot > 0 and x E P - P; i.e., z = x § c~kl = k2 with ki, k2 E K. Clearly, by a standard lattice identity z = x v 0 + otkl A k2 = x + + yk3, for some k3 6 K and for y = f ( z ) - f ( x +) = 1 f (x +) ~> 0. Let k be any point in K. We have that x + + y k ~> 0 and f (x + + y k) = f ( z ) - y f ( k 3 ) + y f ( k ) = 1 and hence x + + y k 6 K. In a similar way we check that x + + y k E x + otK and hence (x + otK) M K D x + + V K. Since any z as above belongs to x + § y K we deduce that (x § c~K) M K = x + § y K as desired. The proof of (ii) =~ (i) can be found in [33], [2, Section 11.3] or [136, Section 9]. As a byproduct of this proof one gets another characterization of compact simplices, in terms of upper envelopes: A compact convex set K is a simplex if and only if for every convex continuous function f on K the upper envelope f is affine (but not necessarily continuous). The proof of (iii) =~ (i) can be found in [99] and [151 ]. Note that in the proof of (ii) =:~ (iii) we did not use the compactness of K. It is proved in [99] (cf. also [ 151]) that any simplex intersects any line in a compact set (i.e., is a line compact set) and that for line compact sets (ii) and (iii) are equivalent. Also, if K is compact then (iii) is equivalent to the weakened version in which we assume c~ = fl = 1. The following is an obvious corollary. m
COROLLARY 3.3. A closed face of a compact simplex is again a simplex. In this connection we mention that it is easy to see (cf. [ 111 ]) that every compact metric convex set can be represented as a section of a compact metric simplex with an affine subspace. Let us prove now another property of compact simplices which is an easy consequence of T h e o r e m 3.2. PROPOSITION 3.4 ([15]). simplices is a simplex.
The intersection of a directed (downward)family of compact
A
PROOF. Let )U be the directed farn~ly of simplices and let K = A K EK: K. x, y E X and assume that (x + a K ) M (y + ilK) ~ 0. Then for every K and a 0 ~< VK ~< 1 so that (x + ot K) A (y + fl K) = ZK + YK K. Take any K C Ko, ZK belongs to the compact set x + otK0 - [0, 1]K0. Thus the
Let 0 < or, 13 ~< 1, E K~ there is a ZK K0 E )U. Then for set {ZK, YK}KEE
Infinite dimensional convexity
615 A
A
has a cluster point (~, 9?). SinceA for every K E/C we haveA (x + otK) n (y + ilK) C ZK + y/( K, it follows that (x + otK) N (y +/3 K) C ~ + 9?K. Similarly since~,,for every K~ ~/C, (x + ot K.) n (y +/3 K~ D z K -Jr- YK K we deduce that (x + c~K) n (y +/3 K) D z + 17K, i.e., (x + otK) N (y -Jr ilK) = ~ + ~'K. D In the w*-topology, the set of probability measures on a compact space F is clearly a simplex K whose set of extreme points is homeomorphic to F. (Any point x 6 F corresponds to the Dirac measure ex in K.) It follows trivially from (ii) =~ (i) of Theorem 3.2 and the fact that for sets K with ext K closed the maximal measures are just the measures supported on ext K, that the converse is true: If K is a simplex with ext K closed then K is (affinely equivalent to) the set of probability measures on ext K in the w*-topology. Simplices K with ext K closed are called Bauer simplices. The three examples we mentioned in Section 2 (completely monotonic, infinitely divisible and positive definite functions) are all Bauer simplices. In a certain sense, every compact metric simplex can be represented by Bauer simplices. PROPOSITION 3.5 ([52]). If S is a compact metric simplex there exists a decreasing sequence S1 3 $2 D ... 3 Sn 3 "" of Bauer simplices such that S is affinely homeomorphic to ASh. The proof of this proposition uses the structure Theorem 3.22 (below). From the structural point of view the Bauer simplices are simple objects with a very transparent structure. More interesting are simplices K with ext K non-compact. The question of what ext K can be in general is answered in the metrizable case by THEOREM 3.6 ([82]). For every Polish (i.e., complete separable metric) space M there is a compact metric simplex K so that ext K is homeomorphic to M. Note that by Proposition 2.9, for every separable compact convex set K the set ext K is a G~ subset of K and thus homeomorphic to a Polish space. It should be mentioned that unlike the case when M is compact, in general M does not determine the simplex K. In other words, two compact metric simplices which are not affinely equivalent may have homeomorphic sets of extreme points (see, e.g., the discussion following Definition 3.14). The proof of Theorem 3.6 is also presented in [4, p. 126]. We pass now to an important class of concrete examples of compact simplices. Let (X2,/3) be a measurable space and 7- a set of measurable maps from S-2 to itself. A probability measure/z on on/3 is called 7--invariant if/z ( T - 1A) = # (A) for every A 6 B and T 6 T. For a given invariant probability measure/z we say that a set A 6/3 is #-invariant i f l z ( A A T - 1 A ) = 0 for each T ET-. (By definition, Al AA2 is (A1 \ A 2 ) U ( A z \ A 1 ) , t h e symmetric difference between A1 and A2.) An invariant probability measure/z is called ergodic if the only #-invariant sets A are trivial (i.e., #(A) = 0 or/z(A) = 1). PROPOSITION 3.7 ([55]). With the notation as above, the set S of T-invariantprobability measures is a simplex and its extreme points are exactly the ergodic measures. Moreover, if I2 is a compact Hausdorff space, 13 is the Baire a-algebra and every T ~ 7- is continuous, then the simplex S is compact (in the w*-topology).
616
V.P. Fonf et al.
PROOF. It is straightforward to check that if/z and v are 7--invariant measures so is/z A v and thus S is a simplex. If/z is not ergodic let A be a/z-invariant set so that 0 < # (A) < 1. Put/z 1 ( B ) = / z (A M B) / # ( A ) and # 2 ( B ) = / z ( B \ A)/(1 - #(A)). One can verify that #1,/z2 E S and /z = /z(A)/z! -k- (1 - # ( A ) ) # 2 , i.e., # r extS. Conversely, assume that # 6 S is ergodic. It is easy to check that every measurable real valued function f on 12 which is 7--invariant (i.e., f(co) = f ( T c o ) a.e./z for every T 6 7-) must be constant. If/z = (/Zl +/1,2)/2 with //~1, # 2 E S, let f be the Radon-Nikod3~m derivative d # l / d # . This f must be 7--invariant and thus a constant, i.e., f = 1 a.e. # (since/zl and # are probability measures) and hence # = #1 = #2. The compactness assertion is obvious. M There are many theorems in the literature which state that, under some conditions on 12 and 7-, every invariant probability measure on I2 has a unique representation as an "integral average" of ergodic measures (see, e.g., [149]). The following theorem is an immediate consequence of Theorem 2.17 and Proposition 3.7. T H E O R E M 3.8. I f 1-2 is a compact Hausdorff space and 7- a family o f continuous mappings from Y2 to 1-2, then to each lZ E M (the set o f all 7--invariant probability measures on Y2 ), there exists a unique probability measure v on the Baire subsets o f M such that
# ( f ) -- fM m ( f ) d r ( m ) f o r each f E C(12) and v ( B ) = 0 f o r each Baire subset o f M which contains no ergodic measures. I f the ergodic measures f o r m a Baire subset o f M (e.g., if 1-2 is metrizable), then the measure v is supported on the ergodic measures.
The special case where 12 is compact metric and T a group was obtained earlier in [79] by using direct integral decompositions of unitary representations. A different approach to obtaining results similar to Theorem 3.8 was suggested in [42], making use of the concept of sufficient statistics. We present an interesting example where Proposition 3.7 produces a Bauer simplex. Let D be a compact metric space, 12 the product space D N ( N the positive integers) and 7the group of automorphisms of 1-2 induced by those permutations of the integers which shift only finitely many integers. In probabilistic language, the 7--invariant measures on D N are those measures which give the distribution of a sequence of D-valued exchangeable random variables. It was proved by Hewitt and Savage [87] that in this case the extreme (i.e., ergodic) measures are exactly the product m e a s u r e s o - u where a is a probability measure on D (see [131, pp. 150-151], for a simple proof). In probabilistic language, this means that the extreme invariant measures are those which correspond to a sequence of D-valued independent and equally distributed random variables. The correspondence (7 ~ O"N is a homeomorphism from the compact set Z7 of probability measures on D (with the usual w*-topology) onto ext K, where K is the simplex of invariant probability measures on D N . Thus, K is a metrizable Bauer simplex. It should be pointed out that for the identification above of ext K it is essential that we consider the infinite product D N.
Infinite dimensional convexity
617
(For the finite product D k of D with 1 < k < cx~ it is easy to see that ~rk is ergodic only if cr consists of a single atom.) Consider now the special case where D consists of just two points {0, 1 }. Then Z can be identified with the interval [0, 1]. In this case, the formula in T h e o r e m 3.8 says the following: For every invariant probability measure # on D N there is a unique probability measure v on [0, 1] for which /z(Ak,n) -- f01 t k (1 - t) "-k d r ( t ) ,
l <~ k <<.n < e~,
where Ak,n C {0, 1} N is defined by Ak,n -- {0 -- (01,02
....
)"
Oi = O, 1 <<.i <<.k; Oi -- 1, k + 1 <<.i <~ n } .
This special case of T h e o r e m 3.8 is a well known result of de Finetti, see [56, p. 225]. Let us mention a simple example where Proposition 3.7 yields a compact metric simplex with non-closed extreme points. Let I = [0, 1] and let J be the unit circle, represented as ~ ( m o d 1). Take Y2 = I x J with the product topology and let T consists of the single (obviously continuous) map T (t, s) = (t, s § t). It is readily checked that for every n, the measure #n on I-2 which gives mass 1/n to the n points (n -1 , k n - 1 ) , 0 ~< k ~< n - 1, is an ergodic measure. The sequence {#n} is w*-convergent to a measure # which is concentrated on {0} x J (it is the Haar measure on J ) which is not ergodic. In this connection, we have the following general result, the proof of which makes use of Proposition 3.5 (above). THEOREM 3.9 ([43]). For every compact metric simplex K there are a compact metric space I-2 and a h o m e o m o r p h i s m T o f I2 onto itself so that the set o f T-invariantprobability measures on S-2 is affinely h o m e o m o r p h i c to K. We shall return to the subject of simplices of invariant measures after first discussing the structure of a particularly interesting simplex. There is a compact metrizable simplex S for which ext S is dense in S. A very natural way to construct such an S is the following (cf. [141]). Let X = 12, En be the span of the first n unit vectors in X, Pn the natural projection from X onto En, n ---- 1,2 . . . . . and K = {x c / 2 : [Xn ] ~< 1/n, n = 1, 2 . . . . }. It is easy to construct inductively a sequence {Sn } of n-dimensional simplices so that (i) Sn C En A K for all n. (ii) S, C Sm, and extSn C extSm for n < m. (iii) Pn S m = Sn for n < m. (iv) For every e > 0 and n there is an m > n so that every point in S, is of distance at most e from ext Sm. The desired S is cl ~ n= 1 Sn. It is easy to check that S is convex and compact and that ext S is dense in S. That S is a simplex follows from the representation S - f"]n~__1 Pn-- 1 (Sn) and from an argument similar to the one used in the proof of Proposition 3.4. We shall see below other ways to construct such an S but the process just discussed is the most intuitive one from the geometric point of view. It is clear that the construction above is not at
V.P Fonf et al.
618
all rigid. There are many possible ways to choose the Sn's (and of course there are other methods of construction). Nevertheless, the surprising fact is that the outcome is always the same. This unique simplex is called the Poulsen simplex and it has many remarkable properties. THEOREM 3.10 ([ 122]). (i) Uniqueness: There is, up to affine equivalence, a unique compact metric simplex S with cl ext S - S (the Poulsen simplex). (ii) Universality: Every compact metric simplex is affinely homeomorphic to a face of
the Poulsen simplex. (iii) Homogeneity: For any two extreme points Sl and s2 of the Poulsen simplex there
is an affine automorphism of the simplex which carries Sl to s2. More generally, if F1 and F2 are two closed proper faces of the Poulsen simplex and if ~p is an affine homeomorphism from F1 onto F2 then ~pcan be extended to an affine automorphism of the Poulsen simplex. The proof of this theorem uses the structure Theorem 3.22 (below) and the tools which enter into its proof. The proof is also presented in [4, Section 3.7]. Since the Poulsen simplex S is uniquely determined, the same must be true for ext S. It turns out that ext S is homeomorphic to II~n0 (in the usual product topology). More precisely, there is a homeomorphism of the Poulsen simplex S onto [0, 1]n0 which carries ext S onto (0, 1)n0. We outline next the proof of the fact that if Z is the set of all the integers and T is the natural shift map of Y2 -- [0, 1 }z onto itself then the T-invariant probability measures
on Y2 form the Poulsen simplex. Let # be a T-invariant probability measure on ~2. For every subset F C Z we can identify {0, 1)F as the subset of {0, 1)z consisting of those functions which depend only on the coordinates belonging to F . Thus for every F , # induces a probability measure # r on {0, 1) F by the formula fs2 f d # - f{0,1lr f d # F for every f E C({0, 1 ) r ) . Let N be a positive integer; then the sets { F j ) j _ _ ~ form a partition of Z, where
Fj - - {iN, j N + 1 , . . , (j + 1)N - - 1) . Clearly Y2 -- {0, 1} z - - U j =~_ c x ) { 0 , 1}Fj Let # N be the probability measure on $2 which is the infinite product of the measures {# rj }j _ _ ~ . From the invariance of # and the definition of #N it is trivially checked that #N is invariant under T N, i.e., #U ( T U A) -- #U ( A ) for every A. In order to get a measure which is invariant under T we define a measure vu by N-1
vN(A) - N -1 Z
#N(TnA)"
n:O
We next show that each invariant measure VN is ergodic and that {VN }N=l ~ converges w* to #, i.e., the ergodic measures are dense in the set of all invariant measures. To prove that w*-lim vu -- #, note that if f is a continuous function on ~ which depends only on the coordinates { - k , - k + 1 . . . . . k - 1, k}, then
fs~ f (Tnc~ d#N - f~ f ( c o ) d # ,
k <~n <~N - k.
619
Infinite dimensional convexity
Hence
n--O
+
t]
E
n--N-k+l
and thus 4k
fs2 f (co) dvN -- fs2 f (o9) d# ~< -~-Ilfl[~ ---> 0,
N ---~ oo.
Since, by the Stone-Weierstrass theorem, the functions which depend only on finitely many coordinates are dense in C(S-2), this proves that indeed Vu --+ #. The fact that VN is ergodic for every N follows from the fact that for every continuous function f on $2 which depends only on finitely many coordinates mN-1
(mN)-' E n=0
N-1
f(Tnog) = N - ' E j =0
m-1
m-' E
f(TkN+jog) --+ J~ f (og) dvN,
k=0
m --+ oo for UN almost every o9 in S2. We omit the verification of this fact. The same proof which was outlined above shows that if I is a positive integer and I21 = {0, 1 }zl, then the probability measures on $21 which are invariant under the action of the group Z 1 form the Poulsen simplex (obviously Z l acts on S-21by translation). This fact was apparently first noted by Ruelle [ 15 3] in the framework of investigating models in classical statistical mechanics. The preceding results are a special case of a general theorem. In order to explain this theorem we have to introduce an important and well-known notion from representation theory. Let G be a (discrete) group and let 7r be a unitary representation of G on a Hilbert space H (i.e., Jr is a homomorphism from G into the group of unitary operators on H). We say that Jr has almost invariant vectors if for every e > 0 and every finite set F C G there is a unit vector x e H so that lirr(g)x - xll ~< e for every g e F. The group G is said to have property T of Kazhdan if every representation Jr of G which has almost invariant vectors actually has a nonzero vector which is invariant with respect to Jr (g) for all g e G. Commutative groups and more generally amenable groups fail to have property T. On the other hand, SL(n, Z) with n ~> 3 has property T. We can now state the following dichotomy result. THEOREM 3.1 1 ([76]). Let G be a countable discrete group. (i) If G has property T, then for any action of G on a compact metric space I2 the set
of G-invariant measures on 1-2 (if nonempty) is a Bauer simplex. (ii) If G does not have property T, then for the natural action of G on the compact metric space S-2- {0, 1}G the set of G-invariant measures on 12 is the Poulsen
simplex.
620
V.P Fonf et al.
By an action of G on a space s we mean a homomorphism from G into the group of homeomorphisms of S-2. In the paper [76] an analogous result is also proved for actions of topological (nondiscrete) groups. The Poulsen simplex arises also as the set of invariant states for many examples of quantum spin systems. Mathematically, this involves the study of G-invariant states of a C*-algebra. For a presentation of these results we refer to [153,23,161] and [162]. An interesting and useful property of compact simplices is exhibited in the following separation result. PROPOSITION 3.12 ([51]). Let K be a compact simplex and let f and - g be convex upper semicontinuous real valued functions on K with f (k) <<.g(k) f o r every k E K. Then there is an affine continuous function h on K so that f (k) <<.h(k) <<.g(k) f o r every k E K. The (quite simple) proof can be also found in [4, p. 72] or [2, p. 90]. COROLLARY 3.13. (i) Let K be a compact simplex and F a closed face o f K. Then f o r every ko E K \ F there is a nonnegative h E A ( K ) so that hlF =- 0 and h(ko) > O. (ii) I f F is a closed G~ face o f a compact simplex K then there is a nonnegative h E A ( K ) so that F = {k ~ K: h(k) = 0}. (Note that every closed face is a G~ set if K is metrizable.) PROOF. (i) By the separation theorem, there exists a ~o E A ( K ) with ~o(k0) > 0 and ~0]F ~< 0. Let f ( k ) = max{~0(k), 0}, k 6 K, and g(k) = 0 if k E F while g(k) = I1~o11~ for k E K \ F. Then f and g satisfy the assumptions of Proposition 3.12. The h we get from this proposition satisfies (i). (ii) From (i) it follows that for every closed (and thus compact) subset D of K which is disjoint from F there is a nonnegative hl) E A(K) of norm 1 so that hD IF -- 0 and hD(k) > 0 for every k 6 D. If K \ F -- Un~ On then the function h - ~n~=t 2-nhon has the desired property. D DEFINITION 3.14. A compact simplex K is called prime if whenever F1 and F2 are two closed G~ faces of K for which K = co(F1 U F2), then either K = F1 or K = F2. Obviously the Poulsen simplex S is prime. Indeed, ext S C F1 U F2 and thus since ext S is dense in S it follows S = F1 U/72, hence either F1 or F2 is equal to S. There are however many other, much simpler examples of prime simplices. For instance, consider the space X of all sequences of reals: x = (x l, x2 . . . . ) for which limi xi = oo )--~i=1 2 - i x i , with the supremum norm. Then X is isometric to A ( K ) for some compact simplex K. The simplest way to see this is to observe that X* is isometric to l l and to apply Proposition 3.23(ii) below. The simplex K consists of the nonnegative elements of X* which are equal to 1 at the sequence (1, 1, 1 . . . . ). The extreme points of K are the evaluation functionals e i (x) -- xi, 1 <, i <, oo. The sequence {e*}i~ is discrete in the w* topology (i.e. , homeomorphic to the set of positive integers) and w*-lim e* = ~i~=1 2 - i e i9 .
Infinite dimensional convexity
621
The simplex K is prime. Indeed, if K -- co(F1 U F2) then either F1 or F2 contains infinitely many of the {e*}~_ 1_ and thus also ~--~iz~1 2 -i e*. A closed face which contains Y~/=I~ 2-i ei* must coincide with K itself. More generally, if ~, -- (~,1, ~,2 . . . . ) is any sequence of nonnegative numbers whose sum is one then Xz, the space of all sequences x so that limi xi -- Zi~176 ~,ixi, is the space A ( K z ) for some compact simplex Kz. The simplex Kz is prime if and only if ~i > 0 for every i. The set ext Kz is discrete (but not closed) in the w* topology if and only if ~i > 0 for at least two indices i. It is trivial to check that Kz is affinely equivalent to Kz, if and only if )t is obtained from )~ by a permutation of the positive integers. In particular, we get many examples of non affinely equivalent compact simplices whose sets of extreme points are all homeomorphic to the positive integers. It is easy to characterize prime simplices K in terms of the order structure of A ( K ) . First, for comparison, we mention PROPOSITION 3.15 ([6]). Let K be a compact simplex. Then the following are equivalent: (i) K is a Bauer simplex. (ii) A (K) (in its natural order) is a lattice. If (i) holds then A ( K ) is order isometric to C ( e x t K ) which is clearly a lattice. Also (ii) =:> (i) is simple, see, e.g., [2, p. 101]. This implication follows also from Proposition 3.23(ii) below. PROPOSITION 3.16 ([54]). Let K be a compact simplex. Then the following are equivalent: (i) K is a prime simplex. (ii) A ( K ) is an antilattice in its natural order This means that whenever f, g ~ A ( K ) and f A g exists in A (K) then either f <~ g or g <~ f . PROOF. (ii) = , (i). Assume that K = co(F1 U F2), with F1 and F2 proper closed faces. Let hi, i = 1, 2, be nonnegative elements in A ( K ) which do not vanish identically and for which hi ]Fi ----O. Clearly, if h i / x h2 exists it must be 0. Hence, since A ( K ) is an antilattice, either h l = 0 or h2 = 0; a contradiction. (i) = , (ii) To prove that (ii) holds it suffices to prove that if h = f / x 0 is in A ( K ) , then either h = 0 or h = f . We first show that h(k) = m i n { f ( k ) , 0} for each k 6 ext K. Indeed, suppose there were k 6 e x t K and r 6 R such that h(k) < r < m i n { f ( k ) , 0}. Applying Proposition 3.12 to the concave continuous function gl = min{f, 0} and the convex upper semicontinuous function fl defined by fl (k) = r and f l = h elsewhere, we obtain an affine continuous function h l 7~ h such that h ~< h l ~< min{f, 0}, a contradiction. Now let Fj = {k: h(k) = f ( k ) } and F2 = {k: h(k) = 0}. Obviously, F1 and F2 are closed G~ faces of K and we have shown that ext K C F1 U F2. Since we assume (i) it follows that h = 0 or h = f , as desired. [] Before we continue, one comment on the term "prime". There exists a quite extensive study of order ideals in the spaces A (K) with K a compact simplex, and the term prime is related to this study. The reader may find details about this in [4] and [2].
622
V.P. Fonf et al.
We conclude the discussion of prime simplices by presenting a dichotomy theorem related to differential equations. Let G be a bounded domain in IKn. The classical Dirichlet problem is the following: Given a continuous f on OG find a harmonic function h on G which is continuous on G for which h = f on OG. If OG is smooth this problem is always solvable. However, in general there may not be a solution. In this connection one defines in potential theory a notion of "good" boundary points (regular points) and "bad" ones. THEOREM 3.17 ([54]). Let G be a connected and bounded open set in R n and let D C O G be the set o f regular points. (i) I f D is closed, then the space o f all harmonic functions on G which extend to a continuous function on G (with the supremum norm) is order isometric to A ( K ) f o r some Bauer simplex K. (ii) I f D is not closed, then the space o f all harmonic functions on G which extend to a continuous function on G (with the supremum norm) is order isometric to A ( K ) f o r some prime simplex K. In [54] there is also a study of related problems where the Laplace equation is replaced by other elliptic (even parabolic) differential equations. We turn next to an important selection theorem. Let K be a convex set and let X be a linear space. A set-valued map ~0 from K to X is called convex if each qg(k) is a nonempty convex subset of X and if ~,qg(kl) -q- (1 - )0~0(k2) C ~00~kl + (1 - )0k2) for every kl, k2 6 K and 0 ~< )~ ~< 1. A selection h of q9 is a function h : K --+ X so that h(k) ~ qg(k) for every k 6 K. The role of simplices in this connection becomes clear if we note that if K is a finite-dimensional simplex, then every convex set-valued map ~0 on K admits an affine selection. Indeed, let {ki}in=o be the extreme points of K. For each i choose any xi ~ qg(ki) and define the affine function h on K by h(~in=_o)~iki) -- Y~4n_o~,ixi, ~,i ~ O, ~-'~in=o~.i -- 1. This procedure obviously does not work if K is not a simplex and easy examples show that affine selections of convex set-valued functions may fail to exist if K is not a simplex (e.g., a square in the plane). In the infinite-dimensional case we are interested of course in continuous affine selections. This forces us to make also some topological assumptions on qg. The resulting selection theorem on simplices proved below reduces to the well-known selection theorem of Michael if the simplex is a Bauer simplex. (To be precise, it yields Michael's theorem only for maps defined on a compact s e t - the extreme points of the simplex - while in Michael's theorem the domain is allowed to be a more general topological space, namely a paracompact space.) For the formulation of the theorem we need one more concept. A set-valued map q9 from a topological space K into another space X is called lower semicontinuous (1.s.c.) if {k E K: ~0(k) N G # 0} is open in K for every open set G C X. THEOREM 3.18 ([108]). Let K be a compact simplex and let ~o be a convex l.s.c, map from K to the subsets o f a Banach space X. Assume moreover that qg(k) is a closed subset o f X f o r every k ~ K. Then q9 admits a continuous affine selection. Note that when X -- •, Theorem 3.18 is exactly Proposition 3.12. This special case will be used in the proof of the general case. The proof we give is taken from [80] and is due to L6ger. The main step of the proof is the following result on approximate selection.
Infinite dimensional convexity
623
LEMMA 3.1 9. Let K, q) and X be as in the statement of Theorem 3.18 but without the assumption that qg(k) is closed f o r every k. Then for every e > 0 there is a continuous affine h : K ~ X so that d(h(k), 99(k)) < e f o r all k ~ K. It easy to see how L e m m a 3.19 implies T h e o r e m 3.18. We have just to note that if h is continuous and affine with d(h(k), qg(k)) < e then the map k -+ ~p(k) n q)(k) is also convex and 1.s.c. where O(k) = {x 6 X: Ilx - h(k)ll < e}. Thus by L e m m a 3.19 we can construct inductively affine continuous maps hn : K --+ X so that d(hl (k), qg(k)) < 1/2 for k 6 K and for n > 1
d(hn(k), q)(k)N {x" Ilx - h n - ,
(k)ll ~
z - n + ' }) < z - n ,
k 6 K.
The sequence {hn}n~ converges uniformly to an affine continuous selection for q). (This last step uses the fact that X is complete and that q)(k) is closed for every k.) PROOF OF LEMMA 3.19. We prove the l e m m a first for finite-dimensional X by induction on the dimension. If d i m X = 1 the l e m m a follows from Proposition 3.12. Now let d i m X = n + 1 and let P be a continuous projection from X onto a o n e - d i m e n s i o n a l subspace. The map from K to the subsets of P X defined by k ~ Pq)(k) is a 1.s.c. convex map. Hence there is an affine continuous g: K -+ P X so that d(g(k), Pq)(k)) < e / 2 for all k ~ K. The map
k --+ lp(k) - q)(k) N {x ~ X" 11Px - g(k)II < e/2} is also a 1.s.c. convex map. By the induction hypothesis there is an affine continuous map f : K ~ (I - P ) X so that d ( f ( k ) , (I - P)gr(k)) < e / 2 for every k. It is easy to check that the continuous affine map h : K --+ X defined by h = f + g satisfies d(h(k), ~0(k)) < e for every k e K. We now pass to the case where X is a general Banach space. By the definition of a 1.s.c. map it follows that for every x e X the set Ux = {k: d(x, qg(k)) < e/2} is an open (perhaps empty) subset of K. Since K is compact there are {xi }i--1 n SO that K = uin_=l Uxi. Let 2; be the standard simplex 2J = {,k: )~i ~> 0, Y~i~z )~i = 1} in R n . Define 7r from K to the subsets of IR~ by
and take in IR~ any n o r m (say the l~ norm). Then it is easy to verify that ~p is a 1.s.c. convex map. Hence there is an affine continuous map g(k) = (gl ( k ) , . . . , gn(k)) from K into 1R~ so that d(g(k), O(k)) < 6 for every k, where 6 = e / ( 2 max/IIxi I[). The continuous affine map h(k) -- ~_~in=l gi(k)xi satisfies d(h(k), q)(k)) < e for every k e K. [3 COROLLARY 3.20. Let K be a compact metrizable simplex. Let F be a closed face of K. Then there is a continuous affine retraction from K onto F. PROOF. We can consider K as a compact set in a Banach space. The set-valued map defined by ~0(k) = k if k e F and ~0(k)= F if k ~ F is 1.s.c. and convex. Any affine continuous selection of this ~0 is an affine continuous retraction onto F. [3
624
V.P. Fonf et al.
REMARK. The corollary and its proof are valid if we just assume that F is metrizable and K a general compact simplex. Without the assumption that F is metrizable the corollary may fail to hold (see, e.g., [37]). Another consequence of Theorem 3.18 is that norm compact simplices behave well in connection with the approximation property. PROPOSITION 3.21 ([111]). Let X be a Banach space and let K be a (norm) compact simplex in X. Then for every s > 0 there is a bounded linear operator T from X into itself with d i m T X < cx~ so that IITx - x [ I ~< s for every x E K. In the paper [ 111 ] it is also pointed out that, in general, norm compact simplices are rather "small" subsets of a Banach space. More precisely, if X is an infinite-dimensional Banach space, then there is a norm compact subset K of X for which there does not exist even a norm bounded simplex S in X which contains K. In this connection it is worthwhile to mention the following result of Christensen [34].
Every compasct set in a Banach space is contained in the convex hull of two Bauer simplices. The proof of this statement runs as follows. A classical result of Grothendieck states that any norm compact set in a Banach space is contained in the closed convex hull of a sequence {Xn}n~_l tending to 0. A minor modification in the usual proof of this result shows that the vectors {Xn }~=l can be chosen to be also linearly independent. A sequence {Xn}n~_l _ is called strongly independent if whenever E~=~ IZnl < ~ and y~'n~__l),.nXn- 0 then )~n = 0 for all n. Clearly, if {Xn}n~=l is a strongly independent sequence which tends to 0 then {y~n~=l )~nXn" ~.n ~> 0, n = 1,2 . . . . , )--~n~_-I•n ~< 1} is a Bauer simplex whose extreme points are {Xn}n~=l and 0. By using a category argument (on {0, 1}s~ = the set of all subsequences of the integers) Christensen shows that any linearly independent sequence {Xn}n~=l tending to 0 can be split into two subsequences {Xn,}n,~=l and {Xn,,}n,~,_l both of which are strongly independent. This concludes the proof. The following structure theorem for general compact metric simplices is proved by a more sophisticated application of the selection Theorem 3.18. THEOREM 3.22 ([113]). For every compactmetric simplex K there is a sequence {Sn }n~__l of n-dimensional simplices and surjective affine maps qgn "Sn+l --+ Sn so that K is the inverse limit of the system
S1 ~ 82 ~ $3 <-- "''.
(3.1)
Conversely every inverse limit of such a system is a compact metric simplex. This theorem gives a very explicit way of representing a general compact metric simplex. Indeed, any affine map qgn from Sn+l onto Sn must map n vertices of Sn+l to the n different vertices {l)i}in_l of Sn. The remaining vertex of Sn+l is mapped by ~0n onto Ein= 1 ai,n Vi where ai,n >/0 and Y~i= n 1 ai,n - 1 It follows that (up to an obvious identification) there is a one-to-one correspondence between systems (3.1) and the set of infinite matrices A - (ai,n) such that
Infinite dimensional convexity ai,n >/O,
~
ai,n -- 1,
ai,n
= 0
if/
625
> n.
i=1
Thus every such matrix A represents a definite compact metric simplex. The correspondence between the representing matrices and the compact simplices is, however, very far from being one-to-one. First, there is a stability result. If A and B are such matrices and if ~-,i,n lai,n - bi,n I < OC, then A and B represent the same simplex. Moreover, there are also completely different matrices which represent the same compact metric simplex. The correspondence between these matrices and the simplices they represent is investigated in some detail in [ 160]. Let us just mention that if A is such that its rows form a dense set in {)~: ~i~__l )~i - 1, )~i ~> 0}, then A represents the Poulsen simplex and we thus get yet another way of constructing the latter. Recall (see text after Definition 2.7) that any compact convex K can be identified canonically with a subset K of the unit ball of A(K)* and that the unit ball of A(K)* is co(K U - K ) . This is useful in the proof of the next proposition. PROPOSITION 3.23 ([109,119,157]). (i) The spaces A(Kl) and A(K2), with Kj and K2 compact simplices, are isometric if and only if K1 is affinely homeomorphic to K2. (ii) A Banach space X is isometric to a space A ( K ) for a compact simplex K if and only if X* is isometric to an L l(IZ) space and the unit ball of X has at least one extreme point. An interesting consequence of this proposition is that one can define a nice multiplication operation on simplices [109]. If K1 and K2 are compact simplices, consider the space X -- A (K1, K2) of continuous real-valued functions on K l • K2 which are affine in each of the two variables separately, with the supremum norm. Clearly the unit ball of X has an extreme point (the function identically equal to 1) and it is not hard to check that X* is isometric to an L1 (#) space. Hence, by Proposition 3.23 there is a unique compact simplex K such that X = A ( K ) . The simplex K is called the tensor product K1 | K2 of the simplices K j and K2. This product is evidently commutative and associative. Clearly ext(K1 | K2) is homeomorphic to ext K1 x ext K2. If K1 ----K2 = S (the Poulsen simplex), then it is easily checked that S @ S is not the Poulsen simplex. Nevertheless, ext(S @ S) is homeomorphic to R s0 x R s0 = R ~0, i.e., to ext S. It is clearly possible to define in the same way the tensor product of an arbitrary family of compact simplices. In view of Proposition 3.23 it is of interest to study general Banach spaces X such that X* is isometric to an L l(#) space. The unit balls of such spaces X can be considered as infinite-dimensional generalizations of finite-dimensional cubes. There are many characterizations of such spaces; we mention here only the following geometric one. THEOREM 3.24 ([119]). Let X be a Banach space. The following three assertions are equivalent: (i) X* is isometric to an L l(lz) space. (ii) Any finite family of mutually intersecting balls in X has a common point. (iii) Any family of four mutually intersecting balls of radius 1 in X has a common point.
V.P Fonf et al.
626
The number four appearing in (iii) above cannot be reduced to three. A classification of the concrete function spaces X for which X* is an L l ( # ) space is given in [123]. For further results related to Theorem 3.24, see [ 116]. The structure Theorem 3.22 has a version for spaces whose duals are L l ( # ) spaces. For every separable space X such that X* is an L l ( # ) space there is a chain E1--~ E2 ~ E3--+ .--,
(3.2)
where En is isometric to I n , u n ' E n --+ En+l is an isometry and X -- cl Un=l oc En. Conversely, any such chain (3.2) defines a space X whose dual is an L1 (#) space. With a suitable choice of the unit v e c t o r s {ei,n}in_=l in l ~ (we are free to choose signs and a permutation) we have Unei,n = ei,n+l + ai,nen+l,n+l for every i ~< n and ~in=l lai,nl <~ 1. Therefore, to every chain (3.2) there corresponds a unique representing matrix A = (ai,n) such that Y']~i-1 n lai,nl ~< 1 for every n and ai,n - - 0 for i > n. If the ai,n are non-negative and Ein__l ai,n -- 1, then the space represented by A is A ( K ) for some compact metric simplex K. (In this case, the chains (3.1) and (3.2) represent exactly the same situation but in a dual form.) There is a special space that plays the same role among the separable spaces X for which X* = L l ( # ) as the Poulsen simplex plays among all compact metrizable simplices. This space is called the Gurarii space. In [81] a separable Banach space is constructed which has a certain nice extension property for isometries on finite dimensional subspaces. This extension property implies in particular that X* is an L1 (/z) space (in view of a characterization of such spaces in [119] by extension properties for compact operators). The decisive step in investigating the Gurarii space was done in [ 126] and [127]. It was proved there that the Gurarii space is unique (up to isometry) and that it is characterized by the fact that it is a separable space X whose dual is an L l ( # ) space with the property that ext B x , is w* dense in B x , . Every separable X with X* = L1 (/z) is isometric to a subspace of the Gurarii space on which there is a projection of norm 1. Finally, any matrix A = (ai,n) whose rows form a dense set in the unit ball of l l represents the Gurarii space. In all the above we considered only real Banach spaces X whose duals are L 1 ( # ) spaces. Naturally, one can look also at the complex Banach spaces whose duals are L1 spaces. Many results go over from the real to the complex case while some have to be modified somewhat. The situation is not obvious and has been investigated in detail in many papers (see, e.g., [ 115,88,92,53] and also [4, Chapter 4]). We present here just one example where the complex case is different from the real one and this is related to Theorem 3.2. Let F be a compact metric space and let X be a closed subspace of the real space C ( F ) which contains 1 and separates the points of F. Let S be the state space of X, i.e., S = {x* e X*: IIx*ll-- x * ( 1 ) - 1}. With the obvious identification of F with elements of X*, we have ext S C F. By Theorem 3.2, each x* 6 S has a unique probability measure which represents it and is supported on B -- ext S iff S is a simplex. It was observed in [72] that similarly, each x* E B x . has a unique signed measure # supported on B which represents x* and for which IIx*ll = I1#11 iff S is a simplex. For the complex case we need first a new notion. DEFINITION 3.25. A CCB set K is said to be a simplexoid if every closed proper face of K is a simplex.
Infinite dimensional convexity
627
Thus, a compact simplex is a simplexoid but so are, for example, the unit balls of strictly convex spaces or the octahedron. With this terminology we can now state the following existence and uniqueness theorem. The existence portion is due to Hustad [91 ], the uniqueness to [72]. THEOREM 3.26. Let X be a closed subspace of the complex space Cc(F) of complex continuous functions on the compact metric space F. Assume that X contains 1 and separates thepoints o f F . Let B C F be the extreme points of the state space S = {x* E X*: IIx*ll = x*(1) = 1}. Then every x* ~ B x , has a (complex) measure lZ supported on B which represents it and and satisfies I1#11 = IIx* II. Moreover, this measure is unique if and only if Bx, is a simplexoid. A proof of this theorem can also be found in [4, Section 4.1 ] or [139]. The theorem also holds if F is just compact Hausdorff, but then the assumption that # is supported on B has to be replaced by requiring that I#l/ll#ll be a maximal measure on S (if # -r 0). We shall return to the setting appearing in Theorem 3.26 (in other contexts) in Section 5. We conclude this section by mentioning briefly a related subject on which much deep research has been done. Let A be a complex C* algebra (with an identity). Let S be the state space of A consisting of those functionals f 6 A* such that IIf l[ = 1 and f is positive (i.e., f ( x * x ) >~0 for every x E A). This is a w*-compact convex set. If A is commutative, then S is just a Bauer simplex. In the simplest noncommutative case, that of all operators on l~, S turns out to be the set of all positive semi-definite n x n matrices of trace 1. The main question here is the following: Given a compact convex set S, when can it be represented as a state space of a C* algebra? In other words what are the geometric conditions which characterize the state space of a C* algebra. A complete answer to this question is known. It is quite complicated to state and the proof is delicate. An important step in this characterization result is to characterize first a more general class of compact convex sets - the state spaces of JB algebras (Jordan-Banach algebras). All this material is explained in detail in the book [3] to which we refer the reader. The results mentioned above concerning the structure of general metrizable compact simplices have applications to the theory of C* algebras. In [ 167] all the Elliott invariants (a concept related to K-theory) of C* algebras have been determined. The main tool used in this determination is the following: PROPOSITION 3.27 ([167]). Let S and E be compact metric infinite-dimensional simplices and let )~: S --+ E be an affine continuous map so that )~(ext S) = ext ZT. Then there exist sequences of finite-dimensional simplices {Sn }n=l ~ and { Sn }n=l'~ affine surjective maps cbn :Sn+l --+ Sn, ~Pn : S,n+l --+ 27n and )~n : Sn --+ Zn so that S (respectively S ) is the inverse limit ofthe system {Sn, dPn}n~=i (respectively { r n , ~n}n~__l), )~n O~bn - ~n o Zn+l for every n and so that the given )~ is the inverse limit of the {)~n}n~__l. In the case where S = r and )~ the identity map, this is Theorem 3.22. The proof of Proposition 3.27 uses the same tools as that of Theorem 3.22 but, naturally, it is somewhat more involved.
628
V.P Fonf et al.
4. Exposed points and the Radon-Nikod~m property In this section we shall discuss extensions of the results of the previous sections in two main directions. One direction will be the generation of convex sets by special types of extreme points. The second direction will be to extend the results to some classes of convex sets which need not be compact (in any topology). We first bring the definition of certain classes of extreme points which will play a central role in this and the next section. DEFINITION 4.1. Let K be a convex set in a Banach space X. (i) A point x E K is an exposed point of K (x E exp K) if there is an f E X* such that f attains its maximum on K at x and only at x. The functional f is said to expose x. (ii) A point x E K is a strongly exposed point of K (x E str exp K) if there is a functional f which exposes x and so that for every e > 0 there is ~ > 0 so that y 6 K and I l y - xll/> 6 implies that f ( y ) <. f ( x ) - e. (iii) If X = Y* is a conjugate space then x is a w*-exposed point of K (x ~ w*-exp K) if the functional f appearing in (i) can be taken to be element in Y.
Clearly ext K D exp K D str exp K. There are simple examples of 2-dimensional convex compact sets such that the inclusion ext K D exp K is proper. For norm-compact convex K one easily checks that exp K = str exp K. Again easy examples show that for closed convex bounded K in infinite-dimensional spaces, the inclusion exp K D str exp K can be proper. Let us note that it follows from Corollary 3.13(ii) that for a norm compact simplex K in a Banach space, ext K = exp K. The notions defined in Definition 4.1 depend on the topology (or norm) of X. There are various variants of the definition which are independent of the topology. The weakest such notion is that of algebraic exposed point (a-exp K). The point x E K belongs to a-exp K if x E exp(K M B) for every 2-dimensional subspace B of X which contains x. Clearly ext K D a-exp K D exp K. We shall not discuss further in this chapter the notion of an algebraic exposed point. A notion obviously related to exposed points is defined in the following definition. DEFINITION 4.2. Let K be a convex set in a Banach space X. A point x E K is said to be a support point of K if there is an f E X* which attains its maximum on K at x. The functional f is said to support K at x. An f E X* is a support functional to K iff f attains its maximum on K. The set of all support functionals of K is denoted by r ( K ) . The notions of w*-supportpoint and w*-supportfunctional (in case X = Y*) are defined in an obvious manner. Clearly every exposed point (respectively w*-exposed point) of K is a support point (respectively w*-support point) of K. The converse is evidently false. The earliest result on exposed points is the following proposition.
Every norm-compact convex set K in a Banach space X is the closed convex hull of its exposed points. PROPOSITION 4.3 ([163]).
Infinite dimensional convexity
629
PROOF. We assume first that X is a Hilbert space. A point x 6 K is called a farthest point if there is a y 6 X so that Ily - x II >~ [ly - z[I for every z 6 K. Any farthest point is an exposed point. Indeed, if x is the farthest point from y then the ball with center y and radius IIY - x II contains K. The tangent hyperplane to this ball at x defines a functional which exposes x in K. We claim that the closed convex hull of all farthest points in K is equal to K. A s s u m e that this is false. Then by the separation theorem there are a point u 6 K (which we m a y assume to be 0), a v E X and a )~ > 0 so that for every farthest point xEK (v, x) >/)~ > (v, u) = 0. Then for t ~> 0 and x a farthest point in K Iltv - xll 2 = t2llvll 2 -+- Ilxll 2 - 2t (v, x ) ~< t 2 IIv II2
_+_ D 2 _
2ti.,
where D is the diameter of K. On the other hand [[tv -0[[ 2 = t2[[Vi[ 2 is for large enough t larger than t 2 [[vi[ 2 nt- D 2 - 2t)~. Hence for large enough t the farthest point to tv in K (which exists since K is compact) does not belong to the set of all farthest points, a contradiction. A s s u m e now that X is a general Banach space. We m a y clearly assume that X is separable. There is a linear b o u n d e d one-to-one operator T from X into 12. The restriction of T to K is an affine h o m e o m o r p h i s m . If x E K is such that T x E exp T (K) then x 6 exp K. Indeed if f exposes T x then T * f exposes x. Since by what we have already proved T (K) = cl co(exp T (K)) we deduce that K = cl co(exp K), as required. D There is an extension of this proposition to locally compact sets which is similar to the extension of the K r e i n - M i l m a n theorem (Theorem 1.6). THEOREM 4.4 ([102]). Let K be a closed convex and locally compact set in a Banach space which does not contain any line. Then K -- cl co(exp K U r exp K).
Here we denoted by r exp K the exposed rays in K (with the obvious definition). Proposition 4.3 is clearly equivalent to the statement that for compact convex sets K in a Banach space ext K C cl exp K. The Borelian structure of the set exp K is not simple even in finite-dimensional K. (Recall that ext K is a G~ set for n o r m compact K, by Proposition 2.9.) A study of the Borelian structure of exp K for compact K in finite or infinite-dimensional Banach spaces is carried out in [32]. We now pass to the question of when certain closed convex sets, which are neither compact nor locally compact, are generated by their extreme or exposed points. For this purpose we introduce the following definition. DEFINITION 4.5. Let C be a closed convex and b o u n d e d set in a Banach space X. The set C has the R a d o n - N i k o d y m property (RNP) if for every probability space (S-2,/3, # ) and
630
V.P. Fonf et al.
every X-valued countably additive measure r on/3 such that r ( A ) / l z ( A ) 6 C for every A ~/3 with #(A) > 0, there is a (Bochner) measurable f ' ~ 2 --+ X so that v ( A ) - fA f (co) d#(co),
AEB.
The space X is said to have the Radon-Nikod3~m property if its unit ball Bx has the RNE Clearly if X has the RNP, then every closed bounded and convex subset C of X has the RNE This notion for C = Bx was introduced (with an equivalent definition) in the basic concepts article [95]. The proof of the following theorem (most of the equivalences) was outlined in Section 7 of [95] in the case C = Bx. Its proof for general C is similar (for complete details and references to the original literature see Section 5.2 of [8]). THEOREM 4.6. Let C be a closed convex and bounded set in a Banach space X. Then the following assertions are equivalent: (1) C has the RNP. (2) For every bounded linear operator T ' L 1 (lz) --+ X such that Tdp ~ C f o r every non-negative r in L| (lZ) of norm 1, there is a g E Lc~(lZ, X) f o r which T f = f n f (co)g(co) dlz(co) f o r every f E L1 (/z). (3) Every martingale { fn }n~_l on some probability space (S2, 13, #) f o r which fn (co) E C f o r all n ~ 1 and co E S-2, converges a.e. (and in L l(#, X)) as n --+ cx~. (4) For every separable subspace Y of X, Y A C has the RNP. (5) Every function f ' [ 0 , 1] --+ X for which ( f ( s ) - f ( t ) ) / ( s - t) ~ C for all 0 <~ t < s <<,1, is differentiablefor almost all t E [0, 1]. The proof of the theorem shows that in (1) and (2) above it is enough to consider only the space [0, 1] with Lebesgue measure in place of (1-2,/3, #). Also it suffices to consider only martingales {fn }~--1 on [0, 1] for which fn is measurable with respect to the a-algebra whose atoms are [k2 -n, (k + 1)2-n], 0 ~< k <~ 2 n - 1. As was pointed out in [95] every separable conjugate space (and therefore every reflexive space) has the RNE Important and easy examples of spaces which fail to have the RNP are co and L 110, 1]. These facts together with the obvious fact that if X has the RNP then so do all its subspaces, gives a quite good picture of which spaces have the RNE To get more precise information on this question see [8] Section 5.4. As for convex sets, more general than unit balls, the most important examples of sets with the RNP are w-compact convex sets (see Proposition 4.18 below). In [95] Section 6 a proof is given of the following proposition for C = Bx. The proof in the general case is identical. PROPOSITION 4.7. Every closed convex set C with the RNP is the closed convex hull of its extreme points.
Infinite dimensional convexity
631
This is obviously a generalization of the Krein-Milman theorem to the setting of RNP sets. Also the stronger result - the Choquet integral representation t h e o r e m - discussed in Section 2 can be generalized to the setting of RNP sets (if C is separable). THEOREM 4.8 ([47]). Assume that C is a separable closed convex and bounded set with the RNP. Then for every x c C there is a Borel probability measure # on C so that # ( e x t C) = 1 and so that x = f c Y d # ( y ) . We note first that the existence of a barycenter for every probability measure/z on C is obvious. The integral f c Y d # (y) is just the Bochner integral. For every continuous linear functional f on the space containing C we have f ( f c Y d # ( y ) ) = f c f (Y) d # ( y ) and thus by the separation theorem f c Y d# (y) E C. That the assertion that # (ext C) -- 1 makes sense follows from LEMMA 4.9. For a separable closed convex set C in a Banach space X the set ext C is universally measurable, i.e., it belongs to the completion of the Borel or-field of C with respect to every Borel probability measure lZ on C. PROOF. Let D be the diagonal of C x C. The set C x C \ D is an open set in the complete metric space C x C and is therefore a Polish space. The map 4~ : (C x C \ D) --+ C defined by dp(x, y) = (x + y ) / 2 is continuous and the range of ~b is exactly C \ extC. Hence C \ ext C is an analytic set (a continuous image of a Polish space) and thus universally measurable. Consequently, ext C is universally measurable. [~ REMARK. The set ext C need not be a Borel set. In [94] there is an example of a closed convex and bounded set C in l l so that ext C fails to be a Borel set. Incidentally, the set ext C in this case is even not analytic (since C \ ext C is analytic and an analytic set whose complement, with respect to a complete metric space, is also analytic must be Borel). We are now going to prove Theorem 4.8. First some notation as a preparation. Let ~1 be the first uncountable ordinal. For every ot < ~1 (i.e., c~ a countable ordinal) consider the discrete set Bc~ = {0, 1 } and suppose that both 0 and 1 have measure 1/2. Let I2 = I-I~ <~1 Bc~ and for 13 < ~l let A~ be the a-algebra of universally measurable sets in I-I~
V.P. Fonf et al.
632
PROOF. Assume that f (co) 6 ext C a.e. and E ( g I A ~ ) = f for some g measurable with respect to some r215 with g >/3. Then for almost every co E 1-]c~<~ Bc~
f (w) -
f
(4.1)
g(co, cot) do'(co'),
where the integration is over I-[~ 0 where A = {co: f(co) r extC}. We shall apply now a standard selection theorem which asserts that if M1 is a Polish space, M2 a metric space and r a continuous map from M1 onto M2 then there is a measurable selection of 4~-1 as a map from M2 to M1 (taking in M1 the Borel (x-field and in M2 the cr-field of universally measurable sets). Taking Mt -- C x C \ D (D the diagonal) and M2 = C \ ext C it follows that there are measurable functions h l and h2 from C \ ext C into C so t h a t x = (hi(x) + h2(x))/2 and h i ( x ) ~ h2(x) for every x 6 C \ extC. We extend hi and h2 to measurable functions from C into itself by putting hi (x) = h2(x) -- x for x E ext C. Consider now the function g : t 2 --+ C which is measurable with respect to Ate+l, defined by g (co, O) = h l o f (co),
g(co, 1) = h2 o f(co),
coE
UBa.
Clearly E(g ]At~) = f but on the set A x {0, 1} g is different from f . Hence f is not maximal. [2 LEMMA 4.1 1. For every measurable f : t2 --+ C there is a measurable g : ~ --+ C so that
f -< g and g is maximal. PROOF. The lemma follows from Zorn's lemma. We shall show that every increasing sequence (with respect to -<) has an upper bound and every transfinite increasing chain is constant after a countable number of steps. Any increasing sequence {36 }~1 with respect to -< is a martingale. Since C has the RNP it follows from the equivalence of (1) and (3) in Theorem 4.6 that g = limi f/ exists a.e. Clearly f/-< g for every i. Assume now that {f• }y is any increasing chain and that f• is measurable with respect to At3• (Vl <, Y2 =~ fl• <, fl• Let A 6 A~ be any measurable set in s By the definition of the order, fa f• (co) dm (co) has the same value for all V such that fly >~~. Call this constant value r (A). We have thus defined a vector valued measure r on t2. By Definition 4.5 there is a measurable g : t 2 ~ C so that r ( A ) = fa g(co)dm(co) for every measurable A. The function g (as every measurable function on t2) is measurable with respect to A~ 0 for some or0. It follows that g = f• (a.e.) for every y with fly >~oto, as required. [2
Infinite dimensional convexity
633
It is now very easy to complete the proof of Theorem 4.8. Let x be any point in C and let f : Y 2 --+ C be the function identically equal to x (note that//~f is simply the Dirac measure ex). By L e m m a 4.11 there is a g >- f which is maximal with respect to this order. By L e m m a 4.10/Zg (ext C) = 1. We clearly have
f c y d/zg ( y ) -
fs2 g ( c o ) d m ( c o ) - fs2 f ( c o ) d m ( c o ) - - x .
The order we used in the proof above is the one introduced in [128]. One could have used also the order among probability measures defined in Definition 2.13 or the notion of dilation of measures (Definition 2.14). These orders were defined in Section 2 for compact convex sets. However they make sense for general closed convex and bounded C in separable Banach spaces. All these three orders coincide. This is proved in Chapter 6 of [21 ]. In this chapter one can find also a discussion of the non-separable case (where much more care is needed). Also the uniqueness theorem (Theorem 3.2) carries over to sets with the RNR THEOREM 4.12 ([22,154]). A closed convex and bounded set C with the RNP in a separable Banach space X is a simplex if and only if every x E C has a unique probability measure # on ext C with x = f y d # ( y ) . Note that Definition 3.1 of a simplex is purely algebraic so it applies also in the present context. It is of some interest to know whether Proposition 4.7 (or for that matter also Theorem 4.8) characterizes convex sets which have the RNR With that purpose in mind we introduce the following DEFINITION 4.13. A closed convex and bounded set C in a Banach space is said to have the Krein-Milman property (KMP) if for every closed convex subset C t of C we have C f = cl co(ext Ct). A Banach space X is said to have the KMP if B x has the KMR Proposition 4.7 states in this terminology that the RNP implies the KMR The question whether the converse is true has received a lot of attention. There are several partial results but the general question is still open. Here are some of the partial results. In [89] it is proved that if a dual Banach space X has the KMP then it also has the RNR In [20] the same result is proved if we assume that X is a Banach lattice (see also [21, Theorem 7.13.1]). Another result of this nature is proved in [155] under the assumption that X is isomorphic to its square X @ X. To state a result of a somewhat different nature we first point out that if D is any closed (not convex in general) subset of a closed bounded convex set with the RNP then D has an extreme point. The notion of an extreme point obviously makes sense also for nonconvex sets. The fact that such D always has an extreme point is not a direct consequence of Proposition 4.7. It will be proved below. It was shown in [90] that this property does indeed characterize the RNP: If every closed subset D of a closed convex and bounded set C has an extreme point, then C has the RNP.
634
V.P. Fonf et al.
We pass now to the main geometric properties which characterize the RNP but first some more definitions. DEFINITION 4.14. Let C be a closed convex bounded set in a Banach S ( C , x * , a ) (where x* 6 X* and ot > 0) of C is a set of the form {x SUpy~c x*(y) - or}. A point x 6 C is said to be a denting point of C if in slices of C of arbitrarily small diameter. A set C is called dentable if arbitrarily small diameter.
space. A slice 6 C: x*(x) > it is contained it has slices of
Clearly if C has a denting point it is dentable. It is also clear that a strongly exposed point of C is a denting point. However the notion of being strongly exposed is stronger than denting since in the definition of the former notion we require that the small slices are all determined by the same functional x* E X*. THEOREM 4.15 ([ 144,129]). A closed convex and bounded set has the RNP if and only if every closed convex subset of C is dentable. For a relatively simple proof of this theorem and additional references see Theorem 5.8 of [8]. It should be pointed out that the notion of dentability (as well as Theorem 4.15) was already used implicitly in an earlier part of this section. In fact, the proof of Proposition 4.7 (presented in [95]) used Theorem 4.15. The following important theorem is also derived from Theorem 4.15. It is much stronger than Proposition 4.7. THEOREM 4.16 ([17,138]). Let C be a closed convex bounded set with the RNP in a Banach space X. Then C is the closed convex hull of its strongly exposed points. The functionals which strongly expose points of C form a dense G~ set in X*. For a presentation of the proof see [8], Theorem 5.17. Note that there is no strengthening of Theorem 4.16 in the spirit of Theorem 4.8 above. An extreme point of C cannot in general be a barycenter of a measure sitting on the exposed (or strongly exposed) points. The only measure which represents an extreme point x is the Dirac measure e x. We can now justify a statement made above. Assume C is a closed convex bounded set with the RNP and D is a closed subset of C. Then every strongly exposed point in cl co D is easily seen to belong to D itself. Any such point is in particular an extreme point of D. The validity of the first assertion of Theorem 4.16 for every closed convex subset of a bounded closed convex set C clearly implies that C has the RNE There is also a strong converse to the second assertion of Theorem 4.16. This converse is due to Bourgain and Stegall (see [21, Theorem 3.5.5]). PROPOSITION 4.17. Let C be a non-dentable closed convex set in a separable Banach space X. Then the set S,(C) of all support functionals of C is of the first category in X*.
Infinite dimensional convexity
635
The first general result on existence of exposed points (strongly exposed points in this case) was Proposition 4.3 proved above. The first result on existence of strongly exposed points in a large class of non-norm compact convex sets was PROPOSITION 4.18 ([118,164]). Every w-compact convex set in a Banach space is the closed convex hull of its strongly exposed points. It follows from Proposition 4.18, in particular, that a w-compact convex set in a Banach space has the R N E There are however other ways (simpler and more direct) to prove this fact. We shall next use dentability to give a geometric proof of a structure theorem on Banach spaces having the R N E THEOREM 4.19 ([74]). Every infinite-dimensional Banach space X with the RNP has a subspace isometric to a separable infinite-dimensional conjugate space. PROOF. We shall present here the proof given in [75]. It is clear that there is no loss in generality in assuming that X is separable. We note first that by using dentability any convex set with the RNP can be decomposed into a countable number of convex sets with a small diameter. To be specific we work with the unit ball Bx of X. For every positive integer k we shall define inductively closed sets F~,k (0 <~ ot an ordinal) as follows F0,k -- Bx,
F~+ 1,k = Fu,k \ {x" ~bu,k(x) > 0 },
where q~c~,k is an affine function of the form ~ , k ( x ) -- x*,k + c~,k with
IIx~*,~II - 1 and is
chosen so that {x E F~,k: q~,k(x) > 0} is non-empty and of diameter smaller than 2 -k. Such a choice is possible as long as F~,k r 0, since every closed convex subset of Bx is dentable. For every V a limit ordinal we put Fy,k -- ~ <• F~,k. Thus, for every k we get a strictly decreasing chain of closed subsets of Bx. There must be a first ordinal fl for which F/~,k = 0. This ordinal 13 = fl(k) is countable since X is separable and hence Bx is Lindel6f. For each x E Bx and integer k we define an ordinal c~k(x) by the condition that x 9 F~k(x~,k A {y: ~b~k(x~,k(y) > 0} (in other words x 9 Fc~,k if and only if ot ~< otk(x)). For every positive integer j let Rj be a 2 - J - n e t in [ - 1 , 1]. We are next going to find a sequence {en }n~__l of unit vectors in X and finite subsets An C X* so that the following hold: (i) SUPx,ea n x*(x) >/[Ixll(1 - 2 -n) for every x 9 span{ei}in=l . (ii) x*(en+l) = 0 for x* 9 An. n ( n (iii) Otk(Y~i=l ziei -+-Zen+l) ~ 0tk ~i=1 ziei) for every 1 ~< k ~< n, every choice of zi 9 Ri and every z with Izl ~< 1. We choose as el any vector of norm 1. We then choose A1 so that (i) holds for n = 1. In general, having chosen {ei}in=l we choose An so that (i) holds for n. We show how to choose en+l once An is defined.
V.P Fonf et al.
636
We choose en+l of norm 1 so that (ii) holds and so that for each of the finite number of functionals X*
n
0~k (~--~i = 1 ziei),k'
l<.k<.n,
ziERi,
l<.i<.n,
we have
The right hand side of the previous inequality is positive by the choice of c~k(Ein_l zi ei). Since there are only finitely many linear inequalities it is clear that there is indeed an en+l of norm 1 which satisfies all of them. This completes the construction. Conditions (i) and (ii) clearly hold. As for condition (iii) note that by definition Ilx* (Ei~=I ziei), k II = 1 and hence for every z with Izl ~< 1,
which in turn implies (iii) by the definition of C~k. By a routine argument (see, e.g., [95, Section 3]) conditions (i) and (ii) imply that {ei }~--1 is a basic sequence. We shall show that (iii) implies that this basic sequence is boundedly complete and thus that [en]n~__l is isomorphic to a conjugate space (see again [95]). Let {zi }~x~=1 be scalars so that SUPn 11y~in__l ziei 11~< 1. We have to show that ~-']~ic~__lziei converges. There is no loss of generality in assuming that zi ~ Ri for every i. By condition (iii) we have for every k that the sequence of ordinals {ak(y'~in=l ziei)}n~k is nonincreasing. Hence there is an no(k) so that for n >~no(k) all the ordinals ~ ( ~ i = l ziei) are the same. This means that all the points ~ i =n 1 ziei belong to the same slice whose diameter is ~< 2 -k. In other words for m ~ n >~n(k), II ~-'~im=nziei II ~< 2 -~ and this proves the claim. Note that by our construction IIen II--+ 1, n --+ oe, where Pn are the basis projections Pn (~--~ai ei) -- Y~in_~_lai ei. Hence lee ]~----1is actually isometric to a conjugate space. [] The proof we just presented of Theorem 4.19 works with small modifications in more general settings. We mention briefly two such situations which are of interest. The first setting is that of spaces with the point of continuity property (PCP). A Banach space X is said to have the PCP if every w-closed and bounded set D in X has a point at which the identity mapping (D, weak) --+ (D, norm) is continuous (point of continuity). Any denting point of D and in particular every strongly exposed point of D (both notions clearly make sense also for non-convex D) are points of continuity. Since any strongly exposed point of cl co D belongs to D (if D is closed and in particular if D is w-closed) it follows that RNP =, PCP. The converse implication holds for large classes of Banach spaces (e.g., for Banach lattices) but there are examples of spaces which have the PCP and fail to have the RNP. The predual of the so-called James tree space has this property (see [19] and [49]). We mention also that in [156] it is proved that any Banach space which has the KMP and PCP necessarily has the RNP.
Infinite dimensional convexity
637
If X has the PCR then for every w-closed and bounded set D in X and every e > 0 * we can find {Xj}l<~j<~m in X* so that D A {x 9 minl~<j~<m Xj(X) > 0 } is non-empty and of diameter < E. Hence, as in the first step of the proof of Theorem 4.19, we can find for any given e > 0, a chain of w-closed sets D~ (or ~< c~(e) some countable ordinal, assuming X is separable) for which Do -- Bx, Dc~+l is strictly contained in Dc~ (if ot < or(e)) and for which the diameter of D~ \ D~+I is at most e. By repeating the argument above with 4~,k finite infima of affine functions (rather than affine functions) we deduce that every infinitedimensional space with the PCP has a separable conjugate subspace. This result is actually also contained in the original paper [74]. In that paper the proof is based on the notion of a G,-embedding which we are not going to define here. A closely related result is contained in [61]. The second setting we want to discuss briefly is that of complex convexity. It has been known for a long time that the theory of convexity has a natural nonlinear analog (for complex Banach spaces) in which linear functionals are replaced by harmonic functions and convex functions by plurisubharmonicfunctions. A real function f on X is said to be plurisubharmonic, written f E PSH(X), if it satisfies the Lipschitz condition and
f ( x ) <~
f ( x +ei~
x, y ~ X.
(In some places in the literature the assumption that f is Lipschitz is weakened to the assumption that f is upper semicontinuous). Clearly the real part of a continuous linear functional belongs to PSH(X). Also f ( x ) = IIx - x011 ( f o r some x0 ~ X) belongs to PSH(X). The class PSH(X) is closed under finite maxima and under finite linear combinations with positive coefficients. The plurisubharmonic convex hull of a bounded set A C X is defined by
fi -- {x ~ X: f ( x ) <<.0 for every f e PSH(X) which vanishes on A}. The set A is said to be PSH-convex if A -- A. Any closed bounded and convex set is PSHconvex. A probability measure # on X is said to be a Jensen measure representing a point x0 if for every f 6 PSH(X)
f (xo) <~Ix f (x) d#(x). It follows from this definition that if f 6 X* then f(xo) - fx f ( x ) d # ( x ) , i.e., x0 is the barycenter of the measure #. The normalized Haar measure on the circle {x0 + e i0 y: 0 ~< 0 ~< 2yr } is an obvious example of a Jensen measure representing x0. It is easy to check that if A is closed and bounded, then any barycenter of a Jensen measure supported on A belongs to ,4. If A is compact, then the converse is also true. Here are the analogues of the notions of extreme and exposed points and also other notions defined in this section in the setting of regular convexity. In what follows A is a bounded set in X and x0 ~ A.
638
V.P Fonf et al.
(1) The point x0 is a complex extreme point of A if there is no 0 # y 6 X for which {x0+rei~ 0~
Infinite dimensional convexity
639
We consider now a special class of martingales {Mn } called analytic martingales. These martingales are defined on S2 = T r~ where T is the unit circle and N the non-negative integers. For 0 = (01,02 . . . . ) E S2 the functions Mn (0) have the form
Mn (0) = ho + ~
hn (01,02 . . . . . On- 1) e i~ ,
n--0,1,2 .....
k--1
where hn " T n-I --+ X are bounded measurable functions. In analogy to (1) r (3) in Theorem 4.6 we have THEOREM 4.22 ([48]). A Banach space X has the ARNP if and only if any analytic martingale {Mn} with SUpn IIM~llg~(~,x~ < ec converges a.e. on X?. By applying Theorem 4.22 one obtains the following analogue of Theorems 4.15 and 4.16.
Let X be a complex Banach space. The following assertions are equivalent: (1) X has the ARNP. (2) Every closed and bounded subset of X is contained in the plurisubharmonic convex hull of its strong barrier points. (3) Every non-empty bounded subset of X has non-empty plurisubharmonic slices with arbitrary small diameter. THEOREM 4.23 ([73]).
By using Theorem 4.23 and the proof we presented for Theorem 4.19 (as well as the special case of Dvoretzky's theorem on existence of 2-dimensional almost spherical sections) one can obtain the following generalization of Theorem 4.19. THEOREM 4.24 ([75]). Any infinite-dimensional space with the ARNP has an infinitedimensional subspace which is isometric to a separable conjugate space.
5. Boundaries and support points DEFINITION 5.1. Let X be a Banach space. A subset B C Sx, is called a boundary for X if for each x E X there is an f E B such that f (x) - IIx II. This definition of a boundary is a generalization of a notion of a boundary for subspaces of the space C ( K ) (or the space Cc(K) of all complex-valued continuous functions on a compact K). A subset B C K is called a boundary for a subspace X C C ( K ) if for each x E X there is a t E B such that Ix(t)l - Ilxll. We also define below another kind of boundary for subspaces of C (K), the so-called Choquet boundary. As usual, for a subspace X C C ( K ) (or Cc(K)) we identify in a natural way the points t 6 K with the appropriate functionals t E X*" t(x) -- x(t), x E X.
640
V.P Fonf et al.
DEFINITION 5.2. Let X be a subspace of C ( K ) (or Cc(K)) which separates the points of K and contains 1. The set S(X) = { f 9 X*: f ( 1 ) -- 1 = IIfll} is called the state space of X. The Choquet boundary B ( X ) is the set of all points in K which are extreme points of the state space S(X). There is an interesting connection between this notion and approximation theory. A subset A C C ( K ) is called a Korovkin set if for each sequence {Tn } of positive operators in C ( K ) such that limTnx = x for each x 9 A, we have limTny = y for each y 9 C ( K ) . For instance, Korovkin's Theorem asserts that the set A -- { 1, t, t 2 } is a Korovkin set for C[0, 1]. More generally, we have THEOREM 5.3 ([158]). Suppose that K is compact metric and that X is a subspace of C ( K ) which separates the points of K and contains 1. Then X is a Korovkin set if and only if B ( X ) = K. The simplest examples of boundaries for a Banach space X are either the whole unit sphere Sx, of the dual space (use the Hahn-Banach Theorem) or the set ext B x , of extreme points of the unit ball of the dual space (use the Krein-Milman Theorem). In general, a boundary B need not consist of extreme points and it is even possible to have B Mext B x , -- 0. (Take X -- Ii ( F ) , F an uncountable set, and B to be the set of all the vectors in l ~ ( F ) with just countably many non-zero coordinates, all of which are +1.) However, all boundaries for X must contain the w*-exposed points of Bx, which, for X separable, form a large part of the set of extreme points. It is clear that an x 9 Sx w*-exposes some f 9 B x , iff x is a smooth point of Sx. In other words, s m S x = {x 9 Sx" 3 f 9 w*-exp Bx, with f (x) - 1}.
(5.1)
If X is separable, then by Mazur's theorem the set sm Sx of all smooth points of Sx is dense in Sx (see [95]). Thus, we have w*-cl(w*-exp B x , ) = w*-cl ext B x , in this case. The notion of a w*-exposed point may also be defined for complex Banach spaces. The functional f is a w*-exposed point of B x , provided I f ( x ) l - - 1 for some x e Sx and g = )~f for some I&l = 1 whenever g 9 B x , and Ig(x)l = 1. In the setting of Definition 5.2 those points k 9 K in the Choquet boundary which correspond to w*-exposed points are called, for obvious reasons, peak points. THEOREM 5.4 ([11]). Let K be a compact metric space and A C Cc(K) be a Banach subalgebra which contains the constants. Then ext BA, = w*-exp BA,. A proof of this theorem can be found in [136, Section 8]. Theorem 5.4 shows that for a Banach algebra A which satisfies the conditions of Theorem 5.4, there is a unique (up to absolute value 1 multipliers) boundary, which is minimal (i.e., it is contained in all other boundaries, up to absolute value 1 multipliers). As simple examples show, in general this is not true. One of the basic results on support points is the Bishop-Phelps theorem.
Infinite dimensional convexity
641
THEOREM 5.5 ([ 13]). Let V be a convex closed bounded subset of a real Banach space X. Then the set of all support points of V is dense in OV and the set 27(V) of all support functionals of V is dense in X*. The proof of the second statement is given in the basic concepts article [95]. Completeness of the space X in the Bishop-Phelps theorem is essential, as the following proposition shows. PROPOSITION 5.6 ([65]). Each incomplete separable normed space contains a closed convex bounded subset V such that Z, (V) = 0. Moreover no non-constant affine functional which is just assumed to be continuous on V attains its supremum on V. The assumption that the Banach space X in Theorem 5.5 is real is also essential. In [ 124] a complex Banach space X and a CCB set V C X are constructed such that Z (V) -- 0. We briefly indicate the construction. Consider the space H ~ of bounded analytic functions on the open unit disc D which vanish at the origin, with the supremum norm. Then H ~ = X* for some X (actually X = L1/H1). For each z E D the functional q9z on H ~ defined by qg(f) = f ( z ) belongs to X. Let V be the closed convex hull of {qgz: z 6 D}. Then for every f E H ~ , ]lf II = sup{If (x)[: x E V }. However, there is no f ~- 0 in H ~ for which 11f ]1 = If (x)[ for some x E V. We show next how the Bishop-Phelps theorem may be used in the study of boundaries. THEOREM 5.7 ([147]). Let X be a real Banach space, K C X* be convex and w*compact and B C K be a norm separable subset such that for each x E X there exists f E B with f ( x ) --sup{g(x): g ~ K}. Then K is the closure in the norm topology ofthe convex hull of B. Rod6 obtained this result from his minimax theorem of "superconvex analysis" (cf. [ 146]); the proof we present here seems to be new. PROOF. Fix an 0 < e < 1 and a s e q u e n c e {gi}~x~=0, 0 < 6i < 6, limi 6i - - 0 . Let {hi}~X~=l be a norm dense subset of B. Put Co = e B x , and for each integer i = 1, 2 . . . . put Ci = Bf-q(hi + e B x , ) . L e t K1 = c o { K U C o } , B 1 =LJi=o(1-+-ei)Ci, V* = w* -clcoB1. Clearly B C co B1 and by the separation theorem K1 C V*. Also it is clear that V* is a w*-closed body in X* with 0 6 int V*. Let V be the polar of V* in X. We shall show that oo
w*-clB1N Z ( V ) NOV* C U w*-cl(1 + 6i)Ci. i=0
Indeed, if (5.2) is not valid then there exist an
f E
w*-clB1 \ U w * - c l ( 1 -3r-6i)Ci NOV* i=0
and an x E V such that f ( x ) = 1.
(5.2)
V.P Fonf et al.
642
Since limi 8i - - 0, f E w*-cl Ui=0 Ci. Indeed, by the definition of f , there is a net {{P•215 which converges to f so that qg• = (1 + ei(•215 with lim• i ( y ) = ~ and ~• E Ci(• Hence f = lim• ~p• Thus f E K1 and by using Kl C V* we deduce that 1 ~ sup{g(x)" g E K1 } ~ sup{g(x)" g E V*} -- 1. Now, B -- Ui~1761 ci implies that for some index i there exists h E Ci with h(x) -- 1. Since (1 -k- ei)h E V* we see that sup{g(x): g 6 V*} ~> 1 + ei, which contradicts the fact that x E V and (5.2) is proved. Since V* - w*-cl co B1, it is the w*-closed convex hull of the compact set w*-cl B1, so by Proposition 2.3, any element of V* is represented by a probability measure supported by w*-cl B1. It follows that if, in addition, f E r ( V ) A 0 V*, then # is supported on (w*-cl B1) 7 / r ( V ) (10V*. (In fact, if x E V and sup{g(x): g ~ V*} = f ( x ) , then # is supported by {g E V*: g(x) = 1} M w*-cl B1 .) From (5.2) we get oo
supp/z C U w*-cl(1 + 8i)Ci. i=0
n-1 Put Do -- (1 + e0)C0, Dn = w*-cl(1 + en)Cn \ U i = 0 w*-cl(1 + 8i)Ci, n = 1, 2 . . . . . Let {7 = {i" ~i = #Di > 0} and let ti be the barycenter of the measure (1//zi)/z restricted to Di, i E {r. It is clear that ti
E
(1 + ei)hi + e(1 + ei)Bx* C hi + (ei Ilhi II-4- e(1
-t-
8i))Bx,
C hi d- e ( L + 2)Bx,, where L = sup{ IIf I1: f E K}. Put
g- Z
#i hi E II. II-cl co B
iEcr
and observe that I I f - gll ~ e(L + 2). Consequently, by the Bishop-Phelps theorem we get that for each f 6 0 V* there exists a g 6 II.ll-cl co B such that IIf - gll ~< e(L + 3). The same holds therefore for every f 6 V* and in particular for f 6 K. Letting e ---> 0, we see that II. II-cl co B = K, which concludes the proof. [2 For the particular case K = Bx, the following generalization of Theorem 5.7 holds. PROPOSITION 5.8 ([35]). Let the real Banach space X admit a boundary B C Bx, which may be covered by a sequence of balls with radius a, 0 < a < 1. Then the dual space X* is separable and the unit ball Bx, is the norm closed convex hull of B. PROOF. Let B C Ui~176 Jr-aBx,). The proof runs along the lines of the proof of Theorem 5.7 (with obvious simplifications), but instead of the inequality Ilf - gII ~< e(L + 3) we get I I f - g l l ~< (1 +e)a < I for sufficiently small e > 0. By the Riesz lemma, X* -- [hi]iC~=l, i.e., X* is separable and we are reduced to the setting of Theorem 5.7. [2
Infinite dimensional convexity
643
We now present James' theorem, the separable case of which is an immediate corollary to Theorem 5.7. THEOREM 5.9 ([93]). Let B be a closed convex and bounded subset o f a real Banach space X. Then the following are equivalent: (1) B is weakly compact. (2) Each functional f E X* attains its supremum on B. In particular, a Banach space X is reflexive iff each functional f E X* attains its supremum on the unit ball o f X. PROOF. (1) =~ (2) is obvious. We prove (2) ~ (1) just for separable B (for the nonseparable case see [93]). Let K be the w*-closure of B in X**. By Theorem 5.7, K -[[. II-cl co B = B, i.e., B is weakly compact. [] James' theorem is also valid (in a natural modified form) for complex Banach spaces. Unlike the situation with the Bishop-Phelps theorem, the passage from the real result to the complex one is completely routine. THEOREM 5.10. Let B be a closed convex and bounded subset o f a complex Banach space X. Then the following are equivalent: (1) B is weakly compact. (2) For every f E X* there is an x E B such that If(x)[ -- sup{[f(y)l: y E B}.
PROOF. Assume that (2) holds. Let B = I[.ll-clco{ei~ 9 0 ~< 0 ~< 27r, x E B}. Clearly, for every f E X*, sup{If(y)l" y 6 B} = sup{lf(y)]" y E B}. Consider X as a Banach space over the reals and denote it by Xr. There is one-to-one correspondence between Xr* and X*" To h E Xr* corresponds f (x) = h (x) - ih (ix) E X*. Each h E Xr* attains its supremum on B. Indeed, let f be the corresponding element in X* and let x E X satisfy [f(x)l = sup{lf(y)l" y E B}. Let 0 ~< 0 ~< 27r be such that I f ( x ) l - f ( e i ~ = h(ei~ Then h(ei~ -- sup{Ih(y)l" y E B}. By Theorem 5.9, B is weakly compact and so is its weakly closed subset B. D The following theorem is a dual version of James' theorem, and is also an immediate corollary of Theorem 5.7. THEOREM 5.1 1. Let B be a separable closed convex and bounded subset o f a dual Banach space X*. Then the following are equivalent: (1) B is w*-compact. (2) Each x E X attains its supremum on B. Let (R) be the class of all reflexive Banach spaces, (RN) be the class of all Banach spaces with the Radon-Nikod~m property, and (NCco) (non-containing co) be the class of all Banach spaces which do not contain an isomorphic copy of co. Clearly, (R) C (RN) C (NCco). By the James theorem, X 6 (R) iff 27(Bx) = X*. By Theorem 4.16 and Proposition 4.17, X ~ ( R N ) iffthe set S , ( B x ) is ofthe second category in X* for each
644
V.P. Fonf et al.
equivalent norm on X (assuming X is separable). The following proposition complements this chain. PROPOSITION 5.12 ([62]). Let X be a real Banach space which does not contain an isomorphic copy of co. Whenever r ( B x ) = Ui~ ci is represented as the union of an increasing sequence o f sets, there exists an index j such that the set w*-cl co -+- (Cj A Sx) contains some ball. Conversely, if X is separable and contains an isomorphic copy of co, then there exists an equivalent norm f o r which the set Z ( B x ) does not have the above property. A common approach in recent years to proving James' theorem and related results is based on the following proposition (which was obtained by streamlining James' original proof). PROPOSITION 5.13 ([ 159]). Let B be a set and let C be a bounded subset o f l ~ (B) which contains ~-'~i=1 )~ixi whenever {xi}~=l C C, )~i > 0 and Y~i~l )~i = 1. Assume also that f o r every x ~ C there is a t ~ B such that x(t) - p ( x ) ---- sup{x(s)" s ~ B}. Let {Xn}n~=l C C and l e t u ( t ) = limsuPnXn(t) f o r t ~ B. Then p(u) >~inf{p(x): x 6 C}. A proof of this proposition is presented, e.g., in [41, p. 15]. We will demonstrate how Proposition 5.13 works by using it to prove the following result. THEOREM 5.14 ([78,83,150]). Let X be a separable real Banach space. The following assertions are equivalent: (1) X does not contain an isomorphic copy of l l. (2) Each convex closed bounded subset K o f X*, such that each x ~ X attains its supremum on K, is w*-compact. (3) For each boundary B f o r X (in any equivalentnorm), B x , is the norm closed convex hull of B. (4) For each boundary B f o r X (in any equivalent norm), X* is the norm closed linear span of B. Recall first some notions from harmonic analysis which will be used in the proof of Theorem 5.14. Let E be a subset of the set of integers Z. We denote by C E ( T ) the space of continuous functions on the circle group T for which f ( n ) = 0 if n ~ E. The set E is called a Sidon set if there is C > 0 such that II ~nm=lan eintllf~T) >~ C Y~nm=l lanl for each trigonometric polynomial P (t) - ~nm__x an e int whose Fourier transform is supported on E. E is a Sidon set iff C E ( T ) is isomorphic to 11(E), iff for each sequence u(n) ~ l ~ ( E ) there is a Radon measure # on T with u (n) - / 2 (n) for every n 6 E. E is called an interpolation set if for each u(n) ~ l ~ ( E ) there is a purely atomic measure v on T such that u(n) = ~(n) for every n ~ E. Clearly, each interpolation set is a Sidon set; the converse is not true [97]. Let L C l ~ (Z) be the space of all functions of the form m
p(n) -- Z
ak e int~ ,
tl, 6 T, n E Z.
(5.3)
k=l
The closure A P ( Z ) of L in l ~ is called the space (it's actually an algebra) of almost periodic functions on Z. We may consider Z as a subset of the dual space AP* (Z). The
Infinite dimensional convexity
645
w*-closure of Z in AP*(Z) is called the Bohr compactification b Z of Z (recall that the w*-closure flZ of Z in l * is the Stone-Czech compactification of Z). By the StoneWeierstrass theorem, A P ( Z ) = C (b Z). THE PROOF OF THEOREM 5.14 [78]. (4) :=> (1). Suppose to the contrary that X D I1. We shall construct an equivalent norm on X such that (4) does not hold even for B = w * - c l e x t B x , . We consider just the particular case X = ll. (For the general case X D ll see [78].) We construct the desired norm on the complex space ll and then, by a standard procedure (see the proof of Theorem 5.10), we deduce the result for the real space 11. Let E be a Sidon set which is not an interpolation set and let Y = C E ( T ) . (In particular, x --+ } is an isomorphism of Y onto the complex space l l (E).) We will show that the norm on l l which comes from Y via this isomorphism works. By a result of Kahane [97], the set E is an interpolation set iff each bounded function u(n) E lot(E) may be extended from E to Z as an almost periodic function on Z. Let ~ :AP(Z) -+ loc = C (I3 Z) be the natural embedding. With the help of the above mentioned result of Kahane and the TietzeUrysohn lemma it is not difficult to see that E is an interpolation set iff the restriction gz*l~E is a homeomorphism of fiE onto bE. (The set fiE is just the w*-closure of E in C*(fiZ) and b E is the w*-closure of E in AP*(Z).) Clearly the restriction ~P*I~E is a homeomorphism iff for each A C E there is a function p of the form (5.3) such that inf{[p(n) - p(k) 1: n ~ A, k ~ E \ A} > 0. As is well known, for each g E ext By, there exist a t ~ T and a complex number a, la] = 1 such that g(x) = ax(t) for each x ~ Y. Therefore each g E span{w*-cl ext By, } has a representation g(x) -- ~-,k=l m akx(tk), where tk ~ T. Put g , ( n ) - - ff-~km=l a k e intk . Clearly s is of the form (5.3). Since E is not an interpolation set, there is an A C E such that inf{lp(n) - p(k)]: n ~ A, k ~ E \ A} = 0 for each p of the form (5.3). In particular, the characteristic function of A does not belong to [w*-cl ext B x , ] (we identify here the space C*E(T ) with lee(E)), i.e. the subspace [w*-clextBy,] is a proper subspace of Y*. (1) => (2). Suppose to the contrary that K is not w*-compact. Then there is an f e w*-cl K \ K. Let u c Bx** be such that
u ( f ) > oe > sup{u(g): g e K}
(5.4)
for some ol. Since X 75 11, by the Odell-Rosenthal theorem [134] there exists a sequence {Xn}neC__l C BX such that u -- w*-limxn. Put C -- {x ~ Bx" f (x) ~ ol}, B -- K and apply Proposition 5.13. (We may clearly assume that Xn 6 C for all n.) We have sup{u(g)" g 9 B} >~ inf{sup{g(x)" g e B}" x ~ C} >~ inf{f(x)" x ~ C} >~ ce, which contradicts (5.4). The other implications are trivial. Next we consider the question how properties of a Banach space may influence properties of the boundaries for this space. THEOREM 5.15 ([121]). Let X be an infinite-dimensional reflexive Banach space. Let B C Sx be a boundary f o r X*. Then B is uncountable. In particular ext B x is uncountable.
V.P. Fonf et al.
646
PROOF. Assume to the contrary that B = {Xi}~=1. Put Fi = { f 6 BX*" f ( x i ) - - I l f l l } , oo
i = 1, 2 . . . . . By the reflexivity of X and by the definition of boundary, Bx, -- U i = l Fi. By the Baire category theorem, there is a j such that the set Fj contains a w*-relatively open set in Bx,. Thus, there are a functional f ~ Fj and a finite set {Yk}~_-i C X such that
g ~ Bx*" max
Ig{Y~)- f(yk)l
< 1} c Fj.
Without loss of generality we may assume that Ilfll - 1 - e, ~ > 0. Put L = ( [ y k ] L 1 ) • n [xj] • It is clear that ( f + L) n Bx, C Fj. Since d i m L > 0 there is an h 6 ( f + L) n Sx,. Therefore 1 = h(xj) -- f ( x j ) = IIfll -- 1 - e < 1, a contradiction. O Actually the same proof is valid for any CCB (not necessary symmetric) body in a reflexive Banach space X. Recall that the set of extreme points of a metric convex compact in a topological vector space is always a Ga set, and hence homeomorphic to a Polish space (see Section 2). Therefore, by Theorem 5.13, if X is an infinite-dimensional separable reflexive Banach space, then for each CCB body V C X, the set ext V (in the weak topology) is homeomorphic to an uncountable Polish space. The converse also holds. PROPOSITION 5.16 ([112]). Each uncountable Polish space is homeomorphic to the set of extreme points (in the weak topology) of some CCB body in 12. This proposition complements Haydon's theorem (Theorem 3.6) and its proof uses this result. We mention that in [57] a stronger version of Theorem 5.15 is proved which actually gives a characterization of reflexivity. In [121] an equivalent norm on Hilbert space was constructed for which the set strexp Bx is countable. The intermediate set exp Bx is the subject of Theorem 5.19 and Corollary 5.23 below. For our next result we need a new notion. Recall that a subset C C X* of a dual space is called r-norming (0 < r <~ 1) for X if inf{sup{If(x)l: f ~ C}: x ~ Sx} -- r. A subset C C X* is called norming if it is r-norming for some r > 0. A subspace E C X* is called norming if its unit ball BE is a norming subset. Each norming subspace is total (i.e., f (x) = 0 for all f 6 E implies x = 0) but the converse is not true. Indeed, the following proposition holds. PROPOSITION 5.17 ([38]). A Banach space X is quasireflexive (i.e., dim X** / X < c~) iff
every total subspace of X* is norming. DEFINITION 5.18 ([62]). A subset C C X* is called thin if it can be covered by an increasing union of non-norming sets, i.e. oo
ccUci, i--1
where inf{sup{[f(x) l" f E Ci }" x E Sx } -- O, i = 1, 2 . . . . .
647
Infinite dimensional convexity It is clear that if X is infinite-dimensional, then each countable set in X* is thin.
THEOREM 5.19 ([66]). Let X be a separable Banach space. The following assertions are equivalent: (1) The space X contains an isomorphic copy o f co. (2) There exists an equivalent norm on X such that the set w*-exp B x , is thin. To prove Theorem 5.19 we need some auxiliary results. We start by a useful characterization of thin sets. LEMMA 5.20 ([68]). Let X be a Banach space and C be a bounded subset o f X*. The following assertions are equivalent: (1) The set C is thin. (2) There exists a linear bounded one-to-one operator T : X --+ E into some Banach space E such that the inverse operator T -1 is u n b o u n d e d a n d T * ( E * ) D C. Let T : X --+ E be a linear bounded one-to-one operator from a Banach space X into a Banach space E such that the inverse operator T -1 is unbounded. We call the topology on X generated by the sets of the form T - I ( G ) , where G ranges over the open subsets of E, the E-topology of X. LEMMA 5.21 ([66]). Let T : X --+ E be a linear bounded one-to-one operator from a separable Banach space X into a Banach space E and let C C Sx, M T* E*. Set
Z - - {x E Sx" 3 f E C f ( x ) - -
l}.
Then Z can be represented as a countable union o f subsets o f Sx which are E-closed in B x . PROOF. Without loss of generality we can assume that T has a dense range and therefore E is separable (as well as X) and T* is a one-to-one operator. Write
Vn--{xESx"
3f ESx, MT*(nBE,)with f(x)--l},
n--1,2 .....
It is clear that Z C U Vn. We shall show that each Vn is E-closed in Bx. Let E- l i m x j = x, xj E Vn, x ~ B x . By the definition of Vn there exists a sequence of functionals { f j } C Sx, M T * ( n B E , ) such that f j ( x j ) = 1, for all j. Let gj = T * - l ( f j ) , j = 1,2 . . . . . Since the sequence {gj} C nBE, is bounded and E is separable, we can assume that there exists g E n BE,
such that w*- lim gj
=
g.
(5.5)
Since l i m T x j = T x it follows from (5.5) that g ( T x ) = l i m g j ( T x j ) = lim f j ( x j ) and therefore g ( T x ) = 1. It is clear that f -- w * - l i m T * gj = T* g E B x , M T * ( n B E , ) . Since f ( x ) = 1 and x E B x we deduce that I l x l l - I l f l l - - 1 and thus x E Vn. D The following proposition is the main step in the proof of Theorem 5.19.
648
V.P. Fonf et al.
PROPOSITION 5.22 ([66]). Let X be a separable Banach space and r be a Hausdorff topology on X that is consistent with the linear structure o f X and is strictly weaker than the norm topology. I f sm Sx C ~_J~ Vn, where each Vn is a r-closed in B x subset o f S x , then X contains an isomorphic copy o f co. PROOF. Without loss of generality we can assume that the sequence {Vn }n~__lis increasing. Let T:12 --+ X be an arbitrary compact linear operator with a dense range such that the image of the unit ball K1 = TBI2 is contained in int B x . Since K1 N V1 = 0 and K1 is compact, there exists a r-closed balanced r-neighborhood of zero G1 such that (K1 + G1) N V1 = 0. Since the set G1 is unbounded, there exists an element Xl e r a n t G1 such that, for K2 ---]-Xl § K1, (1) K2 C intBx, (2) K2 C K1 § r a n t G1, (3) dll.ll(K2, S x ) < 2 -2. By K2 A V2 = 0 (see (1)), compactness of K2 and (2), there exists a r-closed balanced rneighborhood of zero, G2, such that (K2 + G 2 ) C (K1 + r a n t G1) and (K2 + G2) f'l V2 = 0. Since the set G2 is unbounded there exists an element x2 e r a n t G2 such that for K3 = -f-x2 § K2, K3 C int B x , K3 C K2 + r a n t G2, dll.ii (K3, S x ) < 2 -3. In this way we construct a sequence of elements {Xn }, a sequence of compact sets Kn y.~,~-I +xi + K1, n -- 1, 2 . . . . . and a sequence of r-closed sets {Gn} such that (Kl + G1) (K2 § G2) D (K3 § G3) D ..- and (Km § Gm) f-) Vm ----13 for every integer m. Thus, (a) (y~7 +n -']-Xi § gl) C (y~n~ -+-Xi + g l § Gm+l), m , n - 1, 2 . . . . . (b) (y~n -+-xi § KI + Gm+l) f') Vm+l -- 0, m = 1, 2 . . . . . (c) dll.ii( y ~ n -+-xi + K1, S x ) < 2 -m, m = 1, 2 . . . . . Let us assume that the Banach space X does not contain co. Then, by the BessagaPetczyfiski theorem [9], the series Y~xi converges unconditionally and the set F {~_,~ aixi" ai -- -4-1} is compact. By the construction, F + Kl C B x . From (a) and (b) (sending n to infinity) we get (F + K1) A Vm+l -- 0 for all m. Therefore (F + K1) A sm Sx - 0. From (c), using the compactness of the set F + K l, we deduce that (F + K 1 ) N S x ~ 0. Letx e F, y e K1 be such that (x + y ) e Sx. S i n c e ( F + K 1 ) C B x we have y e OaK1 (OAK1 is the algebraic boundary of the set K1). Since x + y is not a smooth point of Sx there exist two distinct functionals f, g e S x , such that 1 - f (x + y) - g(x + y). It is clear that the functionals f and g are supporting functionals for the set x + K1 (recall that (F + K1) C B x ) . But every support point of the set x + Kl is a smooth point of Sx, since K1 is an image of the unit ball of Hilbert space and linK1 is dense in the space E. This contradiction completes the proof. [2 THE PROOF OF THEOREM 5.19. (2) => (1). The proof of this implication is just a direct combination of Lemmas 5.20 and 5.21, equality (5.1) and Proposition 5.22. The inverse
Infinite dimensional convexity
649
implication (1) =~ (2) is a simple consequence of Sobczyk's theorem: co is complemented in each separable Banach space which contains it (for details see [66]). D REMARK. The proof of Theorem 5.19 works just as well if we replace in (2) the unit ball of an equivalent norm by a general CCB body (i.e., not necessary symmetric). By applying Theorem 5.19 to reflexive spaces we get, in particular COROLLARY 5.23 ([66]). Let W C X be a CCB body in an infinite-dimensional reflexive Banach space. Then the set exp W is uncountable. It is perhaps worthwhile to state the following reformulation of Theorem 5.19. THEOREM 5.24. Let X be a separable Banach space which does not contain an isomorphic copy of co and let the set w*-exp B x , be represented as a union of an increasing sequence of sets: oo
w*-exp B x , -- U Di. i=1
Then there exists an integer j such that the
set Dj
is norming.
Theorem 5.24 looks like the Baire Category theorem. This similarity is strengthened if we recall that a set D is norming iff the set w*-cl co{+D} contains some ball. The following consequence of Theorem 5.24 may be considered as a more precise version of the Banach-Steinhaus theorem, for separable Banach spaces that do not contain co. COROLLARY 5.25 ([66]). Let X be a separable Banach space which does not contain an isomorphic copy of co and let A be an unbounded subset of X. Then there exists an f E w*-exp B x , such that the set f ( A ) is unbounded. PROOF. Suppose to the contrary, that for each f E w*-exp B x , , the set f ( A ) is bounded. By writing Dn = {f E w*-expBx,: sup If(A)l < n}, we have w*-exp B x , = U Dn, where {Dn } is an increasing sequence of sets. By Theorem 5.24, there exists an integer m such that the set Dm is norming, i.e.,
3 c >OVx E x ,
Ilxll
C up{lf(x>l f EDm}.
Thus sup{ Ilx II: x E A} ~< Cm, a contradiction. We conclude this section with an application to classical analysis. The following result may be considered as a generalization of a well-known theorem of Zygmund on lacunary trigonometric series. We call a sequence {Xn(t)}n~__l of functions in C[0, 1], a generalized Sidon system if it is equivalent to the natural basis of l l.
V.P Fonf et al.
650
COROLLARY 5.26 ([62,66]). Let {Xn}n~__l C C[0, 1] be a generalized Sidon system and X = [Xn]n~=l. Assume that X separates the points of [0, 1]. Let B C [0, 1] be the set of all peakpoints for the subspace X. If the partial sums of the series ~-,n=l anXn(t) are bounded oo at each point t E B, then the series converges absolutely, i.e., Y~n=l [anl < oo. PROOF. As usual, we identify the interval [0, 1] with a subset of C*[0, 1] and moreover, with a subset of X*. By definition, w*-exp Bx, = +B. Put k
Dn -- { s E B" supk ~ a i x i ( s )
n=l,
....
i=1
It is clear that the sets Dn form an increasing sequence of sets and that w*-exp Bx, = Un%l i D n . Since X, being isomorphic to ll, does not contain co, it follows from Theorem 5.24 that there exists an index m such that the set Dm is norming (say r-norming). In particular, for each n
rll
aixi
i--1
{ I k aixi (S) " s E D m }
~< sup s~p Z
~m.
i=1
Since {Xn } is equivalent to the natural basis of 11, we deduce that ~n~176lan I < cx).
[-1
6. Convex polytopes In this section we study the class of infinite-dimensional polytopes. We shall consider only closed convex bounded (CCB) sets in a Banach space. DEFINITION 6.1. A closed convex bounded set P in a Banach space X is called a polytope if every finite-dimensional section of it is a (usual finite-dimensional) polytope. The Banach space X is called polyhedral if its unit ball Bx is a polytope. This definition was introduced by Klee [105] (who actually considered only unit balls of Banach spaces). There are in the literature other definitions of polytopes. Some of them are isomorphically equivalent to Definition 6.1 (see [45] and [71]), while the others are completely different, see, e.g., Definition 6.27. It seems however, that only the notion we have just defined has led to an interesting theory. In [103] Klee proved that a finite-dimensional CCB set is a polytope iff every twodimensional section of it is a polygon. Thus in Definition 6.1 it is enough to consider only two-dimensional sections of P (and if P is symmetric, just two-dimensional central sections). This observation will simplify several proofs below. It is reasonable to expect that a discussion of polytopes will start with some examples of such sets in "nice" Banach spaces (for example Hilbert space). It will soon become evident that this is impossible; there are no infinite-dimensional polytopes in reflexive Banach spaces. On the other hand, the unit ball of co is easily seen to be a polytope. In some sense
651
Infinite dimensional convexity
it is a typical example of an infinite-dimensional polytope. All this will be clarified by the results below. We start with a result which shows, in particular, that sections cannot be replaced in Definiton 6.1 by continuous images. THEOREM 6.2 ([117]). Every infinite-dimensional Banach space X has a two-dimensional quotient space whose unit ball is not a polygon. The idea of the proof is simple. We start with any two-dimensional quotient which is a polygon and show that we can perturb it slightly to get a two-dimensional quotient with more vertices. In this way we get a sequence of two-dimensional quotient spaces whose unit balls have a monotonely increasing number of vertices which converge to a limiting two-dimensional quotient space whose unit ball is not a polygon. This approximation argument has to be done somewhat carefully since the number of vertices of a polygon is only a lower semi-continuous function (in the Hausdorff metric). In order to present the details of the argument we need three simple lemmas. The first two are simple facts about finite-dimensional polytopes. We shall use below the following notation. For a set C we denote its cardinality by ICI. For a set A in a Banach space and e > 0 we denote by A~ the set {x: d ( x , A) ~< e}. The first lemma is completely elementary. LEMMA 6.3. Let P be a symmetric polygon in the plane (taken, f o r instance, with the usual Euclidean norm). Then there is an e = ~(P) > 0 so that f o r any convex body C in theplane with OC C (OP)~ we have lextC] ~> ]extP]. LEMMA 6.4. Let X be a finite-dimensional polyhedral Banach space. Let T be a linear map from X t o I[~ 2 o f rank 2. Assume that there is a hyperplane Y C X so that T B x = T By. Then f o r every e > 0 there is a linear W : X - - + ~2 SO that IITIy - WlYll ~ s,
OWBx C (OTBx)e,
lext W B x l > l e x t T B x l .
171 PROOF. Let {Vj }j__l be the set ext T B x . If, for some j , T -1Vj 0 B x is not a singleton let ~o be a functional on X which takes on this set both positive and negative values and let u be a vector in the direction of a line in the plane which touches T Bx just at vj. The operator W x = T x + O~p(x)u will have all the desired properties for small enough positive 0. We may thus assume that for every j there is a unique xj ~ ext B x such that T x j = vj. Clearly Y D {Xj}jm=l 9Let T1 be an operator with liT1 - TII ~< e / 2 so that T1 is one-to-one on ext B x . If lext T1Bxl > lext T Bxl we can take W = T1. If this is not the case (and e is small enough) it follows that ext T1Bx - - {Tlxj}jm=l 9Let u be a vector in the plane so that no vector of the form {/'1 (p - q): p, q 6 extBx} is in the direction u, let q9 be a non-zero functional which vanishes on Y and consider T2,z(x) = Tlx + )@(x)u. Clearly for all )~, T2,~. is one-to-one on ext B x . Let )~0 = max{)~ ~> 0: T2,~B x C T1Bx }. For sufficiently small positive 0 the operator W -- T2,)~0+0 will have the desired properties. D
LEMMA 6.5. Let X be an infinite dimensional Banach space and let T be a quotient map from X onto a two-dimensional polyhedral space L. Let 2n = [ext BL] and let m ~ n. Then
652
V.P. Fonf et al.
there is an (m + 1)-dimensionalquotient space Z o f X and quotientmaps Tt: X --+ Z and V" Z --+ L so that T = V T I.
PROOF. Let ext BE = { - q - u i } ~ . Put cr -- {i" T - l u i 0 S X - - ~ } . If cr -- 0 let Y' be any subspace of Ker T of codimension m + 1 in X and put Z -- X~ Y'. If o- 7~ 0, we can choose for every i 6 cra sequence {xJ}~_l- and vectors xi so that T(xi + x j ) -- Txi -- ui for all j and limj ]]xi + x j 11- 1. Since the sequence {xJ}j_l has no w-cluster points we can assume (by passing to a subsequence if needed), that {x j }~c_1 is a basic sequence. By passing to further subsequences we may assume that Y1 -- [x/]iEcr, j--1 .... is of infinite co-dimension. Take as Y' any subspace with Y1 C Y' C ker T so that the co-dimension of Y' in X is (m + 1) and put Z - X~ Y'. [3 PROOF OF THEOREM 6.2. Assume to the contrary that all two-dimensional quotient
spaces of X are polyhedral. By [105] all finite-dimensional quotient spaces of X are polyhedral too. We start with an arbitrary two-dimensional quotient space L 1 of X with the quotient map 7'1 : X --+ L1. Put 2nl = ]ext BLI] and apply L e m m a 6.5 for L -- L1, T = 7'1, n = n,, m = n l + 1. By the choice of m there is a hyperplane (say Y1) in Z1 such that V1Br~ contains ext BL~ and thus V1Br~ = BE1. Put 2el = e(BCl) (the number given by L e m m a 6.3). Next, choose points {xl}l ' in Sx and 61 > 0 such that T(xli 6 Y, for e v e r y / and such that, whenever Ilui - Tlx 1 II < 61 for every i, we have 0 co{u]}l 1 c (Bcl)~I. By L e m m a 6.4 there exists an operator W1 : Z1 --+ L1 such that 2 n 2 - - l e x t W l B g l l > 2n,, and
[IWIT;x: - T, xlll < 6 , / 2
Vi
OWIBz1 C (SL1)ej.
Let T2 - - W 1 T( and L2 be the two-dimensional Banach space with the unit ball BL2 = W1BzI. Apply L e m m a 6.5 again for T = T2, L -- L z, n -- n z, m -- n l + n2 + 1. Let Y2 be a hyperplane in Z2 with T2x ] E Y2 for every i -- 1 . . . . . n l and VzBr2 = B L 2 . Take e2 so small that (BLz)E2 C (Bcl)~I and e2 < e(BL2)/2. Choose 32 > 0 and {X2}l 2 C Sx as in the first step and then by L e m m a 6.4, find an operator W z ' Z 2 --+ L2 such that IIWzTJxki -- Tzx/kl[ < & for k = 1,2 a n d / = 1 . . . . . nk, OWzT~Bx C (Sc2)~2 and such that 2n3 - l e x t W2T~Bxl > 2n2. Continuing in this manner we get sequences {nk}, {6~}, {ek}, {Tk}, {Lk} and {x/k} such that (1) nl < n 2 < . . . . (2) ]ext B/~I - 2n~, k - 1, 2 . . . . . (3) (Sck)ek C (SLk-1)~k-1, e~ < e~-l, k = 2, 3 . . . . . (4) For every symmetric convex body C in the plane such that OC C (Sck)zEk we have lextCI ~> 2nk. (5) {x/~}Tk 1 C Sx and whenever I l u i - T~x~ill < 6k for i - - 1 . . . . . n~, we have O(co{+ui 1i=1) ~k C (OTkBx)~k.
/
(6) IIT~x - T~-lX J II < a j / 2 ~-~ i = 1
nj j -- 1
k-
1
Infinite dimensional convexity
653
By (3), IITkll ~ (1 + el)liT111 for every k and hence, by w* compactness of the unit ball B x , , the sequence {Tk} has a limit point T in the strong operator topology. By using (3) again we get that OTBx C (1 + ek)OTkBx for each k. Also, by (6), IlZxki -- Tkxkill < 6k for every i and k and hence by (5),
T B x ~ co{+Txki }i%, ~ ( 1 - - e k ) B L , . We have thus that O T B x C (SLk)2ek for every k and therefore, by (1) and (4), the set of extreme points of cl T Bx is not finite, i.e., this set is not a polygon. This contradiction completes the proof. D
No infinite-dimensional dual (and in particular reflexive) Ba-
COROLLARY 6.6 ([117]).
nach space is polyhedral. The next proposition is of a general nature but we shall see that it is very useful in the study of polytopes. PROPOSITION 6.7 ([69]). Let V be a CCB symmetric subset of an infinite-dimensional Banach space X such that int V = 0. Then there exists a subspace E C X such that the section V f-) E is infinite-dimensional and compact. PROOF. Without loss of generality we may assume that X is separable. Therefore there exists a countable norming subset {gk} C X* for X. Let Y be the Banach space span V with the norm generated by V as the unit ball and let T : Y --+ X be the natural embedding. Thus the image T By = V is closed and by the H a h n - B a n a c h Theorem T ' X * is a 1-norming linear manifold. Also, it is clear that the restriction of T to any finite codimensional subspace of Y is not an isomorphism. By using these two facts we construct a sequence {Yk} C Sy as follows. Fix a sequence {Ej }oc j=0 of positive numbers tending fast to zero with e0 = 1/2. Take Yl E Sy with []Tyll] < 2 -1 and let hl E ST*X* be such that h l ( y l ) > 1 - E l . Choose y2 E Sy f-) [hi, T*gl] • with ]]Tyzl] < 2 -2 and put E2 -- [Yl, Y2]. Find {h2 . . . . . hm2} C BT*X* such that {hzlE2 . . . . . hmz[E2} is an ez-net for BE~. Take Y3 E Sy A [T*gl, T*gz, hl . . . . . hm2] • with I[Ty3[[ < 2 -3, and continue in an obvious way. A standard calculation shows that {yk } is a basic sequence and that for the partial sum operators we have [ISn I[ <<, 1/(1 - en). From the construction it is clear that limk~oc gj(Tyk/llTykl]) =--0 for every j. Since {gj} is norming we may thus assume that { Ty~ } is a basic sequence (otherwise pass to a subsequence). Next we construct a bounded biorthogonal sequence {y~} C T ' X * for {Yk}. Take y ~ - h l / h l ( y l ) and suppose that the functionals {y~ }~ have already been constructed such that: (1) y~ -- Zn kl akihi, Z n k l laki] <, (2 - ek-1)/((1 -- e k - , ) ( 1 -- ek)), k -- 1 . . . . . n. (2) ([y~17) • D [YJln~l" Put En -- [yj ]~ n -- 1 2,
and let ~o be the functional in E*n+l defined by ~ole. - 0 and ~O(yn+l) - 1. For y -- Y~nl+l biYi E SEn+I we have: ,
,
9
9
9
]go(y)]- Ibn+l I - Ily - Snyll ~ 1 + IISn II ~ 1 + 1/(1 - ~ ) .
654 Since
V.P. Fonf et al.
{hj[En+lImn+l *j=l
is an en+l-net for BE,+I it follows that q9 has a representation
mn+l
mn+l
q9 -- E cjhj[En+l, j--1
[cj[ <<,( 2 - 6 n ) / ( ( 1 - 8n)(1 - 8 n + l ) ) .
E 1
* m ~-'~mn+l Put Yn+l z-.,j=l cjhj. Clearly [[y~ll ~< 3 for all k and the functionals {x~} C X* biorthogonal to the basic sequence xk = TYk/llTyk [I, k -- 1 . . . . . may be taken such that T * x~* -- IlTyklly~. * Put E -- [xk]~__l and take any x -- Y~k=l or akx~ ~ E N V. Then x -- Ty, y ~ B r . We have
l a k l - Ix;(x>l
-
Ix;(ry)l-
Ilry
llly;(y)l <<.3 . 2
-~:,
and this shows that V A E is compact. Obviously it is also infinite-dimensional.
Fq
REMARK. The assumption that V is closed may be weakened [69]: It is enough to assume that V is a G~-set in X and closed in its linear span. COROLLARY 6.8 ([69]). Each symmetric polytope V in a Banach space X has a nonempty interior in its linear span. Therefore, V induces an equivalent polyhedral norm in its linear span. PROOF. Denote X1 = [V] and suppose to the contrary that intxl V -- 0. By Proposition 6.7 there exists a subspace E C X1 such that the section W -- E A V is infinite-dimensional and compact. Let Y be the Banach space span W with the unit ball W and Z be the Banach space of all affine continuous (in the norm topology of X) functions on W which vanish at zero. It is not difficult to see that Z* -- Y and that (since B r = W), Y is polyhedral, which contradicts Corollary 6.6. D For separable polytopes the symmetry assumption may be omitted. THEOREM 6.9 ([69]). Each separablepolytope has a non-empty interior in its affine span. In particular this affine span is closed and is isomorphic to (a translate of) a polyhedral Banach space. Since each closed face of a polytope is itself a polytope, we have COROLLARY 6.10 ([69]). rior in its affine span.
Each closed face o f a separable polytope has a nonempty inte-
Combining Corollary 6.6 with Theorem 6.9, we see that a reflexive space contains no infinite-dimensional polytopes. Moreover, from Corollary 6.8 and Theorem 6.9 it is clear that, if we want to know more about polytopes, we have to pass to the study of polyhedral spaces. The following is a trivial sufficient condition for the polyhedrality of a Banach space.
Infinite dimensional convexity
655
PROPOSITION 6.11. Let the Banach space X admit a 1-norming set B C Sx, with the following property:
(*)
Each w*-cluster point of B of norm 1 (if any) does not attain its norm on the unit ball B x.
Then the set B is a boundary for X and the space X is polyhedral.
For other conditions for polyhedrality and the relations between them see [45] and [71 ]. The simplest Banach space which satisfies the condition in Proposition 6.11 is co; we shall meet other examples later. For the next theorem (which describes the geometric structure of the unit ball of a polyhedral space) we need the following notation. For a functional f ~ Sx, put Ff--{xEX:
f(x)--l},
Vf--FfNSx.
THEOREM 6.12 ([60]). Let X be a polyhedral Banach space with density character co. Then there exists a boundary B for X with card B - - c o (in particular, if X is separable then B is countable) and such that for each f ~ B, intr s yf ~ 0. Recall that the "opening" of two subspaces L and M of a given Banach space X is just the Hausdorff distance between their unit spheres: O(L, M) = max{sup{d(x, SM)" x E SL 1, sup{d(y, So)" y E SM }1.
The set of all subspaces of a Banach space equiped with this metric is a complete metric space. To prove the theorem we need two lemmas. The first is just a consequence of the definition of O(L, M). LEMMA 6.13. Let X be a polyhedral Banach space and L be a finite-dimensional subspace of X. Then there exists an e = e(L) > 0 such that for each subspace M C X with O(L, M) <~c(L), lext BM, I ~> lext BL, I. The second lemma is related to Mazur's theorem on smooth points (see [95]). DEFINITION 6.14. A finite-dimensional subspace M of a Banach space X is called smooth if each smooth point of the sphere SM is a smooth point of the whole sphere Sx. Equivalently, M is smooth iff each f 6 ext BM, has a unique extension to a norm one functional on the whole space X. LEMMA 6.15. Let X be a polyhedral Banach space and L be a two-dimensional subspace of X. Then for each e > 0 there exists a (two-dimensional) smooth subspace M with O(L, M) < e such that if {if/[M}in 1 - - extBM,, fi E Sx,, i = 1 . . . . . n, then {+ fi IL}inl D ext BL,.
V.P. Fonf et al.
656
PROOF. Suppose that L is not smooth. Then there exist two functionals f, g ~ Sx, with f ~ g but f l L = glL E extBc,. Take any y 6 X with f ( y ) ~ g(y) and put G = span{L, y}. G is a three-dimensional polyhedral space and a simple consideration shows that there exists a two-dimensional subspace L1 C G such that: (1) O(L, Ll) < min{e(L)/4, s/4}. (The function s(L) comes from Lemma 6.13.) (2) lext BeT1 > lext BL, I. (3) For each subset A C Sx, with {flL~: f 6 A} D extBc 7 we have {f[L: f 6 A} D ext BL,. If L1 is not smooth, then starting with L1 and repeating the procedure, we construct a two-dimensional subspace L2 which has the properties (1)-(3), where the following substitutions are made: L by L1, L1 by L2, min{s(L)/4, s/4} by min{s(L)/8, s(L1)/8, s/8}, and so on. If in some step we get a smooth subspace Ln, the lemma is proved. Otherwise we construct a sequence of two-dimensional subspaces {Ln }~ with the following properties: (a) For each integer n/> k we have O(Lk, Ln) < s(Lk). (b) For each integer n we have O(Ln, Ln+l) < s/2 n+2. (c) For each integer k we have lext BL;I ~> 2k. It follows from (b) that 0 - limLn = L0 exists and from (a), (c) and Lemma 6.13 it follows that lext BL~I -- cx~, a contradiction. [] REMARK. It is proved in [67] that in a separable polyhedral space X the set of finitedimensional smooth subspaces of X forms a dense Gs set in the set of all finite-dimensional subspaces of X. PROOF OF THEOREM 6.12. Let 99 be the set of all two-dimensional smooth subspaces of X. Put
B = { f ~ Sx*: f l L E extBL,, L 6 qg}. By Lemma 6.15, B is a boundary for X. Let f 6 B, L 6 q9 and x 6 SL \ ext SL be such that f ( x ) = 1. Since L is smooth it follows that for each y 6 Ker f the point x is not an extreme point of Six,y]. By the Baire Category theorem we get that intr I yf ~ ~J and, in particular, that card B ~< co. It may be proved that actually card B = co. V] REMARKS. (i) Theorem 6.12 remains true (with obvious changes) for a non-symmetric polytope. (ii) Since intry yf ~ 0 for every f 6 B it follows that every boundary for X must contain B, i.e., this B is a minimal boundary. (iii) An alternative proof of Theorem 6.12, which uses the Mazur theorem directly is given in [ 166]. COROLLARY 6.16. Every extreme point of a separable polytope is an exposed point of
the polytope. PROOF. Let V be a polytope in X with X separable and 0 an interior point of V. By Theorem 6.12 and Remark (i) above there is a sequence of functionals {fn }n~_-i of norm 1
Infinite dimensional convexity
657
in X* so that for every x E 0 V there is an n with
f n ( x ) = 1.n = sup{ fn(y)" y ~ V}. For each n put Vfn = 0 V N {y 6 X: fn (Y) = )~n}. Let x0 6 ext V and let Ixo -- {n; x0 6 yfn }. We shall show that {xo } = A n ~Ixo vfn and thus
f -- ~_~n~ixo 2 -n fn is a functional which exposes x0. Assume that the intersection contains a point x0 r x l E 0 V. Consider the two-dimensional section of V determined by 0, x0 and xl. This section is a polygon and x0 is an extreme point of it. Let x2 be a (relative) interior point of the edge of this polygon which contains x0 but not Xl. Let m be such that X2 E Yfm" Then clearly m ~ Ixo but x l 6 VJ~, and this contradicts the choice of X l . [-I The following result holds for every boundary of a polyhedral space. PROPOSITION 6.17 ([67,166]). Let X be a polyhedral Banach space and B a boundary f o r X. Then Bx, is the norm closed convex hull of B. By remark (ii) above it is enough to consider just the minimal boundary B. If X is separable then B is countable. Thus Proposition 6.17 gives information beyond that of Theorem 5.7 only for non-separable polyhedral spaces. COROLLARY 6.18 ([60]). separable too.
Let X be a separable polyhedral Banach space. Then X* is
In general the converse to Theorem 6.12 is not valid. However for a countable boundary we have PROPOSITION 6.19 ([58]). If a Banach space X has a countable boundary, then f o r each e > O, X is (1 + e)-isomorphic to a polyhedral space. PROOF. Let B = {-+-hi}~ be a boundary for X. Take a sequence {ei} of positive numbers which tends to zero and ei < e, i - 1,2 . . . . . Put B1 -- {-+-(1 -k- e i ) h i } ~ and V* = w*-cl co B1. Clearly Bx, C V* C (1 + e ) B x , . Since V* is w*-closed it induces an equivalent norm l][.lii on X which is (1 + e)-isomorphic to the given norm. By definition, the set B1 is 1-norming for (X, 101.]16).Let x 6 X, x ~ 0. Since B -- {-+-hi} is a boundary, there exists an i such that I h i ( x ) l - Ilxll, so IIIxlll t> I(1 + ei)hi(x)l > I h i ( x ) l - Ilxll. Let f be any w*-cluster point of the set B1; then since 6i ~ 0, f ~ B x , . Suppose that for some x E X with IIIxlll- 1 we had f ( x ) 1. Since f E B x , it follows that Ilxll/> f ( x ) = 1. Therefore Ilxll >f IIIxlll, which contradicts what we just proved. Thus, no w*-cluster point of the set Bl attains its norm II1.111and by applying Proposition 6.11 we deduce that (X, II1.111) is polyhedral. D COROLLARY 6.20 ([58]). Each predual of the space ll is isomorphic to a polyhedral space whose dual is still isometric to l l.
658
V.P Fonf et al.
REMARKS. (i) No infinite-dimensional C ( K ) space is polyhedral in its natural sup-norm. But, as follows from Corollary 6.20, C ( K ) with K compact and countable is isomorphic to a polyhedral space. (ii) If X* is isomorphic to I i then X need not be isomorphic to a polyhedral space. Indeed, Bourgain and Delbaen [ 18] constructed a separable space X so that X* is isomorphic to l l but X fails to have a subspace isomorphic to co and hence (see Theorem 6.21 below), X is not isomorphic to a polyhedral space. The following theorem (which follows directly from Theorems 5.19 and 6.12) strongly restricts the possibility that a Banach space be isomorphic to a polyhedral space. THEOREM 6.21 ([60]). Let X be a polyhedral Banach space. Then X contains an isomorphic copy of the space co. REMARKS. (i) Theorem 6.21 is of course a strengthening of Corollary 6.6. We thought it instructive to present here in detail the proofs of both of these results. The proof of Corollary 6.6 (or equivalently of Theorem 6.2) is finite-dimensional in nature. The proof of Theorem 6.12 was inspired by the proof of Theorem 6.2; it is more functional analytic in nature. (ii) In [ 114] it is proved that there is an Orlicz sequence space X which has a separable dual and every infinite-dimensional subspace of it contains a copy of co (i.e., X has the properties described in Corollary 6.18 and Theorem 6.21) but is nevertheless not isomorphic to a polyhedral space. We return to Proposition 6.11. As simple examples show, this proposition gives just a sufficient condition for polyhedrality. But in some classes of Banach spaces this condition is also a necessary one. PROPOSITION 6.22 ([77]). Let X be a Banach space with X* = L1 (#) for some measure Iz. Then X is polyhedral if and only if it has the property appearing in Proposition 6.11. Polyhedral spaces play an important role in the class of preduals of L l(#), as the following theorem shows. PROPOSITION 6.23 ([110,119]). Let X be a Banach space with X * - L 1 (#). The following two assertions are equivalent: (1) X is polyhedral. (2) For each pair of Banach spaces Y C Z and for each compact operator T" Y --+ X there exists a compact extension T" Z --+ X with II T II - II T II. Both assertions yield (3) X* - ll ( F ) for some set F. Certainly (3) does not imply (1) isometrically, but isomorphically it does in the following cases:
Infinite dimensional convexity
659
(1) For/-" a countable set. (2) For X = C ( K ) with K a closed interval of ordinals [59]. In general, even isomorphically, (3) does not imply (1): the space C ( K ) , with K the Kunen dispersed compact [133], is not isomorphic to a polyhedral space, while C*(K) = ll (K). (This example was communicated to us by M. Jimenez Sevilla.) We consider next a very natural question about approximating a CCB body by polytopes. DEFINITION 6.24. Let V be a CCB body in a Banach space X such that 0 6 int V. We say that a CCB body P, e-approximates V if V C P C (1 + e) V. Moreover, if in addition each maximal face of P is tangent to V, we say that P is an e-approximating tangent body for V. It is clear that the problem of approximating of a CCB body by polytopes may be posed only in polyhedral spaces. The following theorem gives a full solution in the separable case.
THEOREM 6.25 ([40]). Let V be a CCB body in a separable polyhedral Banach space. Then for each e > 0 there exists an e-approximating tangent polytope. We pass now to a local property of the norm in a Banach space which is strongly connected with polyhedrality. We say that the norm on a Banach space X locally depends on a finite number of coordinates if for each x E X, x --/: 0, there exist a neighbourhood W of x and a finite number of continuous linear functionals {f/}~n such that whenever y, z E W and f / ( y ) = f/(z), i = 1 . . . . . m, we have I[yll = Ilzll. THEOREM 6.26 ([64,85]). Let X be a separable Banach space. The following assertions are equivalent: (1) X has an equivalent norm which locally depends on a finite number of coordinates. (2) X is isomorphic to a polyhedral space. The polytopes discussed above all have nonempty interior and thus in particular are never compact in infinite dimensions. In [ 137] a definition is given of classes of compact convex sets based on the notion of a simplex. DEFINITION 6.27. (1) A convex set K is called an ~-polytope if there exist a compact simplex S and a continuous affine map from S onto K having finite-dimensional fibres. (2) A convex set K is called a fl-polytope if it is a finite codimensional slice of a compact simplex. It is of interest to compare this definition with the observation following Corollary 3.3. No inclusion relation holds between the classes of or- and /3-polytopes in infinitedimensional spaces. In finite-dimensional space the notions of ot-polytope and fl-polytope both coincide with the usual notion of a polytope. For more details see [137].
660
V.P Fonf et al.
7. Miscellaneous topics The subject of infinite-dimensional convexity covers a large number of different directions, most of which we do not have space to even mention in this survey. We chose to present in this section four additional topics in a very brief form. These topics are all related to the material presented in the previous sections. The two last topics contain interesting and challenging open problems. a. Stable convex sets
DEFINITION 7.1. A convex set K in a locally convex space X is called stable if the map (x, y) --+ (x + y ) / 2 is an open map from K • K into K. It is easy to see that an equivalent definition is the following: K is a stable iff for every relatively open subset G of K the set co G is also relatively open in K. Trivially, if K is stable then ext K is closed. Indeed, (assuming X is metrizable) let {Un } C K with Un --+ u q~ ext K, i.e., u = (x + y ) / 2 with x r y in K. By the definition of stability there are sequences {Xn} and {Yn} in K with Xn --+ x, Yn --+ Y and Un = (Xn + Yn)/2. Thus Un q~ext K for large enough n. In some important examples, the fact that ext K is closed already implies stability. For example it is not hard to prove that a Bauer simplex is stable and so is every CCB body K in a Banach space for which ext K = 0 K. However, there are already finite-dimensional compact convex sets K with ext K closed which are not stable. It can be shown that a compact convex n-dimensional set K is stable iff for every m ~< n the m-frame K (m) of K (i.e., the union of all faces of K of dimension ~< m) is closed. For compact convex sets K we have THEOREM 7.2. Let K be a compact convex set in a Hausdorff locally convex space. Then the following three assertions are equivalent: (1) K is stable. (2) For every continuous function f on K the upper envelope f is also continuous. (3) The map from P ( K ) (the Baire probability measures on K in the w*-topology) into K, which assigns to each measure lZ its barycenter, is open. By using the selection Theorem 3.18 and the definition of stability it is easy to prove THEOREM 7.3. Let S be a compact simplex and let K be a stable CCB set in a Banach space. Denote by A(S, K ) the (convex) set of affine continuous maps from S into K. Then T 6 extA(S, K ) i f f T ( e x t S ) C extK. Also the symmetric version of this theorem is valid. In this version S is replaced by B x , in the w*-topology, where X is a space whose dual is an L1 (/z) space, K is assumed to be symmetric and one considers only those affine continuous T which satisfy TO = 0. For more results on stable convex sets as well as references to the literature on this subject, we refer to the survey paper [ 135].
Infinite dimensional convexity
661
b. Sets with dense extreme points It is a purely infinite-dimensional phenomenon that the set of extreme points of a convex set K may be dense in K. We already encountered two such examples in Section 3, namely, the Poulsen simplex and the unit ball of the dual of the Gurarii space in the w*-topology. A much simpler example is the unit ball of 12 in the w-topology. Our aim here is to examine how common this phenomenon is. We start with compact convex sets in a Banach space. A CCB set K in a Banach space X is called strictly convex if every proper support hyperplane of K meets K in a single point. (By a proper support hyperplane we mean a hyperplane which does not contain K.) The support points of proper support hyperplanes of a strictly convex K belong to ext K (actually to exp K) and (by the Bishop-Phelps theorem) are dense in the boundary of K. Thus, for a compact strictly convex K in an infinite dimensional Banach space, ext K is dense in K. THEOREM 7.4 ([104]). In the Hausdorff metric, the set of compact convex subsets of a Banach space X is a complete metric space. The set of strictly convex compact sets is a dense G~ set in this space. Thus the compact convex sets K with cl ext K ~: K form a set of the first category in the space of all compact convex sets. The situation for convex but not necessarily compact sets in a Banach space is clarified in the next result. THEOREM 7.5 ([69]). Let K be a CCB set in a Banach space. Then K is the limit in the Hausdorff metric of convex sets with dense extreme points iff K does not have an interior point in any subspace of finite codimension. Actually, if K has this property, then it can be approximated in the Hausdorff metric by sets in which the strongly exposed points are dense. The basic idea of the proof is to use a method similar to the one described in Section 3 for the geometric construction of the Poulsen simplex. It is also proved in [69] that if K is a CCB set in X so that cl(K - K) does not have an interior point in any subspace of finite codimension, then there is a sequence Kn ~ K so that str exp Kn is dense in Kn, Kn ---->K in the Hausdorff metric and for every n there is a continuous affine retraction from Kn onto K.
c. Convex tiling DEFINITION 7.6. A covering of a normed space by bodies (i.e., sets which are the closures of their nonempty interiors) is called a tiling if the interiors of the bodies are disjoint. A tiling is called convex (respectively bounded) if all the bodies (which are also called tiles) are convex (respectively bounded). A covering (in particular a tiling) is called locally finite if each point in the space has a neighbourhood which meets only finitely many sets of the covering. We call a tiling uniformly bounded if the diameters of the tiles are uniformly bounded. A filing is called uniformly bounded from below if there is a r > 0 so that all tiles contain a ball of radius r.
V.P. Fonf et al.
662
THEOREM 7.7 ([70]). Each normed space has a convex tiling which is bounded and uniformly bounded from below. The proof shows that the tiling can be done in a uniform way for all normed spaces with a given density character. There is a tiling r of l ~ so that for every separable normed space X there is an isomorphism T from X into l ~ so that {T -1 (B): B 6 r} forms a bounded and uniformly bounded from below tiling of X. The situation is similar for all spaces X having a given uncountable density character. If X has the RNP then it is easy to see that each bounded convex body can be tiled by bodies having diameters less than 1, say. Thus, if X has RNP, it follows from Theorem 7.7 that it has a convex uniformly bounded tiling. In the process of cutting down tiles we usually lose the property that they are uniformly bounded from below. The space co has an evident tiling by translates of its unit ball. In other spaces it seems hard to find such nice tilings. In particular we mention the following open problem:
"Does 12 (or any other separable reflexive space) have a convex tiling which is uniformly bounded from below and above ?" If A were a discrete subset of 12 for which IIx - y II/> 1 for x r y in A and for which sup{d(z, A): z 6 12} < ~ , and if in addition A were proximinal (see Subsection d below), then it is easy to check that the Voronoi cells Va = {z 6 12: IIz - all = d(z, A)}, a E A, would form a convex tiling of 12 which is bounded from above and below. The trouble is that, as observed in [69], a proof similar to that of Theorem 5.15 shows that no such A exists. On the other hand, it was shown in [ 107] that if c~ is a cardinal with ots~ -- or, then such a set A (and thus also a tiling uniformly bounded from above and below) exists in 12(F) for a set F of cardinality or. Let us mention here also the following surprising fact: The space ll ( F ) ( F as above) can be tiled by a collection of disjoint (!) translates of its unit ball. For covering a Banach space with convex sets there is the following result. PROPOSITION 7.8 ([36]). For any covering of a reflexive space by CCB sets there is a finite-dimensional parallelotope that intersects infinitely many members of the covering. In particular such a covering is never locally finite. For locally finite tilings we have THEOREM 7.9 ([64]).
For a separable Banach space X the following assertions are
equivalent. (i) X is isomorphic to a polyhedral space. (ii) X has a locally finite convex bounded tiling. d. Proximinal sets and related sets Let K be a set in a Banach space X. A point y 6 K is a nearest point to x E X if ]ix - y I[ --
d(x, K). DEFINITION 7.10. The set K is called proximinal if every x E X has a nearest point in K. The set K is called a Chebyshev set if every x has a unique nearest point in K.
Infinite dimensional convexity
663
Obviously a proximinal set has to be closed. It is trivial to check that a closed convex set in a strictly convex reflexive Banach space is Chebyshev. The question whether conversely, every Chebyshev set is (under suitable assumptions on the space) convex, is a challenging open problem. PROBLEM. Is every Chebyshev set in a smooth Banach space X convex? In particular, what is the case for X = 12? Under some assumptions of compactness or continuity of the map which assigns to each point its nearest point in K it is known that the answer is positive. For example: PROPOSITION 7.1 1 ([ 169]). is convex.
In a smooth Banach space every w-compact Chebyshev set
The survey paper [169] has much more information on this topic. Subsequent to this paper there has been very little progress towards the solution of the problem mentioned above. There is also a well-known problem concerning farthest points. Given a bounded set K C X, we say that y E K is a farthest point from x E X if Ilx - y II = sup{ Ilx - z I[: z E K }. PROBLEM. Assume that each x E X has a unique farthest point qK(x) in K. Must K consist of a single point? Here, too, some partial results are known. In case K is convex and compact it follows trivially from the Schauder fixed point theorem that K consists of a single point. Using this idea and passing from K to co K it is not hard to show (cf. [106]) that if X is strictly convex and II. II-cl K is compact then the answer to the problem is positive. The same holds if the assumption that X is strictly convex is replaced by the assumption that X is smooth (cf. [130]). Also if dim X < c~, the answer is positive [5]. An interesting connection exists between the problem of convexity of Chebyshev sets and the unique farthest point problem, but only in the setting of Hilbert space. PROPOSITION 7.12 ([106]). The following assertions are equivalent: (i) Every Chebyshev subset of a Hilbert space H is convex. (ii) Every subset K of H such that any x E H has a unique farthest point in K consists of a single point. The proof of this equivalence is obtained by exploiting the properties of inversion with respect to a sphere in a Hilbert space. We pass now to the question of existence of nice discrete proximinal sets in a Banach space. The vectors in co all whose coordinates are integers form a nice discrete proximinal set. In the discussion on tiling we mentioned that 12 does not have good proximinal discrete sets. More generally we have PROPOSITION 7.13 ([69]). A separable infinite-dimensional Lp(lZ) space, 1 <~ p < cx~, does not contain a proximinal set A such that Ilx - y ll ~> 1 for all x ~ y in A and sup{d(z, A): z E Lp(/Z)} < cx~.
V.P. Fonf et al.
664
The proof for p # 2 of Proposition 7.13 is considerably harder than the case p = 2. The main part of the argument is similar to the proof of Proposition 5.22. For p # 2 there is no connection between Proposition 7.13 and problems of convex tiling. Proposition 7.13 may very well hold for every separable infinite-dimensional Banach space which has no subspace isomorphic to co. We mentioned above that (for certain uncountable F ) there is a tiling of 11 ( F ) by disjoint translates of the unit ball of this space. The centers of these balls clearly form a discrete Chebyshev set. This is an example of a highly non-convex Chebyshev set in (the nonsmooth) space ll ( F ) . We pass now to the notion of antiproximinal sets. DEFINITION 7.14. A subset K of a Banach space X is called antiproximinal if no point in X \ K has a nearest point in K. The first example of a CCB body which is antiproximinal was constructed in [46] in the space co. In the same paper it was observed that a CCB set K in X is antiproximinal iff (7.1)
r(K)Ar(Bx)={O}.
It follows from this observation and Theorem 4.16 that if X has the RNP, then it does not contain a CCB set which is antiproximinal. Note that if we consider unbounded sets, then even if the space X has the RNP there may be closed convex sets K l and K2 so that there is no linear functional which attains its m a x i m u m on both K1 and K2, though there are many linear functionals which are bounded on both sets. Such an example in case X = 12 may be found in [ 140]. It is known (cf. [16]) that there are spaces X which fail to have the RNP and still do not contain an antiproximinal CCB set. For example, every space with the convex point of continuity property fails to contain an antiproximinal CCB set. (As was mentioned in Section 4, there are such spaces which fail to have the RNE) Among examples of spaces X which do contain antiproximinal CCB sets, we mention in particular the space C[0, 1]. In order to get a characterization of Banach spaces which contain sets of a similar form we define a stronger property (see condition (7.1)). DEFINITION 7.15. A CCB subset K of a Banach space X is called strongly antiproximinal if there exists a bounded one-to-one linear operator T : X -+ Y to some Banach space Y so that
T*(Y*) D r ( K ) , THEOREM 7.16 ([63]).
T*(Y*) A ~7(Bx) = {0}.
Let X be a separable Banach space. The following assertions are
equivalent: (i) X has a subspace isomorphic to co. (ii) A space isomorphic to X contains a strongly antiproximinal symmetric CCB body.
Infinite dimensional convexity
665
The paper [63] contains also several references to other papers on the subject of antiproximinal sets.
References [1] E.M. Alfsen, On the Dirichlet problem on the Choquet boundary, Acta Math. 120 (1968), 149-159. [2] E.M. Alfsen, Compact Convex Sets and Boundary Integrals, Ergeb. Math. Grenzgeb., B. 57, SpringerVerlag, Berlin (1971). [3] E.M. Alfsen and E Shultz, State Spaces of Operator Algebras, Birkhiuser Verlag (2001). [4] L. Asimov and A.J. Ellis, Convexity Theory and its Applications in Functional Analysis, Academic Press (1980). [5] E. Asplund, Sets with unique farthest points, Israel J. Math. 5 (1967), 201-209. [6] H. Bauer, Schilowscher Rand und Dirichletsches Problem, Ann. Inst. Fourier (Grenoble) 11 (1961), 89136. [7] R. Becker, COnes Convexes en Analyse, Travaux en cours 59, Hermann, Paris (1999). [8] Y. Benyamini and J. Lindenstrauss, Geometric Nonlinear Functional Analysis, Vol. 1, Colloquium Publications, Vol. 48, Amer. Math. Soc. (2000). [9] C. Bessaga and A. Petczyfiski, On bases and unconditional convergence of series in Banach spaces, Studia Math. 17 (1958), 151-164. [10] C. Bessaga and A. Petczyfiski, Selected Topics in Infinite-Dimensional Topology, Monografie Matemat. 58, Polish Sci. Pub., Warsaw (1975). [11] E. Bishop, A minimal boundary for function algebras, Pacific J. Math. 9 (1959), 629-642. [ 12] E. Bishop and K. de Leeuw, The representation oflinearfunctionals by measures on sets of extreme points, Ann. Inst. Fourier (Grenoble) 9 (1959), 305-331. [13] E. Bishop and R.R. Phelps, The support functionals of convex sets, Proceedings of Symposia in Pure Mathematics, Vol. 7 (1963), 27-35. [14] E E Bonsall, On the representation ofpoints of a convex set, J. London Math. Soc. 24 (1963), 265-272. [15] V. Borovikov, On the intersection of a sequence ofsimplices, Uspehi Mat. Nauk 7 (52) (1952), 179-180 (Russian). [16] J.M. Borwein, M. Jim6nez Sevilla and J.P. Moreno, Antiproximinal norms in Banach spaces (in preparation). [17] J. Bourgain, On dentability and the Bishop-Phelps property, Israel J. Math. 28 (1977), 265-277. [18] J. Bourgain and F. Delbaen, A class ofspecial 12oo spaces, Acta Math. 145 (1980), 155-176. [19] J. Bourgain and H.P. Rosenthal, Geometrical implications of certain finite-dimensional decompositions, Bull. Soc. Math. Belg. 32 (1980), 57-82. [20] J. Bourgain and M. Talagrand, Dans un espace de Banach rdticul~ solide, la propridtd de Radon-NikodSm et celle de Krein-Milman sont gquivalentes, Proc. Amer. Math. Soc. 81 (1981), 93-96. [21] R.D. Bourgin, Geometric Aspects of Convex Sets with the Radon-Nikod•m Property, Lecture Notes in Mathem. 993 (1983). [22] R.D. Bourgin and G.A. Edgar, Non-compact Simplices in Banach Spaces with the Radon-Nikodym Property, J. Funct. Anal. 23 (1976), 162-176. [23] O. Bratteli and D.W. Robinson, Algebraic Methods in Quantum Mechanics I, Springer-Verlag (1979). [24] A.V. Bukhvalov and A.A. Danilevich, Boundary properties of analytic and harmonic functions with values in Banach spaces, Mat. Notes 31 (1982), 104-110. [25] E Cartier, J.M. Fell and P.A. Meyer, Comparison des mesures portde par un ensemble convex compact, Bull. Soc. Math. France 92 (1964), 435-445. [26] G. Choquet, Theory of capacities, Ann. Inst. Fourier 5 (1955), 131-295. [27] G. Choquet, Existence des reprdsentations intdgrales and moyen des points extremeaux dans les cones convexes, C. R. Acad. Sci. Paris 243 (1956), 699-702. [28] G. Choquet, Unicitd des reprdsentations intdgrales au moyen des points extrgmeaux dans les cones convexes, C. R. Acad. Sci. Paris 243 (1956), 555-557.
666
V.P Fonf et al.
[29] G. Choquet, Le thdorkme des reprdsentations intdgrales dans les ensemble convexes compacts, Ann. Inst. Fourier 10 (1960), 333-344. [301 G. Choquet, Remarks d propos de le ddmonstration de l'unicite' de P.A. Meyer, S6minaire BrelotChoquet-Deny (Theorie de Potential), Paris 6 (1962). [311 G. Choquet, Lectures on Analysis (3 volumes), W. Benjamin, New-York (1969). [32] G. Choquet, H.H. Corson and V.L. Klee, Exposed points of convex sets, Pacific J. Math. 17 (1966), 33-43. [331 G. Choquet and RA. Meyer, Existence et unicit~ des representations intdgrales dans les convexes compacts quelconques, Ann. Inst. Fourier (Grenoble) 13 (1963), 139-154. [341 J.RR. Christensen, Compact convex sets and compact Choquet simplices, Invent. Math. 19 (1973), 1-4. [351 M.D. Contreras and R. PayL On upper semicontinuity of duality mappings, Proc. Amer. Math. Soc. 121 (1994), 451-459. [361 H.H. Corson, Collections of convex sets which cover a Banach space, Fund. Math. 49 (1961/61), 143-145. [371 H.H. Corson and J. Lindenstrauss, On simultaneous extensions of continuous functions, Bull. Amer. Math. Soc. 71 (1965), 542-545. [381 W.J. Davis and J. Lindenstrauss, On total nonnorming subspaces, Proc. Amer. Math. Soc. 31 (1972), 109111. [391 L. de Branges, The Stone-Weierstrass theorem, Proc. Amer. Math. Soc. 10 (1959), 822-824. [40] R. Deville, V.R Fonf and R H~ijek, Analytic and polyhedral approximation of convex bodies in separable polyhedral Banach spaces, Israel J. Math. 105 (1998), 139-154. [411 R. Deville, G. Godefroy and V. Zizler, Smoothness and Renormings in Banach Spaces, Pitman Monographs and Surveys in Pure and Applied Mathematics 64 (1993). [421 E.B. Dynkin, Sufficient statistics and extreme points, Ann. Probab. 6 (1978), 705-730. [431 T. Downarowicz, The Choquet simplex of invariant measures for minimal flows, Israel J. Math. 74 (1991), 241-256. [441 J. Dugundji and A. Granas, Fixed Point Theory, Monografie Matemat. 61, Polish Sci. Pub., Warsaw (1982). [45] R. Duffer and L. Papini, Polyhedral norms in an infinite dimensional space, Rocky Mountain J. Math. 23 (3) (1993), 863-875. [46] M. Edelstein and A.C. Thompson, Some results on nearest points and support properties of convex sets in c0, Pacific J. Math. 40 (1972), 553-560. [471 G.A. Edgar, A non-compact Choquet theorem, Proc. Amer. Math. Soc. 49 (1975), 354-358. [481 G.A. Edgar, Analytic martingale convergence, J. Funct. Anal. 69 (1986), 268-280. [491 G.A. Edgar and R.E Wheeler, Topological properties of Banach spaces, Pacific J. Math. 115 (1984), 317350. [501 D.A. Edwards, On the representation of certain functionals by measures on the Choquet boundary, Ann. Inst. Fourier (Grenoble) 13 (1963), 111-121. [511 D.A. Edwards, S~paration des fonctions rdelles ddfinies sur un simplexe de Choquet, C. R. Acad. Sci. Paris 261 (1965) 2798-2800. [521 D.A. Edwards, Systkmes projectifs d'ensembles convexes compacts, Bull. Soc. Math. France 103 (1975), 225-240. [531 E.G. Effros, On a class of complex Banach spaces, Illinois J. Math. 18 (1974), 48-59. [541 E.G. Effros and J.L. Kazhdan, Applications of Choquet simplexes to elliptic and parabolic boubdary value problems, J. Differential Equations 8 (1970), 95-134. [551 J. Feldman, Representations of invariant measures, 17 pp. dittoed notes (1963). [56] W. Feller, An Introduction to Probability Theory and its Applications, Vol. 2, Wiley (1966). [57] V.P. Fonf, A necessary and sufficient condition for reflexivity of a Banach space in terms of extreme points of the unit ball, Ukrainian Math. J. 5 (1978), 531-533. [581 V.R Fonf, Massiveness of the set of extreme points of the dual ball of a Banach space and polyhedral spaces, Funct. Anal. Appl. 12 (1978), 237-239. [59] V.P. Fonf, One property of spaces of continuous functions on intervals of ordinals, Siberian Math. J. 6 (1980), 230-232 (Russian). [60] V.P. Fonf, Polyhedral Banach spaces, Math. Notes Acad. Sci. USSR 30 (1981), 809-813. [61] V.P. Fonf, Semi-embeddings and G~-embeddings of Banach spaces, Mat. Zam. 39 (1986), 302-307. [62] V.P. Fonf, Weakly extremal properties of Banach spaces, Math. Notes Acad. Sci. USSR 45 (1989), 488494.
Infinite dimensional convexity
667
[63] V.P. Fonf, Strongly antiproximinal sets in Banach spaces, Math. Notes Acad. Sci. USSR 47 (1990), 212217. [641 V.E Fonf, Three characterizations of polyhedral Banach spaces, Ukrainian Math. J. 42 (9) (1990), 11451148. [65] V.P. Fonf, On supportless convex sets in incomplete normed spaces, Proc. Amer. Math. Soc. 120 (1994), 1173-1176. [66] V.E Fonf, On exposed and smooth points of convex bodies in Banach spaces, Bull. London Math. Soc. 28 (1996), 51-58. [67] V.P. Fonf, On the boundary of a polyhedral Banach space, Extracta Math. 15 (2000), 145-154. [68] V.P. Fonf and M.I. Kadets, Two theorems on massiveness of a boundary in reflexive Banach space, Funct. Anal. Appl. 17 (1983), 77-78. [691 V.E Fonf and J. Lindenstrauss, Some results on infinite-dimensional convexity, Israel J. Math. 108 (1998), 13-32. [70] V.E Fonf, A. Pezzotta and C. Zanco, Tiling infinite-dimensional normed spaces, Bull. London Math. Soc. 29 (1997), 713-719. [71] V.E Fonf and L. Vesely, Infinite-dimensional polytopes (in preparation). [72] R. Fuhr and R.R. Phelps, Uniqueness of complex representing measures on the Choquet boundary, J. Funct. Anal. 14 (1973), 1-27. [73] N. Ghoussoub, J. Lindenstrauss and B. Maurey, Analytic martingales and plurisubharmonic barriers in complex Banach spaces, Contemp. Math. 85 (1989), 111-130. [74] N. Ghoussoub and B. Maurey, Gr-embeddings in Hilbert space, J. Funct. Anal. 61 (1985), 72-97. [75] N. Ghoussoub and B. Maurey, A non-linear method for constructing certain basic sequences in Banach spaces, Illinois J. Math. 34 (1990), 607-613. [76] E. Glasner and B. Weiss, Kazhdan's property T and the geometry of the collection of invariant measures, Geom. Funct. Anal. 7 (1997), 917-935. [77] A. Gleit and R. McGuigan, A note on polyhedral Banach spaces, Proc. Amer. Math. Soc. 35 (2) (1972), 398-404. [78] G. Godefroy, Boundaries of a convex set and interpolation sets, Math. Ann. 277 (1987), 173-184. [79] R. Godement, Sur la th{orie des representations unitaires, Ann. Math. 53 (1951), 68-124. [801 M. Goullet de Rugy, Gdomdtrie des Simplexes Lecture Notes, University of Paris 6 (1968). [811 V.I. Gurarii, Space of universal disposition, isotropic spaces and the Mazur problem on rotations of Banach spaces, Sibirskii Mat. Zhurnal 7 (1966), 1002-1013. [82] R. Haydon, A new proof that every Polish space is the extreme boundary of a simplex, Bull. London Math. Soc. 7 (1975), 97-100. [83] R. Haydon, Some more characterization of Banach spaces containing l|, Math. Proc. Cambridge Phil. Soc. 80 (1976), 269-276. [841 R. Haydon, An extreme point criterion for separability of a dual Banach space, and a new proof of a theorem of Corson, Quarterly J. Math. (Oxford) 27 (1976), 377-385. [851 E H~ijek, Smooth norms that depend locally on finitely many coordinates, Proc. Amer. Math. Soc. 123 (1995), 3817-3821. [86] M. Hervr, Sur les reprdsentations int~grales 6 l'aide des points extrdmeaux dans un ensemble compact convex mdtrisable, C. R. Acad. Sci. Paris 253 (1961), 366-368. [871 E. Hewitt and L.J. Savage, Symmetric measures on Cartesian products, Trans. Amer. Math. Soc. 80 (1956), 470-501. [88] B. Hirsberg and A.J. Lazar, Complex Lindenstrauss spaces with extreme points, Trans. Amer. Math. Soc. 186 (1973), 141-150. [89] R.E. Huff and E Morris, Dual spaces with the Krein-Milman property have the Radon-NikodSm property, Proc. Amer. Math. Soc. 49 (1975), 104-108. [90] R.E. Huff and P. Morris, Geometric characterizations of the Radon-NikodSm property in Banach spaces, Studia Math. 56 (1976), 157-164. [91] O. Hustad, A norm preserving complex Choquet theorem, Math. Scand. 29 (1971), 272-278. [92] O. Hustad, Intersection properties of balls in complex Banach spaces whose duals are L! spaces, Acta Math. 132 (1974), 283-313. [931 R.C. James, Weakly compact sets, Trans. Amer. Math. Soc. 113 (1964), 129-140.
V.P F o n f et al.
668
[94] J.E. Jayne and C.A. Rogers, The extremal structure of convex sets, J. Funct. Anal. 26 (1977), 251-288. [951 W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84.
[96] M.I. Kadets and V.E Fonf, Some properties of the set of extreme points of the unit ball of a Banach space, Mat. Zametki 20 (1976), 315-320. [97] J.P. Kahane, Ensembles de Ryll-Nardzewski et ensembles de Helson, Colloq. Math. 15 (1966), 87-92. [98] O.-H. Keller, Die Homoiomorphie der kompakten Mengen im Hilbertschen Raum, Math. Ann. 105 (1931), 748-758. [991 D.G. Kendall, Simplices and vector lattices, J. London Math. Soc. 37 (1961), 365-371. [100] V.L. Klee, Some topological properties of convex sets, Trans. Amer. Math. Soc. 78 (1955), 30-45. [1011 V.L. Klee, Extremal structure of convex sets, Arch. Math. 8 (1957), 234-240. [1021 V.L. Klee, Extremal structure of convex sets, II, Math. Z. 69 (1958), 90-104. [103] V.L. Klee, Some characterizations of convex polyhedra, Acta Math. 102 (1959), 79-107. [104] V.L. Klee, Some new results on smoothness and rotundity in normed linear spaces, Math. Ann. 139 (1959), 295-300. [lO5] V.L. Klee, Polyhedral sections of convex bodies, Acta Math. 103 (1960), 243-267. [lO6] V.L. Klee, Convexity of Chebyshev sets, Math. Ann. 142 (1961), 292-304. [lO7] V.L. Klee, Do infinite-dimensional Banach spaces admit nice tilings?, Studia Sci. Math. Hungarica 21 (1986), 415-427. [lO8] A.J. Lazar, Spaces ofaffine continuous functions on simplices, Trans. Amer. Math. Soc. 134 (1968), 503525. [109] A.J. Lazar, Affine products ofsimplices, Math. Scand. 22 (1968), 165-175. [110] A.J. Lazar, Polyhedral Banach spaces and extensions of compact operators, Israel J. Math. 7 (1969), 357364. [111] A.J. Lazar, Sections and subsets ofsimplices, Pacific J.Math. 33 (1970) 337-344. [112] A.J. Lazar, Extreme boundaries of convex sets in 12, Israel J. Math. 20 (1975), 369-374. [113] A.J. Lazar and J. Lindenstrauss, Banach spaces whose duals are L! spaces and their representing matrices, Acta Math. 126 (1971), 165-193. [114] D.H. Leung, Some isomorphically polyhedral Orlicz sequence spaces, Israel J. Math. 87 (1994), 117-128. [115] ,~. Lima, Complex Banach spaces whose duals are L 1 spaces, Israel J. Math. 24 (1976), 59-72. [116] ,~. Lima, Intersection properties of balls and subspaces in Banach spaces, Trans. Amer. Math. Soc. 227 (1977), 1-64. [117] J. Lindenstrauss, Notes on Klee's paper 'Polyhedral sections of convex bodies', Israel J. Math. 4 (1964), 235-242. [118] J. Lindenstrauss, On operators which attain their norm, Israel J. Math. 1 (1963), 139-148. [119] J. Lindenstrauss, Extension of compact operators, Mem. Amer. Math. Soc. 48 (1964). [120] J. Lindenstrauss, A short proof ofLiapounoff's convexity thoerem, J. Math. Mech. 15 (1966), 971-972. [121] J. Lindenstrauss and R.R. Phelps, Extreme points of convex bodies in reflexive Banach spaces, Israel J. Math. 6 (1968), 39-48. [122] J. Lindenstrauss, G.H. Olsen and Y. Sternfeld, The Poulsen simplex, Ann. Inst. Fourier (Grenoble) 28 (1978), 91-114. [123] J. Lindenstrauss and D. Wulbert, On the classification of the Banach spaces whose duals are L 1 spaces, J. Funct. Anal. 4 (1969), 332-349. [124] V. Lomonosov, A counterexample to the Bishop-Phelps theorem in complex spaces, Israel J. Math. 115 (2000), 25-28. [125] L.H. Loomis, Unique direct integral decompositions on convex sets, Amer. J. Math. 94 (1962), 509-526. [126] W. Lusky, On separable Lindenstrauss spaces, J. Funct. Anal. 26 (1976), 627-635. [127] W. Lusky, The Gurarii spaces are unique, Arch. Math. 27 (1976), 627-635. [128] R Mankiewicz, A remark on Edgar's extremal integral representation theorem, Studia Math. 63 (1978), 259-265. [129] H.B. Maynard, A geometrical characterization of Banach spaces with the Radon-NikodSm property, Trans. Amer. Math. Soc. 185 (1973), 493-500.
Infinite dimensional convexity [1301 [1311 [132] [1331 [1341 [1351 [136] [137] [138] [1391 [1401 [141] [1421 [143] [1441 [145] [1461 [1471 [1481 [149] [150] [151] [152] [153] [154] [155] [156] [157] [158] [1591 [1601 [161] [162] [163] [164] [165] [166] [167] [168]
669
S. Mendelson, The farthest point problem, M.Sc. Thesis, The Technion (1994) (in Hebrew). RA. Meyer, Probability and Potentials, Blaisdell (1966). M.A. Naimark, Normed Rings, Noordhoff, Groningen (1959). S. Negrepontis, Banach spaces and topology, Handbook of Set-Theoretic Topology, K. Kunen and J.E. Vaughan, eds, North-Holland, Amsterdam (1984), 1045-1142. E. Odell and H.R Rosenthal, A double-dual characterization of separable Banach spaces containing ll, Israel J. Math. 20 (1975), 375-384. S. Papadopoulou, Stabile konvexe Mengen, Jahresber. Deutsch. Math.-Verein. 84 (1982), 92-106. R.R. Phelps, Lectures on Choquet's Theorem, Van Nostrand Math. Studies, Vol. 7 (1966). R.R. Phelps, Infinite dimensional convex polytopes, Math. Scand. 24 (1969), 5-26. R.R. Phelps, Dentability and extreme points in Banach spaces, J. Funct. Anal. 17 (1974), 78-90. R.R. Phelps, The Choquet representation in the complex case, Bull. Amer. Math. Soc. 83 (1977), 299-312. R.R. Phelps, Counterexamples concerning support theorems for convex sets in Hilbert space, Canad. Math. Bull. 31 (1988), 121-128. E.T. Poulsen, A simplex with dense extreme boundary, Ann. Inst. Fourier (Grenoble) 11 (1961), 83-87. J. Rainwater, Weak convergence of bounded sequences, Proc. Amer. Math. Soc. 14 (1963), 999. M. Rao, Measurable selection of representing measures, Quarterly J. Math. Ser. 2 22 (1971), 571-572. M.A. Rieffel, The Radon-Nikodym theorem for the Bochner integral, Trans. Amer. Math. Soc. 131 (1968), 466-487. J. Roberts, A compact convex set with no extreme points, Studia Math. 60 (1977), 255-266. G. Rod6, Superconvex analysis, Arch. Math. 34 (1980), 452-462. G. Rod6, Superconvexit6t und schwache Kompaktheit, Arch. Math. 36 (1981), 62-72. M. Rogalski, Opdrateurs de Lion, Projecteurs Bordliens, et Simplexes Analytiques, Publication of Fac. de Sciences Orsay (France) (1967-68). V.A. Rohlin, Selected topics in the metric theory of dynamical systems, Usp. Mat. Nauk 4 (1949), 57-128. English translation: Amer. Math. Soc. Transl. Ser. 2 49 (1966), 171-240. H.P. Rosenthal, Pointwise compact sets of the first Baire class, Amer. J. Math. 99 (1977), 362-378. H.P. Rosenthal, On the Choquet representation theorem, Longhorn Notes, The University of Texas at Austin, 1986-87, Lecture Notes in Mathem. 1332 (1988), 1-32. W. Rudin, Functional Analysis, 2nd ed., McGraw-Hill (1991). D. Ruelle, Statistical Mechanics, Rigorous Results, Benjamin Inc. (1969). J. Saint Raymond, Reprdsentation int~grale dans certain convexes, Sem. Choquet, University of Paris VI, Vol. 2 (1974-75). W. Schachermayer, For a Banach space isomorphic to its square the Radon-Nikod~m property and the Krein-Milman property are equivalent, Studia Math. 81 (1985), 329-339. W. Schachermayer, The Radon-Nikodjm property and the Krein-Milman property are equivalent for strongly regular sets, Trans. Amer. Math. Soc. 303 (1987), 673-687. Z. Semadeni, Free compact convex sets, Bull. Acad. Polon. Sci. 13 (1964), 141-146. Yu.A. Shashkin, The Milman-Choquet boundary and approximation theory, Funct. Anal. Appl. 1 (1967), 170-171. S. Simons, A convergence theorem with boundary, Pacific J. Math. 40 (1972), 703-708. Y. Sternfeld, Characterization of Bauer simplices and some other classes of Choquet simplices by their representing matrices, Notes on Banach Spaces, H.E. Lacey, ed., Univ. of Texas Press (1980), 306-358. E. St0rmer, Large groups ofautomorphisms of C* algebras, Comm. Math. Phys. 6 (1967), 194-204. E. Stcrmer, Symmetric states of infinite tensor products of C* algebras, J. Funct. Anal. 3 (1969), 48-68. S. Straszewicz, Uber exponierte Punkte abgeschlossener Punktmengen, Fund. Math. 24 (1935), 139-143. S.L. Troyanski, On locally uniformly convex and differentiable norms in certain non-separable Banach spaces, Studia Math. 37 (1971), 173-180. J. van Mill, Infinite Dimensional Topology, North Holland, Amsterdam (1989). L. Vesely, Boundary of polyhedral spaces - an alternative proof, Extracta Math. 15 (2000), 213-217. J. Villadsen, The range of the Elliott invariant, J. Reine Angew. Math. 462 (1995), 31-55. G.F. Vincent-Smith, Measurable selections of simplical maximal measures, J. London Math. Soc. (Ser. 2) 7 (1973), 427-428.
670
V.P Fonf et al.
[169] L.E Vlasov, Approximative properties of sets in normed linear spaces, Russian Math. Surveys 28 (1973), 1-66. [170] J. von Neumann, Zur Operatorenmethode in der Klassischen Mechanik, Ann. Math. 33 (1932), 577-642. [171] D.V. Widder, The Laplace Transform, Princeton University Press, Princeton, NJ (1941).
CHAPTER
16
Uniform Algebras as Banach Spaces
T.W. Gamelin Department of Mathematics, The University of California, Los Angeles, CA, USA E-mail: twg@math, ucla. edu
S.V. Kislyakov* POMI, Fontanka 27, 191 O11, Saint Petersburg, Russia E-maih skis @pdmi. ras. ru
Contents 1. Uniform algebras
...............................................
2. Analytic functions on Banach spaces 3. Characterization of proper subalgebras
673
.................................... ...................................
4. Tight subspaces and subalgebras of C(K) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. The Petczyfiski and Dunford-Pettis properties
...............................
6. Absolutely summing and related operators on the disk algebra
......................
675 677 681 685 688
7. Interpolation of Hardy-type subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
694
8. Bourgain projections
699
.............................................
9. Perturbation of uniform algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
702
10. The dimension conjecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
703
References
704
.....................................................
*Supported in part by the Russian Foundation for Basic Research, grant 96-01-00693. H A N D B O O K OF THE GEOMETRY OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 671
672
T. W. Gamelin and S. V. Kislyakov
Abstract Any Banach space can be realized as a direct summand of a uniform algebra, and one does not expect an arbitrary uniform algebra to have an abundance of properties not common to all Banach spaces. One general result concerning arbitrary uniform algebras is that no proper uniform algebra is linearly homeomorphic to a C(K)-space. Nevertheless many specific uniform algebras arising in complex analysis share (or are suspected to share) certain Banach space properties of C ( K ) . We discuss the family of tight algebras, which includes algebras of analytic functions on strictly pseudoconvex domains and algebras associated with rational approximation theory in the plane. Tight algebras are in some sense close to C (K)-spaces, and along with C(K)-spaces they have the Petczyfiski and the Dunford-Pettis properties. We also focus on certain properties of C (K)-spaces that are inherited by the disk algebra. This includes a discussion of interpolation between H P-spaces and Bourgain's extension of Grothendieck's theorem to the disk algebra. We conclude with a brief description of linear deformations of uniform algebras and a brief survey of the known classification results.
Uniform algebras as Banach spaces
673
1. Uniform algebras A uniform algebra is a closed subalgebra A of the complex algebra C (K) that contains the constants and separates points. Here K is a compact Hausdorff space, and A is endowed with the supremum norm inherited from C(K). The algebra A is said to be proper if A ~: C (K). Uniform algebras arise naturally in connection with problems in approximation theory. The main examples of proper uniform algebras come from complex analysis. The prototypical proper uniform algebra is the disk algebra, which we denote simply by Ca, consisting of the analytic functions on the open unit disk in the complex plane that extend continuously to the boundary F. More generally, if K is a compact subset of C n, we denote by A ( K ) , or by C a ( K ) , the algebra of functions continuous on K and analytic on the interior of K. Also, we may consider the uniform closure of the restriction to K of some algebras of "elementary" holomorphic functions, such as analytic polynomials, or rational functions with singularities off K. The uniform closure of the rational functions with singularities off K is denoted by R(K). If K is a compact subset of the complex plane, Runge's approximation theorem asserts that R ( K ) includes the functions that are analytic in a neighborhood of K. It may occur that R ( K ) is a proper subalgebra of C ( K ) even when K has empty interior. Other uniform algebras, associated with a domain D in C n, are the algebra A (D) of analytic functions on D that extend continuously to the closure of D, and the algebra H ~ (D) consisting of all bounded analytic functions on D. Endowed with the supremum norm on D, the algebra H ~ (D) becomes a uniform algebra on the smallest compactification of D to which the functions extend continuously. In the case of the open unit disk A = {JzJ < 1}, we may identify H ~ ( A ) with a closed subalgebra H ~ ( d 0 ) of L ~ (d0) via nontangential boundary values, where dO is the arc-length measure on the unit circle. One of the goals in studying uniform algebras is to use the tools of functional analysis and the Gelfand theory in order to prove approximation theorems or to understand why approximation fails. Mergelyan's theorem, that R ( K ) = A ( K ) whenever K is a compact subset of the complex plane whose complement has a finite number of components, was eventually given a proof, by Glicksberg and Wermer, that depends on uniform-algebra techniques and a less difficult theorem of Walsh on approximation by harmonic functions. We quote two other approximation theorems, whose proofs depend on the algebra structure and on techniques of functional analysis. THEOREM (Wermer [50]). Let A be a (not necessarily closed) algebra of analytic functions on the unit circle F in the complex plane. Suppose that A separates the points of F, and that each function in A extends to be analytic in a neighborhood of F. Then either A is dense in C ( F ) ; or else there is a finite bordered Riemann surface with border F such that the functions in A extend to be analytic on the surface. THEOREM (Davie [ 14]). Let )~I( be the area measure on a compact subset K of the complex plane, and let H~ be the weak-star closure of R ( K ) in L~ Then each f ~ H~ is approximablepointwise a.e. on K by a sequence offunctions fn E R ( K ) such that I[fn no~< Hfi II.
T. W. Gamelin and S. V. Kislyakov
674
In some sense, uniform algebra theory can be regarded as an abstract study of the maxim u m principle for algebras. For an incisive account of this aspect of the theory, see [ 1]. Uniform algebra theory has also served as a source of interesting problems. One such problem, originally raised by S. Kakutani, asked whether the open unit disk A is dense in the spectrum of the algebra H ~ ( A ) . This problem became known as the "corona problem". It was answered affirmatively by L. Carleson (see [22]). While the corona theorem per se has not played a really significant role in analysis, the techniques that were devised to solve the problem have played an important role in function theory. The question arises as to the extent that properties of various uniform algebras depend only on their linear structure. From the point of view of Banach spaces, how special are uniform algebras? We will see in Section 2 that genetic uniform algebras are as bad as generic Banach spaces, in the sense that any Banach space is isometric to a complemented subspace of a uniform algebra. On the other hand, we will show in Section 3 that a proper uniform algebra is distinct as a Banach space from C ( K ) . In fact, no proper uniform algebra is linearly isomorphic to a complemented subspace of a C(K)-space, or even to a quotient of a C(K)-space. It is of interest to know which properties of C ( K ) are inherited by uniform algebras. Towards answering this question, a great deal of effort has gone into determining which properties of C ( F ) are passed down to the disk algebra CA. Some of the known results are summarized in the table below. At present there is no unified theory but rather only fragmented results on uniform algebras as Banach spaces. Our aim in this article is to present a selection of results in order to give an idea of what has been studied and what problems are currently open. Sometimes proofs or indications of proofs are also given, to convey the flavor of the techniques that are employed. Section 2 includes a brief introduction to algebras of analytic functions on domains in Banach spaces. In Section 3 we show how some basic facts about p-summing and p-integral operators lead to several characterizations of proper uniform algebras. Sections 4 and 5 are devoted to certain properties that involve weak compactness. The properties are well-known for C(K)-spaces, and they are inherited by algebras generated by analytic functions on planar sets and on strictly pseudoconvex domains, where the key ingredient is the solvability of a 0-problem (Table 1). Table 1 Property of X = C(K) or X
=
CA
X is a (linear) quotient of C (S) X** is complementedin a Banach lattice X* has the Dunford-Pettis property X has the Petczyfiski property X* is weakly sequentially complete X has a basis X verifies Grothendieck's theorem X has a complementedcopy of C (F)
C(K)
CA
Reference
Yes Yes Yes Yes Yes Yes if K is metric Yes Yes if K is metric and uncountable
No No Yes Yes Yes Yes Yes Yes
w w w w w [51] w [51]
Uniform algebras as Banach spaces
675
In Sections 6 through 8 we talk of properties that are more specific to the disk algebra CA. We focus on Bourgain's extension of Grothendieck's theorem to the disk algebra obtained in [8,9]. This subject was treated in detail in the survey [30]. The exposition here will follow the ideas of [30] with slight simplifications at some points, and with emphasis on the interpolatory nature of the proofs. The extension of Grothendieck's theorem is discussed in Section 6, and the results on interpolation used in the proofs are dealt with in Section 7. Section 8 includes a brief account of Bourgain projections, the main technical tool used by Bourgain to transfer results from continuous to analytic functions. A good deal of the material presented here has already been discussed in various monographs and expository papers. We mention particularly the early lecture notes of Petczyfiski [39], and the research monographs of Wojtaszczyk [51] and of Diestel, Jarchow, and Tonge [15]. For background on uniform algebra theory, see [19,49], and [1]. For Davie's theorem, see also [ 18]. The basic Hardy space theory that we refer to is covered in [17,22], and [24]. One reference for several complex variables and pseudoconvexity is [42]. The expository papers [30] and [31] also cover in part the material of the present article. We collect here some standard notation and conventions. We denote by F the unit circle {Iz[ = 1} in the complex plane, and by dm = d0/2zr the normalized arc-length measure on F. The Hilbert transform on L 1(dm) is denoted by 7-/. We denote by CA the disk algebra, which is the uniform closure of the analytic polynomials in C (F). The algebra of bounded analytic functions on a domain D is denoted by H~C(D). If A is a linear space of functions on K, and cr is a measure on K, then HeC(A, or) will denote the weak-star closure of A in L ~ (~). The generic Banach space, or quasi-Banach space, is denoted by X. The closed unit ball of a Banach space X is denoted by Bx, and the open unit ball by B~. The image of a Banach space X in its bidual X** under the canonical embedding is denoted by X. A
2. Analytic functions on Banach spaces
We wish to develop some examples of algebras of analytic functions defined on domains in a Banach space. A complex-valued function on an open subset of a Banach space X is analytic if it is locally bounded and its restriction to every complex one-dimensional affine subspace of X is analytic. In other words, f is analytic on D if f is locally bounded, and if for every x0 c D and direction x E X, the function )~ ~-~ f (x0 + )~x) depends analytically on )~. Sums, products, and uniform limits of analytic functions are analytic. A locally bounded function f on D is analytic just as soon as its restriction to D fq Y is analytic for every finite-dimensional subspace Y of X. Thus any statement about analytic functions that involves only a finite number of points of X will hold in general once it holds for analytic functions of several complex variables. Let f be analytic on a domain D in a Banach space X. Suppose 0 6 D, and suppose If(x)[ <~ C for [[xl] < r. For fixed x ~ X, the function ~" ~ f ( ~ x ) has a Taylor series
T W. Gamelin and S. V. Kislyakov
676 expansion O0
f (~x) -= ~
Am(x)~ m,
14"1 < r/llxll.
m=0
One checks easily that each Am(x) is m-homogeneous, that is, Am()~x) =- )~mAm(x). From the Cauchy estimates we have JAm(x)] ~ C]]xllm/r m, and consequently Am(x) is locally bounded. If we restrict the expansion f (x) = y~ Am (x) to a finite dimensional subspace of X, we obtain the usual expansion of an analytic function as a series of m-homogeneous polynomials. In particular, Am (x) depends analytically on x in any finite-dimensional subspace of X, and since it is locally bounded, it is analytic. Let Pm = Pm (X) denote the space of analytic functions on X that are m-homogeneous. We endow Pm with the supremum norm over the unit ball Bx of X, and then Pm becomes a Banach space. The Cauchy estimates show that the correspondence f F-~ Am is a normone projection from the space H~176 of bounded analytic functions on the open unit ball of X to Pro. The first Taylor coefficient in the expansion of an analytic function f at 0 coincides with the usual Frdchet derivative f1 (0) of f at 0, whose defining property is that f (x) = f(O) + f'(O)(x) + o(]]x[]) as x --+ 0. The Fr~chet derivative f ' ( 0 ) is a continuous linear functional on X, and the space P1 coincides with the dual space X* of X. THEOREM 2.1 (Milne [34]). Any Banach space is isometric to a complemented subspace
of a uniform algebra. To prove the theorem, we let K = Bx, be the closed unit ball of the dual space X* of X, endowed with the weak-star topology. Recall that X denotes the canonical image of X A in X**. The restriction of X to K is a closed subspace of C(K). Let A be the uniform algebra on K generated by X. The functions in A are analytic on the open unit ball of X*, hence have Taylor expansioAns f(x*) = ~ Am(x*). The norm-one projection f e-->A1 into P1 (X*) is the identity on X, and it projects any m-homogeneous polynomial in elements of X to 0 if m -r 1. Thus if we pass to uniform limits of sums of polynomials in elements of X, we obtain a norm-one projection of A onto X ~ X. We could as well obtain the same result by considering the algebra A(Bx,) of weakstar continuous functions on Bx, that are analytic on the open unit ball of X*. It is not difficult to check that the projection of A(Bx,) onto P1 (X*) -~ X** maps A(Bx,) onto linear functionals that are weak-star continuous on Bx,, thus onto X. However, it is shown in [3] that the algebra A(Bx,) need not coincide with the algebra generated by the weakstar continuous linear functionals. Along these lines, it is not even known whether the maximal ideal space of A (Bx,) coincides with Bx,. There is an expanding literature about polynomials on Banach spaces and about uniform algebras associated to Banach spaces (see [16,23]). The study of polynomials focuses on the spaces Pm, which can be viewed as spaces of multilinear functionals on X. Every continuous m-homogeneous polynomial f on X is the restriction to the diagonal of a
677
U n i f o r m a l g e b r a s as B a n a c h s p a c e s
unique continuous symmetric m-linear functional F on X • ... • X. This F is given by the polarization formula
F(Xl . . . . . Xm) --
1 m!2 m
E
e l " " " gm f ( g l X l + " " " -'[- F,m X m ) ,
the summation being extended over the 2 m independent choices of 6 j = 4-1 (exercise). The same formula shows that the norm in Pm is equivalent to the multilinear functional norm, though the spaces are not isometric in general. The space Pm can also be viewed as the dual space of the m-fold symmetric projective tensor product of X with itself. For background, see [20,35].
3. Characterization of proper subalgebras In this section we shall prove that if a uniform algebra is proper, then it differs as a Banach space from any space C (K). The crux of the proof is that every absolutely summing operator from C(K) to a reflexive space is compact, while we construct on any proper uniform algebra an operator to ~2 that is absolutely summing but not compact. Let X and Y be Banach spaces, and let 0 < p <~ q < ~ . An operator T :X --+ Y is said to be (q, p)-summing if 4Csup
{tz*
t
Ix (xj)lp 1/p. x 6
Ilx l l 4 1
/
(3.1)
for every finite collection {X j} of elements of X. The best constant C is denoted by 7 r q , p ( T ) . The class of such operators forms an operator ideal, in the sense that precompositions and postcompositions with bounded operators remain within the class, and further the usual estimates for norms hold. If p = q, (3.1) coincides with the definition of a p-summing operator (see Basic Concepts); in this case the notation 7rp(T) is used. By absolutely summing we mean 1summing 9 We remind the reader that the p-summing operators are characterized as those that factor through a part of the inclusion L ~ ( # ) r L p (/z), in the sense that T can be represented as a composition T 9X
u>
M
r
Mp
v>
y,
(3.2)
where # is a probability measure, M is a subspace of L ~ ( # ) , and Mp is a subspace of L p (/z) containing M, or alternatively Mp is the closure of M in L p (/z). For p ~> 1, T is said to be strictly p-integral if it factors through the entire inclusion L ~ ( # ) r LP(lz), that is, we can take M -- L ~ ( / z ) and Mp = LP(I z) in (3.2). This notion differs slightly from that of a p-integral operator, as defined in Basic Concepts. However, the two notions coincide if, say, Y is reflexive. As explained in Basic Concepts, for p ~> 1 every p-summing operator on C(K) is strictly p-integral, so that the p-summing and the strictly p-integral operators on C (K) coincide.
678
T. W. Gamelin and S. V. Kislyakov
We recall that any absolutely summing operator is weakly compact. For 1-integral operators, we can use the Dunford-Pettis property of L 1(#) to say more. LEMMA 3.1. A (strictly) 1-integral operator from X to a reflexive Banach space is compact. In particular, an absolutely summing operator from C (K) to a reflexive Banach space is compact. To see this, consider the factorization (3.2) above, with p - 1, M -- L ~ ( # ) , Mp = L p (#). Since Y is reflexive, V is weakly compact. By the Dunford-Pettis theorem, V maps weakly compact subsets of L 1(#) to norm-compact subsets of Y. Since bounded subsets of L ~ (#) are weakly precompact in L 1 ( # ) , the composed operator T maps bounded subsets of X into norm-compact subsets of Y, and T is compact. The prototype for an absolutely summing operator that is not 1-integral is the Paley operator P on the disk algebra CA. The Paley operator assigns to a function f on the unit circle F the sequence of 2kth Fourier coefficients {f(2k)}~_l . Paley's inequality is
cxz
)1/2 CEil fill,
k=l
fell
l(m),
for some constant ce > 0. In other words, the restriction of the Paley operator P to H 1(m) is a bounded operator from H 1(m) to g2. For the proof, see [24,53]. Let M be the closed linear span in L2(m) of the exponential functions exp(i2k0), k ~> 1. It is a classical fact (see [53]) that the LP-norms on M are equivalent, for 0 < p < oo. Further, for 1 < p < oo there is a continuous projection Qp of LP (m) onto M. Thus for 1 < p < oo, we can factor the Paley operator on C ( F ) through the inclusion L~~ LP(kt),
P'C(I-') ~
L~
~
LP(m) QP> M
> g2,
and P operating on C ( F ) is p-integral for 1 < p < oo. On the other hand, P is not compact, so P is not absolutely summing on C (F). The story is different if we restrict P to CA. Paley's inequality yields the factorization P ),~A ~ - +n"tl ' m "
v
2
where V is the Paley operator on Hi(m). Thus P is absolutely summing o n CA. On the other hand, P maps the exponential functions exp(i2k0), k ~> 1, to the standard basis vectors of g2, so that P is not compact, and P is not 1-integral. This shows incidentally that CA is not complemented in C ( F ) , nor even isomorphic to any quotient space of a C(K)-space, or else the composition of the projection and P would produce an absolutely summing operator on C (K) that is not compact. Our aim is to transfer the Paley operator and this final observation to an arbitrary uniform algebra. Paley's inequality transfers directly, as follows.
Uniform algebras as Banach spaces
679
LEMMA 3.2. Let A be a uniform algebra on K. Let f E A satisfy Ilfll ~ 1, and let r be any measure on K orthogonal to A. Then the sequence
P(r)-{f
f2k dr ]k~__l
belongs to ~2, and IIP(v)II2 ~ cP Ilvll, where cp is the best constant for Paley's inequality. The estimate persists for f E H ~ ( I r l ) , the weak-star closure of A in L ~ ( I r l ) .
PROOF. Define U ' C ( F ) --+ C ( K ) by (Ug)(s) - ~ ( f ( s ) ) for g E C ( / ' ) and s E K, where is the Poisson integral of g. Then IIUII ~< 1, U(z n) -- f n , and U(~ n) = f n . Now U * ' M ( K ) --+ M ( F ) sends A • to C ) , which by the E and M. Riesz theorem is identified with H~(m). Thus U * ( r ) - h d m for some h E H i ( m ) , and
J K j g2k dr -- f r ~2kh(•) dm,
k~>l.
Paley's inequality for h yields IIP(v)II2 ~ cpIIhlll = cpIIU*(v)ll ~ cpIIvll. This proves the first statement of Lemma 3.2, and the second is obtained by applying the first to the uniform algebra H ~ ( I r l ) and noting that r generates a functional on L ~ ( I r l ) orthogonal to H ~ ( I r l ) . D THEOREM 3.3 (Kislyakov [29]). If A is a proper uniform subalgebra of C(K), then there is an absolutely summing operator from A to ~2 that is not compact, hence not 1-integral. The proof depends on the Paley operator associated with an extremal function F for a certain dual extremal problem. Since A is proper, there is a measure/z on K such that # 2- A but the complex conjugate /2 o f / z is not orthogonal to A. We assume that the functional f ~ f f d/2 on A has unit norm. By the Hahn-Banach and Riesz representation theorems, there is a measure )~ on K such that IlZll = 1 and )~ - / 2 2_ A. Let {fn} be a sequence of functions in A such that IIf~ II ~< 1 and f fn d/2 --+ 1, and let F E H ~ (llzl + I~1) be a weak-star limit point of the sequence {fn}. Then IFI ~< 1, and f F d)~ = 1, from which it follows that IFI-- 1 a.e. d)~. Now f()~ - / 2 ) 2_ A for all f E A, hence for all f E H ~ ( I # I + I~1). In particular, F2()~ - / 2 ) 2_ A, and r = / z + F2()~ - / 2 ) 2_ A. We define
gdr
/
,
gEA.
(3.3)
k=l
By Lemma 3.2, applied to F E H ~ (I/~1 + I~1) and the orthogonal measure gr, the sequence T(g) is square summable and IIT(g)ll2 ~< cllgrll -- c f Igl dlrl. Thus T can be factored through the closure H j (Irl) of A in L1 (Irl), T'A ~
Hl(Irl)
> g2,
T. W. Gamelin and S. V. Kislyakov
680
and T is absolutely summing. To see that T is not compact, we compute the kth component of T ( F 2k-1) (to be rigorous, rather we must consider l i m n ~ T(f2k-1))"
T(F2~-~)~= f F2~-~-ff2~dr--f F2k+~-ff2kdX + f F2~-~-ff2*d# _ f F2k+~-ff2kdfz.
(3.4)
Let E be the set on which ]F] -- 1. Since )~ is carried by E, the first integral on the right is f F d)~ - 1. Since ]F]n --+ 0 off E as n --+ cx~, (3.4) tends to
l + f E f f d # - f E F d [ z - - l + 2 i I m ( f E f f d# ),
(3.5)
which is not zero. Since the kth components of the vectors T (Fn), n ~ 1, do not tend to zero uniformly in n, the vectors T (F n) do not lie in a compact subset of ~2, and then neither does the image of the unit ball of A under T. Thus T is not compact, and by Lemma 3.2, T is not 1-integral on A. If we analyze the proof of Theorem 3.4, we find that it extends to any closed subspace B of a uniform algebra A providing there is a function f E B such that f A _ B while f ~ A. Indeed, let I be the set of all f satisfying f A ___B. This is a closed ideal in A. We choose # _1_A such that/2 generates a norm-one functional on I, and we proceed as before. For some time it was an open problem, known as the Glicksberg problem, as to whether a proper uniform algebra on a compact space K can be complemented in C (K). Theorem 3.4 settles the Glicksberg problem, and it does even more. THEOREM 3.4. If A is a proper uniform subalgebra of C(K), then A is not isomorphic to a quotient ofa C(J) space. In particular, A is not complemented in C(K). Indeed, suppose A is the quotient of C (J). If we compose the operator T from Theorem 3.3 with the quotient map, we obtain an operator
C(J)
>C ( j ) / Z ~ A
T> ~2
that is absolutely summing, hence compact, by Lemma 3.1. Since the projection is an open mapping, the operator T must be compact, and this contradicts Theorem 3.3. We mention some further conclusions that can be drawn from this circle of ideas. Recall (see Basic Concepts) that a Banach space X has Gordon-Lewis local unconditional structure (GL 1.u.st.) if, roughly speaking, its finite dimensional subspaces are well embeddable in spaces with unconditional basis. This occurs if and only if X** is a complemented subspace of a Banach lattice. Thus C (K) has GL 1.u.st., as do all LP-spaces, 1 ~< p <~ o~. THEOREM 3.5. If A is a proper uniform subalgebra of C(K), then A does not have GL
1.u.st.
Uniform algebras as Banach spaces
681
The idea of the proof is to look for a factorization of the Paley operator P ' C A ~ s through A** with the help of the second adjoint of the operator T defined in the proof of Theorem 3.3,
P'CA
U> A** T**> s
However, in this way we only obtain an operator quite similar to P (but not P itself). We redefine F to be a weak-star limit point of the sequence {fn } in A**, and we use the realization of A** as a weak-star closed subspace of C** described in the next section. The operator T** may still be defined by (3.3), where now g belongs to A**, and the functions being integrated in (3.3) are the projections of the functions in C** into L ~ (It 1). The correspondence p ( Z ) ~ p ( F ) mapping a polynomial in the coordinate function Z to p ( F ) is of norm at most 1, hence extends to a bounded operator U from CA into A**. Now the hypothesis of GL 1.u.st. implies by the Gordon-Lewis theorem (see [ 15, Theorem 17.7]) that the absolutely summing operator T** factors through an L 1-space, and we obtain
T**U "CA u > A** v > L 1(v) w>
s
.
By Bourgain's extension of the Grothendieck theorem (Theorem 6.5), the composition V U mapping CA into an L 1-space is 2-summing, hence weakly compact. By the DunfordPettis theorem, the weakly compact operator W maps weakly compact subsets of L l(v) to norm-compact subsets of s Thus the composition WVU is compact, contradicting the noncompactness of T**U (see (3.4)-(3.5)). Along similar lines, it can also be proved that if a proper uniform algebra A is a quotient of a Banach space X having GL 1.u.st., then X contains a complemented copy of 11. The crucial observation here is that if X fails to have a complemented copy of 11, then every operator from X to L 1 (13) is weakly compact. To see this, combine the Petczyfiski property of L ~ (v) with [33, Proposition 2.e.8]. In another direction, Garling [21 ] showed that the dual A* of a proper uniform algebra is not a subspace of the dual of a C*-algebra. The proof is modeled on an earlier argument for CA and uses the basic objects ()~, #, and F) appearing in the proof of Theorem 3.4.
4. Tight subspaces and subalgebras of C (K) The classical Hankel operator corresponding to a function g on the unit circle operates from HZ(d0) to HZ(d0) • sending f to g f - P ( g f ) , where P is the orthogonal projection from LZ(d0) onto H2(d0). The Hankel operator is equivalent to the operator f --> g f + H 2 from H 2 to the quotient space L 2 / H 2. The analogue of these operators, acting on subspaces of C (K), has proved a key to understanding uniform algebras. Let A be a closed subspace of C(K). To each g ~ C ( K ) we associate a generalized Hankel operator Sg from A to the quotient Banach space C ( K ) / A by
S g f - - g f -+-A,
f eA.
682
T.W. Gamelin and S. V. Kislyakov
We say that A is a tight subspace of C(K) if the operators Sg are weakly compact for all g ~ C(K). We say that A is a compactly tight subspace of C(K) if Sg is compact for all g E C(K). Tightness was introduced in [12]. Our discussion is based on that paper, and on [46,47]. We will use the representation of the bidual C** of C = C(K) as a uniform algebra. This representation is realized as follows. The dual space of C is the space M(K) of finite (regular Borel) measures on K, with the total variation norm, and this can be regarded as the direct limit of the spaces L 1 ( # ) , # E M(K). The bidual C** is then represented as the inverse limit of their dual spaces L ~ ( # ) , # E M(K). A "simple-minded" way to express this is to say that each element F 6 C** determines for each # 6 M(K) a function Fu E L~ and these satisfy the compatibility condition that Fv = Fu almost everywhere with respect to v whenever v << #. Conversely, each uniformly bounded compatible family {Flz }, Fu 6 L ec (#), determines an element of C**. The norm of F in C** is the supremum of the norms of Fu in Lee(#). The multiplication in the spaces Lee(#) determines an obvious multiplication in C**. Let A be a subspace of C. Recall that H~ #) denotes the weak-star closure of A in L ec (/z). The bidual A** can then be identified with the weak-star closed subspace of C** consisting of F E C** such that Fu E HeC(A, #) for all /z 6 M(K). The bidual of the quotient space C/A is isometric to C**/A**, and the canonical embedding maps C/A isometrically onto C~ A**. In particular, C~ A** is a closed subspace of C**/A**, and consequently A** + C is a closed subspace of C**. Now the operator Sg is weakly compact if and only if the image of A** under Sg* is contained in the canonical image of C/A. Identifying C with its canonical image in C**, we see that for g 6 C,
Sg is weakly compact
r
gA** _ A** + C.
(4.1)
From this it follows that the g's for which Sg is weakly compact form a closed subalgebra of C(K). Thus A is tight just as soon as Sg is weakly compact for any collection of g's that generates C (K) as a uniform algebra. If A is a subalgebra of C, then each space H ~ ( A , #) is an algebra, and A** is a subalgebra of C**. In this case we obtain from (4.1) the following theorem. THEOREM 4.1. A uniform algebra A on a compact space K is a tight subalgebra of C(K)
if and only if A** + c ( g ) is a closed subalgebra of C(K)**. For algebras of analytic functions, tightness is related to solving a 0-problem. Roughly speaking, the connection is as follows. Functions that belong to an algebra of analytic functions A are characterized as the functions f satisfying 0 f - 0. Suppose that 0-1 is a solution operator for the 0-problem, which need not be linear. Let g be a smooth function. If f 6 A, then from the Leibnitz rule we obtain O(fg) - fO(g) + gO(f) - fO(g). This shows that the one-form fO(g) is 0-closed. We apply the solution operator and obtain a function h -- ~ - l ( f ~ g ) satisfying Oh -- O(fg), so that h - f g ~ A, and S g f - h + A. Thus the action of Sg on f amounts to multiplying f by 0g, applying ~ - l , and projecting into the quotient space C/A. It follows that if there is a weakly compact solution operator
Uniform algebras as Banach spaces
683
for the 0-problem, then each Sg is weakly compact, and A is tight. By the same token, if there is a compact solution operator for the 0-problem, then A is compactly tight. If D is a bounded strictly pseudoconvex domain in complex n-space with smooth boundary, the 0-problem can be solved by means of integral operators, with H61der estimates on the solutions, so that there are compact solution operators; see [42]. The argument outlined above can be made precise. It shows that the algebra A(D) associated with any such domain is compactly tight. The connection between tightness and solving the 0-problem is not completely understood, but our line of reasoning does establish the following theorem. THEOREM 4.2. Let D be a bounded domain in complex n-space. Suppose there is a weakly compact subset E of C( D ) such that the equation Oh - co on D has a solution h E E for every O-closed smooth (0, 1)-form co on D that extends continuously to D and satisfies IicoI]oc ~< 1. Then A(D) is a tight subalgebra of C( D ). If E is compact, then A(D) is compactly tight. If D is strictly pseudoconvex with smooth boundary, the 0-problem solution techniques can be used to show that any f E H~ can be approximated pointwise on D by a bounded sequence of functions in A (D) that extend analytically across 0 D, with uniform convergence on D if f E A (D). In the strictly pseudoconvex case, every point p in 0 D is a peak point for A(D), that is, there is f E A(D) satisfying f ( p ) = 1 and Ifl < 1 on D\{p}. Thus the following theorem applies to strictly pseudoconvex domains with smooth boundaries. THEOREM 4.3. Let D be a bounded domain in complex n-space for which the O-problem is solvable as in Theorem 4.2, and let ~ be the volume measure on D. Suppose that the functions in A (D) that extend analytically across OD are pointwise boundedly dense in H~ Then A(D)* is the direct sum of L I ( ~ ) / A ( D ) • and an Ll-space, and the bidual A(D)** is isometrically isomorphic to the direct sum of H~ and an L~ Further, if every point of OD is a peakpoint for A(D), then HOe(D) + C(D) is a closed subalgebra of L~ and A(D)** + C(D) is isometrically isomorphic to the direct sum of H~ + C(D) and an L~ The idea of the proof is as follows. Let Bs be the band of measures on D that are singular to every measure in A(D) • and let Ba be the band of measures generated by A(D) • There is a direct sum decomposition M ( D ) -- Ba (~ Bs, with a corresponding decomposition A(D)** -- H~ @ L~C(Bs). We claim that the summand H~ is isometrically isomorphic to H~ We regard H~C(D) as a subalgebra of L~ If F E A(D)**, there is a bounded net {fa} in A(D) that converges weak-star in A(D)** to F. Then {fa } converges weak-star to F~ in L ~ (~). Since the fa's are uniformly bounded, they are equicontinuous at each point of D, and consequently any limit function is analytic on D. Thus F~ E H ~ (D). The hypothesis of pointwise bounded density implies that the projection F ~-> F~ maps A(D)** onto H~ Suppose F E A(D)** satisfies F~ --0. Let # E A(D) • Let g be a smooth function, and let {fa } be a bounded net in A(D) that converges weak-star to F. Then fa --+ 0 on D. Choose ha E E such that O(fag) 0(ha) on D. Passing to a subnet, we may assume that ha --+ h weakly, where h c C(D).
684
T. W. G a m e l i n a n d S. V. K i s l y a k o v
Then gfa - ha is analytic on D, and in the limit, h is analytic on D. Now gfa + h ha --+ gFu weak-star in Lee(#). Thus # 2_ gFu, this for all smooth functions g, so that Fu = 0. It follows that F~ = 0 for all measures v ~ Ba, and consequently the projection of F in Hee(Ba) is 0. Thus the algebra homomorphism Hee(Ba) --+ Hee(D) is one-toone and onto. Since any homomorphism of uniform algebras that is one-to-one and onto is an isometry, Hee(Ba) is isometric to Hee(D). The final statement of the theorem, that Hee(D) + C(D) is isometric to a direct summand of A(D)** + C(D), is equivalent to the statement that Hee(Ba) --+ Hee(D) is a "local" isometry, in the sense that it is an isometry at every point of 0 D. An easy way to guarantee this is to assume that every point of 0 D is a peak point for A (D). Another class of examples of tight algebras are the algebras R(K) and A(K) associated with a compact subset K of the complex plane. For these, the solution operator for the 0-problem is the Cauchy transform operator
(~_lh)(() _ __1 Jr
f f h(z) dx dy, ( ~ K, z -- (
which is a compact operator on C (K). Again the line of reasoning outlined above, together with a few technical details, establishes the following theorem. THEOREM 4.4. Let K be a compact subset of the complexplane. Then the algebras R(K) and A(K) are compactly tight. If ~r is the area measure on K, and A is either of these algebras, then A* is isometric to the direct sum of Ll (cr)/A • and an Ll-space. The bidual A** is isometrically isomorphic to the direct sum of Hee (cr) and an Lee-space. Finally, Hee(cr) + C(K) is a closed subalgebra of Lee(~r), and A** + C(K) is isometrically isomorphic to the direct sum of Hee(cr) + C(K) and an Lee-space. Here Hee (or) is the weak-star closure of A in Lee(cr). In the case of A(K), the measure cr can be taken to be the area measure on the interior of K, or the harmonic measure on the boundary of the interior of K. In the case of R(K), cr can be taken to be the area measure on the set of nonpeak points of R(K), which serves in some sense as an interior for K with respect to R(K). The proof of Theorem 4.4 is similar to that of Theorem 4.3, except that Davie's theorem is used to obtain the isometric isomorphism of Hee(Ba) and Hee (or). The proof of the final statement depends upon estimating solutions of the 0-problem for some specific bump functions. Summarizing, we can say that very many standard uniform algebras of analytic functions are tight. We turn to an example of a tight subspace that is not an algebra and that has a different flavor. Let Uc be the space of continuous functions f (e i0) on the unit circle F for which the symmetric partial sums Tn f -- Y]n n f ( k ) eik~ of the Fourier series of f converge uniformly. Normed by Illfill = sup IITn f llee, the space Uc becomes a Banach space. We may regard Uc as a subspace of a C ( K ) - s p a c e as follows. For each n, 0 ~< n ~< oo, let Fn be a copy of the unit circle F , and let K be the disjoint union of the Fn 's, with the natural topology determined by declaring that Fn ---> Fee as n ---> oc. Each f E Uc determines F ~ C(K) by setting F = Tn f on In, 0 ~< n < oc, and F = f on Fee. Then Uc is isometric to a closed subspace of C(K).
Uniform algebras as Banach spaces
685
THEOREM 4.5. Let Uc be the Banach space of functions on F with uniformly convergent Fourier series, regarded as a closed subspace of C ( K ) as above. Then Uc is a tight subspace of C(K), though Uc is not compactly tight. Further, the weak-star closure H ~ (Uc, dO) of Uc in L ~ (dO), where dO is the arc length measure on F~, coincides with the space of functions f E L oc (d0) such that the symmetric partial sums of the Fourier series of f are uniformly bounded. The bidual U c is isometric to the direct sum of H~ dO) andan LeO-space. If g ~ C ( K ) is supported on one ofthe circles Fn for n finite, then the operator Sg is finite dimensional. Thus to check that Uc is tight, it suffices to show that the operators Sz and $5 are weakly compact, where z = e i~ on each circle Fn. Let f E Uc have Fourier series ak e ik~ and denote the corresponding function in C ( K ) by 45(f) = [T0f, T l f , . . . . f]. With this notation, ZCIO(f)
-- cI)(zf)
--
[aoe i0 --a-1 . . . . . ake i(k+l)0
--a-k-1
e -ikO
0]
Since this expression is in C ( K ) for all two-tailed g2-sequences {ak}, and since ~ [akl 2 < cc for f E Uc, the operator S~ factors through g2. Thus Sz is weakly compact, as is $5, and Uc is tight. To see that Sz is not compact, apply Sz to the sequence of exponential functions {eik~ There is another way to see that Sz is weakly compact. Since only two Fourier coefficients appear in each component above, we obtain II&fll ~
Iz~(f)- g'(zf)[]oc
~< 2[IfllL'(dO/2zr).
This estimate shows that Sz is an absolutely summing operator, in fact, an integral operator. Since bounded subsets of L~ are weakly compact in L 1, Sz is weakly compact. A similar theorem holds for the space UA of analytic functions on the unit disk with uniformly convergent Taylor series. Again we may regard UA as a subspace of C ( K ) as above, and UA is a tight subspace. In this case the weak-star closure H~ dO) coincides with the functions in f c H~ such that the partial sums of the power series of f are uniformly bounded. The bidual U~* is isometrically isomorphic to the direct sum of H~ dO) and an L~ The proof (see [47]) depends on a generalization of the E and M. Riesz theorem due to Oberlin [37], asserting that any measure on K orthogonal to UA is absolutely continuous with respect to dO on Foc.
5. The Pelczyfiski and Dunford-Pettis properties As Banach spaces, tight subspaces of C ( K ) share a number of properties of C(K). We discuss the Petczyfiski property, which is shared, and also the Dunford-Pettis property, which is partially shared. A Banach space X has the Petczyhski property if whenever T is an operator from X to another Banach space that is not weakly compact, there is an embedding co ~ X such that the restriction of T to co is an isomorphism. The spaces C ( K ) have the Petczyfiski
686
T. W. Gamelin and S. V. Kislyakov
property (see [51, III.D.w 33]). However, L 1-spaces do not have the Pdczyfiski property unless they are finite-dimensional. Reflexive Banach spaces have the Pdczyfiski property, by default. THEOREM 5.1 (Saccone). Any tight subspace of C ( K ) has the Petczyhski property. Before saying something about the proof, we discuss the Pdczyfiski property in more detail. A series ~ xk in X is weakly unconditionally convergent, or a wuc series, if ~ x* (xk) converges unconditionally for all x* 6 X*. In this case, ~ x*(xk) converges absolutely for each x* 6 X*, and the closed graph theorem shows that the operator x* w-> {x*(xk)} is continuous from X* to e 1. In particular, there is fl > 0 such that y~ [x*(xk)[ ~< ill[x*[[ for all x* E X*. The preadjoint operator T : co ~ X, defined on the standard basis vectors ek of co by T (ek) = xk, is then seen to be continuous and satisfy [[T [[ ~< ft. Conversely, any (continuous) operator T :co ---> X determines a wuc series ~ T (ek). Thus wuc series correspond to operators from co into X. Let E be a subset of X*. If E is weakly precompact in X*, and if y~ Xk is a wuc series in X with corresponding operator T, then T* (E) is weakly precompact in s 1. Consequently T* (E)(ek) tends to 0 as k ---> cx~, that is, sup Ix* (xk)] --> 0
as k ~ ~ .
(5.1)
x*EE
With a little more effort, it can be shown that if (5.1) fails, then there is an s sequence {x~} in E, that is, a sequence that is equivalent to an s These statements characterize weak compactness precisely when X has the Pdczyfiski property. We state this result formally. THEOREM 5.2. The following statements are equivalent, for a Banach space X. (i) X has the Petczyhski property. (ii) l f E is a subset of X* such thatfor any wuc series ~ x k in X we have x*(xk) --> 0 (as k --+ cx~) uniformly for x* E E, then E is weakly precompact. (iii) If E is a subset of X* that is not weakly precompact, then there is an s 1-basic sequence in E. We refer to [51 ] for the proof. A related result in this circle of ideas is that if X has the Petczyfiski property, then its dual space X* is weakly sequentially complete. Now we return to Saccone's theorem, which is proved in [46]. The crux of the matter is to find, for a given E that is not weakly compact, a wuc series in X for which (5.1) fails. To do this, Saccone begins with a characterization of weak compactness due to R.C. James, and eventually he throws the proof back on some difficult work of Bourgain [10,7], for which there is a clear treatment in [51, III.D. w167 29-32]. Recall that an operator T :X --+ Y is completely continuous if T maps weakly convergent sequences in X to norm convergent sequences in Y. If X is reflexive, then the completely continuous operators coincide with the compact operators, while every operator
Uniform algebras as Banach spaces
687
on X is weakly compact. In contrast, for X = C(K), the completely continuous operators coincide with the weakly compact operators. An isomorphism of co cannot be completely continuous, as the standard basis of co converges weakly to 0. Thus if X has the Petczyfiski property, then any completely continuous operator from X to another Banach space is weakly compact. As a corollary to Saccone's theorem, we then obtain the following corollary. COROLLARY 5.3. If B is a tight subspace of C(K), then any completely continuous operator T : B -+ Y is weakly compact. A Banach space X has the Dunford-Pettis property if every weakly compact operator from X to another Banach space is completely continuous. This occurs if and only if whenever the sequence {xn } in X converges weakly to 0, and the sequence {x* } in X* converges ak weakly to 0, then xn (x,) ~ 0. The spaces C(K), and any L 1-space, have the DunfordPettis property. If a dual Banach space has the Dunford-Pettis property, then its predual does also. Reflexive Banach spaces do not have the Dunford-Pettis property unless they are finite-dimensional. See Basic Concepts. Not every tight subspace of C (K) has the Dunford-Pettis property. In fact, any infinitedimensional reflexive subspace of C(K) is tight but fails to have the Dunford-Pettis property. However, the Dunford-Pettis property does hold under hypotheses that are somewhat stronger than tightness. The following statement can be extracted from Bourgain's work in [10]. THEOREM 5.4 (Bourgain). Let A be a subspace of C(K). If Sg* is completely continuous for all g E C, then A* and A have the Dunford-Pettis property. The collection of g E C such that Sg* is completely continuous is a closed subalgebra of C. It is called the Bourgain algebra associated with A. We wish to develop some criteria that guarantee that Sg* is completely continuous. The following condition is a variant of the notion of a rich subspace, which stems from [51 ]. An operator T from A to another Banach space is nearly dominated if there is a probability measure # on K such that if {fro } is a bounded (!) sequence in A that converges to 0 in L1 (#), then I[T fro [[ ~ 0. Trivially, absolutely summing operators are nearly dominated. If Tj : A --+ X is nearly dominated by # j , and if Tj --+ T in operator norm, then T is nearly dominated by ~ ~j/2 j . It is straightforward to show that the collection of g E C such that Sg is nearly dominated forms a closed subalgebra of C. The pointwise bounded convergence theorem can be used to show that nearly dominated operators are completely continuous. THEOREM 5.5. Let A be a subspace of C(K). Each of the following conditions guarantees that Sg* is completely continuous for all g E C, hence that A* and A have the Dunfo rd-Pettis property. (i) The subspace A is compactly tight. (ii) The operators Sg are absolutely summing for a family of functions g E C that generates C as a uniform algebra.
688
T. W. Gamelin and S. V. Kislyakov
(iii) The operators Sg are nearly dominated for a family of functions g E C that generates C as a uniform algebra. For (i), observe that if Sg is compact, then Sg* is compact hence completely continuous. For (ii), we use the fact that if Sg is absolutely summing, then Sg* is absolutely summing hence completely continuous. The proof under the condition (iii) is straightforward. Note that either of the conditions (ii) or (iii) covers the subspaces UA and Uc discussed earlier. Each of the three conditions covers the algebras R ( K ) and A ( K ) from rational approximation theory.
6. Absolutely summing and related operators on the disk algebra Now we consider some properties of the Banach space C (F) of continuous functions on the unit circle F that are inherited by the disk algebra CA. A prototypical theorem along these lines is the following (see [39]): THEOREM 6.1 (Mityagin-Pelczyfiski). For 1 < p < cx~, every p-summing operator from the disk algebra CA to a Banach space Y extends to a p-summing operator from C ( F ) to Y, hence is strictly p-integral. The Paley operator shows that the statement fails at the endpoint p = 1. The proof of the theorem depends on the boundedness of the Riesz projection R from L p (dO) onto H p (dO), 1 < p < cxz, together with some Hardy space theory. Indeed, let T be a p-summing operator, and let # be the measure on F for T given by the Pietsch theorem (see Basic Concepts), so that T extends to a continuous operator from the closure H p (#) of CA in L P ( # ) to Y. Let # = w dO + #s be the Lebesgue decomposition o f / z with respect to Lebesgue measure dO. The Hardy space theory gives H P ( # ) = H P ( w d O ) G LP(lZs). Further, if logw r Ll(d0), then H P ( # ) :- LP(#), and T is p-integral. On the other hand, if log w 6 L1 (d0), and h = exp(log w + iT-/(log w)) is the "outer" function in H 1(dO) such that Ihl = w, then Q p ( f ) = h - 1 / p R ( h l / p f ) projects LP(wdO) onto HP(wdO), and this projection allows us to factor T, T'CA ~
L~(#) ~
L p(#)
QpGI>H p ( # )
> y,
again showing that T is strictly p-integral. Our primary focus will be on Bourgain's extension of the Grothendieck theorem to the disk algebra, with emphasis on the interpolatory nature of the proofs. We will sketch the proofs modulo the interpolation theorems, which we defer to the next section. The Grothendieck theorem (see Basic Concepts) asserts that any operator from an L 1-space to E2 is absolutely summing. A dual version of Grothendieck's theorem asserts that any operator from a C(K)-space to E 1 is 2-summing. In fact, we can replace E 1 in this statement by any space of cotype 2. In reading the following version of Grothendieck's theorem, recall that among the spaces LP, precisely those with 1 ~< p ~< 2 are of cotype 2.
Uniform algebras as Banach spaces
689
THEOREM 6.2. Every operator from C ( K ) to a space of cotype 2 is 2-summing. Theorem 6.2 is a simple consequence of the following two lemmas. LEMMA 6.3. If Y is of cotype 2, and if p ~ 2, then any p-summing operator from a Banach space X to Y is 2-summing. LEMMA 6.4. For every Banach space Y and everyfinite rank operator T : C ( K ) --+ Y we have 7rp(T) <~ rc2(T)~176 f o r 2 < p < ~ , where O = 2/p. If the lemmas are proved and Y is of cotype 2, we combine the lemmas to obtain 7t'z(T) ~< czr4(T) ~< c(7rz(T)llTII) 1/2 for every T : C ( K ) --+ Y of finite rank, whence 7rz(T) ~< c211TII. The finite rank assumption is easily lifted by approximation. The first lemma is proved by an easy concatenation of inequalities, one of which is the Khinchine inequality (see [51, III.F w 36], and also Basic Concepts). It yields the estimate 7t'z(T) ~< cpCq(Y)Tcp(T), where Cq(Y) is the cotype constant of Y, and Cp depends only on p. To prove the second lemma, we choose by the Pietsch theorem a probability measure # on K such that the operator T acts from L2(#) to Y with norm 7r2(T). Also, T acts from C ( K ) to Y with the norm IITII, and consequently T extends to L ~ ( # ) with (at most) the same norm, T'L2(#)
Jrz(T)> y,
(6.1)
T" L ~ ( # )
IITII>y.
(6.2)
By interpolation, T acts from L p (#) to Y with norm not exceeding 7r2(T) 0 IlZll 1-0, where 0 is given by the convexity condition 1/p = 0(1/2) + (1 - 0)(1/cx~) = 0/2. This proves the lemma and with it Grothendieck's theorem. Now we turn to Bourgain's version of the theorem for the disk algebra. THEOREM 6.5 (Bourgain). Every operator from the disk algebra CA tO a space of cotype 2 is 2-summing. As previously, the proof is an easy consequence of Lemma 6.3 and the following analog of Lemma 6.4. LEMMA 6.6. For every Banach space Y and every finite rank operator T : CA --+ Y we have top(T) <~ crcz(T)~ 1-~ f o r 2 < p < cx~, where 0 -- 2 / p and c is a universal constant. For the proof, we start as in the proof of Lemma 6.4 with a probability measure/z on the unit circle F such that T acts from the closure H2(/z) of CA in L2(#) to Y with norm zr2(T). Our first problem is that the measure # need not be absolutely continuous with respect to arc length dO. For this, we invoke the following:
T.W. Gamelin and S. V. Kislyakov
690
ABSOLUTE CONTINUITY PRINCIPLE. In problems like this, the singular parts of measures can be disregarded. One way to justify this is to refer to the decomposition HP(#) = HP(wdO) G LP(#s) used above and to the Hardy space theory underlying this decomposition. However, sometimes other arguments are also applicable. In the case under consideration, we may work with the operators T n f = T(Kn * f ) in place of T, where Kn is the nth Fej6r kernel. For them the above measure becomes absolutely continuous, and moreover the Tn's may be regarded directly as operators on HX(d0). Thus we assume that # = w dO, where w ~> 0 is a weight, f w dO = 1. We arrive at the following analogs of (6.1) and (6.2):
T" H2(wdO) 7r2(T)> y,
(6.3)
T.H~(wdO)
(6.4)
IITII>y.
The question now is whether we can interpolate between (6.3) and (6.4) as we did between (6.1) and (6.2). The answer is that we can replace w by a weight v ~> w, f v ~< C, such that for this new weight the above interpolation is possible. This will follow from results in the next section. Indeed, since L 1(d0) is BMO-regular (see Proposition 7.4), there is a majorant v for w such that log v belongs to BMO, and on account of Theorem 7.7 the desired interpolation holds for this majorant v. This proves Bourgain's theorem. In a standard way, Bourgain's theorem implies that every operator from C~ (or from L 1/ H~) to 12 is absolutely summing. Then from the relations CA ~ (CA ~ CA ~ ' " ")co,
L'/H
IL'/H 9 L /H0' 9 ) , ,
(see [51, III.E. w 12]), it is also standard to conclude that L 1 / H I and C~ are of cotype 2. See [51, III.I, w 14] for more details. We mention another approach to the Grothendieck theorem, due to Maurey. This method gives some information about operators T : C ( K ) ---> Y, where Y is a space of arbitrary finite cotype. In particular, it applies to Y -- LP for any 1 ~< p < c~. For the proof of the following theorem, see [15, Chapter 10] or [30]. THEOREM 6.7 (Maurey). For 1 <<,p < q, the class of (q, p)-summing operators defined on C ( K ) does not depend on p and is contained in the class of (q + e)-summing operators for any e > 0 . A related result, due to Pisier (see the above references), is that T : C ( K ) ---> Y is (q, p)summing if and only if T factors through the inclusion C ( K ) ~ Lq,1 (#) (the Lorentz space) for some probability measure # on K. If Y is of cotype q, the identity operator of Y is (q, 1)-summing. Thus Maurey's theorem shows that every operator from C ( K ) to Y is (q + e)-summing. We recover the Grothendieck theorem by setting q = 2 and applying Lemma 6.3. The following theorem allows us to transport these results to the disk algebra CA.
Uniform algebras as Banach spaces
691
THEOREM 6.8 (Kislyakov). For an arbitrary Banach space Y and q > p ~ 1, every
(q, p)-summing operator T : CA ---> Y extends to a (q, p)-summing operator from C(F) toY. According to the Mityagin-Petczyfiski theorem, the theorem remains true if q = p > 1. The remainder of this section is devoted to an outline of a proof of Theorem 6.8, with some simplifications compared to the exposition in [30]. LEMMA 6.9. Under the conditions of Theorem 6.8, the operator T is lOq-summing. We break the proof of L e m m a 6.9 into four steps. First note that the family of (q, p)summing operators grows as p decreases, so we may assume that p = 1. We assume also that 7rq,1 (T) = 1. When convenient, we use (., -) to denote the pairing between vectors and functionals. Step 1. There is a probability measure )~ on F such that
IlZxll q ~ < q [ ( 1 - ~ , ~ > l
forallx, rpECA satisfying I x l + l ~ l ~ 1.
(6.5)
To see this, we use a trick invented by Pisier to prove his characterization of (q, p)summing operators on C(K) mentioned above. Let
9 X l .....
XnE CA, Z I xj(t) l ~ 1}.
Clearly C n / z zrq,1 (T) -- 1. We choose 6~ "x 1 and for every n find xl n) . . . . . x(~n) E CA .,. (n) . (n) such that ~ j Illxj IIq > 1 and y~'j I~j (t)l ~< 6,/Cn. Then we c h o o s e / ~: (j ' ) E Y* such that ~ j
I1~: ~.!n)) _ 1 , and define a functional )~n on CA by the ~ j !') IIq' ~< 1 and ~ j (TxJ n) ~ ~j
formula )~n ( ~ ) -- ~j(T(Trx~. ")), "J~:!'))" Then )~, (1) -- 1 and IIZ, II ~< ~,. We consider a weak-star limit point of the sequence {)~}. This is a functional on CA. We extend it to C(F) with preservation of norm, obtaining a measure )~ on F . Since )~(1) = 1 and I1~11 ~< lim 6n -- 1, )~ is a probability measure 9 Now let x, 99 E CA satisfy Ixl + I~ol ~< 1. We define Yl . . . . . Yn+I 1 ~< j ~< n and Y,+I = x . Then
( Z l l T y j l l q ) 1/q ~< Cn+, s u p ~ t
E CA
by yj
--
qgx)n) for
ly/(t)] <~c.+,a./c.,
which implies that I(go,)~n)l q + IlZxll q ~ ( C n § q and, in the limit, I(go,~)l q + IlTxll q ~< 1. Finally I(~p,)~)lq = I1 - (1 -qg,)~)l q ~> 1 - q l ( 1 -qg,)~)l, from which (6.5) follows. Step 2. We apply the absolute continuity principle. It can be assumed that )~ = v dO, f v dO -- 1, and that (6.5) is valid for x, rp E H a (dO). (Again, we can convolve with Fej6r kernels to ensure this.)
T. W. Gamelin and S. V. Kislyakov
692
Step 3. There is a ~> v,
fa
~< C, such that
1-1/8q 1/8q IITxll ~
x 9
HOC ,
ii1/8 ). To see this, we assume that Ilxll~ ~< 1/2. It suffices to show that IITxll q <~ Cllx,,g~(a We shall deduce this from (6.5) by a careful choice of qg. Denote by 7-[ the harmonic conjugation operator. As in the proof of Proposition 7.4, there exists a ~> v, f a ~< Co, such that for b = a 1/2 we have lT-/(b)[ ~< Cob, and Ib + iT-/(b)l ~< (1 + Co)b. We put ot = log(1 - Ixl), so that ot ~> 0, and also ot ~< Cllxl since Ixl ~< 1/2. We define successively -
~p --
c~4b -+- i7-/( ~
)
,
4~ -- ~p
1/4
,
q) -- e x p ( - A 4 5 ) ,
b + iT-/(b) where the constant A > 0 will be chosen momentarily. Since 7t is the quotient of functions with values in the right half-plane, it omits the negative axis, and we choose the branch of 45 whose argument ranges between-+-zr/4. Then Req~ ~> 1~1/~/2 and l 1 - qgl ~< C21r Now 17~1 >/or4/( 1 -+- Co), so R e r ~> C3ot. We set A = 1/C3, and then I~01= e x p ( - A R e q ~ ) ~< e x p ( - A C 3 o t ) = 1 - Ixl. Thus we may apply (6.5) with this q9 and )~ = v dO:
IIrxllq f II- t dO
f I*ladO C4 I*tadO
Now, 1~lSa : 1~12b 2 ~ (ot4b) 2 + 7"-/(ot4b)2. Using the L2-estimate for 7-/, we obtain
f
IqblSad0 ~<
f((=4b)2
+7-/(ot4b)2) d0 ~<
2f(4b)
dO
<. f lxl adO
IITxll ~ C'llxllL~q ,
x E CA.
Thus T is 10q-summing, and L e m m a 6.9 is established. Next we need to introduce "vector coefficients". For a Banach space X, we consider the space CA (X) of all X-valued continuous functions f on the unit circle F that extend
Uniform algebras as Banach spaces
693
analytically to the unit disk, that is, that satisfy f f ( z ) z n dO - - 0 for n ~> 1. An operator T ' C A (X) ~ Y is said to be (q, p, X)-summing if
( Z IlZxjllq)1/q~ Csup(~ tEF
Ilxj(t)ll~)1/p
(6.6)
for any finite collection {Xj} in CA (X). LEMMA 6.10. If T" CA (X) --+ Y is (q, p, X)-summing, then there is a probability mea-
sure # on F such that
IITxll ~< C(f IIx(t)llxlOqd#(t))l/lO q
Indeed, it is routine to carry through the above proof of L e m m a 6.9. For this, note that some functions will remain scalar-valued, as for instance the function q) in (6.5). The condition on x and q) in (6.5) becomes ]lx(t)llx + Iqg(t)] ~< 1 for t E F . For any operator T ' C A ~ Y, we define the operator T" {xj} ~ {Txj} on sequences of functions in CA. Evidently T is (q, p)-summing if and only if T maps CA (gP) to ~q (Y). It is quite easy to see that even more is true. LEMMA 6.1 1. If T is (q, p)-summing, then T is (q, p, g~P)-summing. Now we apply L e m m a 6.10 to T and proceed as in the proof of the Mityagin-Petczyfiski theorem. As before, we may assume that the measure # is absolutely continuous, and even that # - a dO with log a E L 1(d0). Then T has the factorization
T " CA (g~P) ~
H |Oq (g~P,a) ~ e q (Y),
(6.7)
where the first mapping is the identity embedding and the second is the extension of T by continuity. To complete the proof, we need a projection. LEMMA 6.12. If 1 < p, s < ec and log a E L l(dO), then there is a projection Q from
LS(g~P,a) onto HS(g~P,a) having the form Q ( { f j}) - { Q f j} for a projection operator Q acting on scalar-valued functions. We take Q to be the projection Q,2 in the proof of the Mityagin-Petczyfiski theorem (Theorem 6.1). The boundedness of Q follows from standard techniques. It is now easy to establish Theorem 6.8 in the case p > 1. If p > 1 in (6.7), then T extends to some operator U ' C ( g p) ~ ~.q (Y) of the form U{xj} = {Sxj}, where S acts from C ( F ) to Y. The boundedness of U means that S is (q, p)-summing. Clearly S extends T. It remains to treat the case where p -- 1. The facts already proved and Maurey's Theorem 6.7 show that for 1 < r < q the class of (q, r)-summing operators from CA to Y does not depend on r. It suffices to extend this statement to r - 1. For this, let T ' C A ~ Y be of finite rank, and let 1 < r < s < q. Then we have
T " CA (~s)
7rq's(T))eq (y),
T'CA(s l) 7rq'l(T))eq(Y).
(6.8)
694
T.W. Gamelin and S. V. Kislyakov
By the remark after the proof of L e m m a 7.6 (where H~ are involved, but this does not matter too much), we can interpolate as if we had C (U) and c ( e l ) . This shows that 7rq,r(T) ~ C~q,s(T)l-OT~q,1(T) ~ for some 0 < 0 < 1. Since the n o r m s ~q,r and ~q,s a r e equivalent, we obtain the desired result.
7. Interpolation of Hardy-type subspaces Several times in Section 6 we had to interpolate either between weighted Hardy spaces HP(adO), or between Hardy spaces of vector-valued functions Hp(er). To cover both cases, we consider the measure space ( F x I2, m x #), where d m = d0/27r is normalized arc-length measure on the unit circle F , and (12, #) is some fixed o--finite measure space. Since we wish to use the full range 0 < p ~< + c ~ , we will refer to quasi-Banach spaces where appropriate. A lattice of measurable functions on ( F x $2, m x / z ) is a quasi-Banach space X of measurable functions such that if f E X, g is measurable, and Igl ~< I f I, then g E X and Ilgl]x ~< Cllfllx. (Note that this is not the same as a Banach lattice, as defined in Basic Concepts, whose elements are measurable functions. Since we treat the term as an inseparable unit, lattice-of-measurable-functions, there should be no confusion.) The examples we have in mind are the spaces L p (w dm d/z), and the spaces L p (dm, L r (/z)) of measurable functions x(t, co) such that y(t) - ( f Ix(t, co)lr d/z(co)) 1/r is in LP(dm). Let N + be the Smirnov class of analytic functions on the unit disk (see [17,41 ]), which we identify with their boundary value functions on the circle. (For our purposes the class N + could be replaced by Up>0 HP.) We call a function on the circle analytic if it belongs to N +. If X is a lattice of measurable functions on ( F x S2, m x / z ) , we define its analytic subspace XA t o be the set of functions f E X such that f (., co) 6 N + for almost all co. In the case of functions of one variable, as w h e n / z is a point mass, we have LPA -- H p. In the sequel we also impose on X the following conditions: (i) if f E X, then f r l~ + If(t, co)l dm(t) < ec a.e. on S2, (ii) if fn --+ 0 in X, then f r l~ + Ifn(t, co)l dm(t) --+ 0 in/z-measure, (iii) if f E X, there exists g 6 X such that Ifl ~< Igl, Ilgllx ~< CIIfllx, and log Ig(', co)l L 1(m) for a.a. co. These conditions serve to exclude various degenerate possibilities. Under these conditions it is easy to prove, for instance, that XA is closed in X. Now let X and Y be lattices of measurable functions. The fundamental problem we consider is to determine when interpolation properties of the couple (X, Y) are inherited by the couple (XA, YA). We shall deal with real interpolation only. We remind the reader of the definition of the real interpolation spaces (X0, X1)O,q for a couple (X0, X1) of compatible quasi-Banach spaces. By compatible we mean that X0 and X1 are linear subspaces of some ambient space, so we may define the K-functional K (x, t; X0, X1) for t > 0 and x E X0 + X1 by
K ( x , t ; Xo, X1) : inf{llxollo + tllxllll" xo-+-xl - - x , xo e Xo, x l
EX1}.
For 0 < 0 < 1 and 0 < q ~< cx~, we define the interpolation space (X0, X1)O,q t o consist of x 6 X0 + X1 such that t o K(x, t; Xo, X1) belongs to L q (dt/t), and we define the norm
Uniform algebras as Banach spaces
695
of x in (X0, X1)O,q to be the norm of t o K ( x , t; X0, X1) in L q (dt/t). Actually the specific expressions for the K-functional and the norm will not play a role for us. Let Y0 C X0 and Y1 C X1 be closed subspaces. We say that the couple (Y0, Y1) is Kclosed in (Xo, X 1) if there is C > 0 such that any decomposition y = x0 + x 1 of an element y E Y0 + Y1 with xi E Xi can be modified to a decomposition y = Y0 + Yl with Yi E Yi and [[Yi 11/ ~ C]]xi 11i, i = 0, 1. In this case we have
(Yo, Y1)o,q = (Yo -+- Y1) A (Xo, Xl)o,q, with equivalence of norms, and the interpolation properties of the couple (X0, X1) and its subcouple (Y0, Y1) are identical. Our basic problem can be formulated as follows. PROBLEM. When is the couple
(XA, YA) K-closed in (X, Y)?
We shall see that this happens fairly often. We begin with a useful duality result. Assume that X0 and X1 are Banach spaces and that X0 A X1 is dense in both X0 and X1. Then, in a natural way, the spaces X 0 and X 1 are included in (X0 A X1)* and, consequently, form a compatible couple. If Yi C Xi (as above), we denote by Yi• the annihilator of Yi in X*, that is, the set of L 6 Xi* such that
L = O o n Yi. LEMMA 7.1. The couple (Yo, Y1) is K-closed in (Xo, X1) if and only if the couple
(Y~-, Y(-) is K-closed in (X~, X~). The proof is left to the reader (see [31,38]). If X is a Banach lattice of measurable functions on ( F • 12, m • #), it often happens that under the duality (f, g) = f f f g dm d#, X* is also a lattice of measurable functions on the same measure space. We will assume that this is the case, and further that both X and X* satisfy the conditions (i)-(iii) above. Then one easily sees that, as in the classical case of the HP-spaces on the circle, we have X ] -- Z(X*)A, where Z is the coordinate function on F . Thus L e m m a 7.1 relates interpolation properties of the couples (XA, YA) and ((X*)A, (Y*)A), X and Y being two lattices as above. The class of BMO functions will play an important role in what follows. Recall that a function f on F is in BMO if f = u + 7-/v, where u, v 6 L ~ . As usual, we disregard the constant functions and define IIf IlBgO to be the infimum of Ilu II + Ilvll over all such representations, where II 9II is the norm in L ~ modulo the constants. LEMMA 7.2. Let w > 0 be a measurable function on F. Then log w ~ BMO if and only if there exist constants C > 1, 0 < p < 1, and a function f > 0 such that w~ C <<,f <~ C w and lT-/(fP)l ~< C f p. Moreover, C and p are controlled in terms of II log WIIBMO, and vice versa. PROOF. Suppose log w ~ BMO. Then log w = ot + u + ~ v , where ot is a real constant, u and v are real functions, f v = 0, and Ilull~ + Ilvll~ ~< 211 logwllBMo. We pick p so small that pllpll~ < ~r/4, and we set F = e x p ( - i p ( v + iT-/v)) = ePT-tv[cos(pv) + i sin(pv)] and
T.W. Gamelin and S. V. Kislyakov
696
f = e~(Re F ) 1/p = we-U[cos(pv)] 1/p. Since F is analytic in the unit disk and F(0) is real, we have I m F = 7-/(Re F). Since I sinpvl ~< cos pv, we have I I m F I ~< Re F, and consequently lT-/(fP)l ~< f P . The estimates 1 ~> cos(pv) ~> 1/~/2 lead to w / C <, f <, Cw. Conversely, given f as in the lemma, we put G = f P + iT-/(fP). Since 17-/(fP)l ~< C f p, the values of G lie in a sector in the right half-plane, and the principal branch of log G is analytic. Writing log G = log IGI + i arg G, we have l arg GI ~< tan -1 C and 7-/(arg G) = - l o g l G I + loglG(0)l, so loglGI E BMO. Since f P ~< IGI ~< ~/1 + C Z f p, we see that p log f - log IG] is bounded, and log f E BMO. Finally, log f - log w is bounded, so log w E BMO, with BMO-norm bounded in terms of C and p. [2 From the proof we see that we can always reduce p, at the expense of increasing C and changing f . A weight is a function w > 0 on F • 1-2 such that log w(., co) E L l(dm) for a.a. co. For 0 < p < cx~ we denote by LP(w) the usual space L P ( w d m d/z), with norm denoted by Ilfllp,w, though we shall denote by L ~ ( w ) the space of functions f on F z 12 such that f / w is bounded, with the norm I l f l l ~ , ~ - e s s s u p { I f ( f f , co)l/w(ff, co)" ((,09) ~ F • s2}. With this notation, L ~ (w) = L 1(w)* under the non-weighted duality (f, g) = f f f g dm d/z. We denote L p (w) by H p (w), 0 < p <~cx). For a weight w, we say that log w is uniformly (or C-uniformly) in BMO if the function log w (., co) is in BMO for almost all co, with BMO-norm bounded by C. In this case, the analog of L e m m a 7.2 holds, where the function f can be chosen to depend measurably on the parameter co. We will use this extended version of Lemma 7.2. A quasi-Banach lattice of measurable functions X on F • g2 is said to be BMO-regular if for every x ~ X, there exists u E X such that Ix[ ~< u, Ilullx ~< CIIxllx, and logu is Cuniformly in BMO, where C depends only on X. The function u will be referred to as a BMO-majorant of x. As an easy consequence of (the extended version of) Lemma 7.2, we have the following. LEMMA 7.3. A lattice X is BMO-regular if and only if there are C, p > 0 such that for every x E X there exists u E X with Ixl ~< u, I[u[I ~< Cllxll, and lT-/(uP(., co))l ~< CuP(.,co) for a.a. co E I2. At the expense of increasing C, we can take p to be arbitrarily small. For a quasi-Banach lattice of measurable functions X and 0 ~ < cx~, we define X ~ to be the space of functions f such that Ifl ~ E X, with quasi-norm 1/~
llltl - [llll ii .
Thus if X = L p, then X/~ = L p~ . Clearly X/~ is BMO-regular if and only if X is. Our main examples of BMO-regular spaces will be based on the following proposition. PROPOSITION 7.4. If the operator 7-[ (acting in the first variable) is bounded on X ~, then X is BMO-regular.
Uniform algebras as Banach spaces
697
PROOF. It suffices to show that Y -- X ~ is BMO-regular. We verify the conditions formulated in L e m m a 7.3. Taking y E Y, we put Y0 = lyl, Yn+l --lT-/(Yn)l for n ~> 0, and v = ~ 6nyn, where 6 > 0 is a fixed small constant. Then lYl ~< v, Ilvllr" ~< CllYlly, and I~vl ~< v/S. D LEMMA 7.5. / f l o g w is uniformly in BMO, then L p (w) is BMO-regularfor 0 < p <~ cxz. PROOF. If 1 < p < ~ and w = 1, we may apply Proposition 7.4, with/~ = 1. The case 0 < p < cx~ and w arbitrary then follows easily from the definitions. For p = cx~, the definition of the n o r m in L ~ ( w ) as sup(Ixl/w) shows that Ilxll~w is a BMO-majorant of x E L c~ (w). [] LEMMA 7.6. The space L p (dm, L r (S2)) is BMO-regularfor 0 < p < cx~, 0 < r <~cx~. PROOF. The case where r < cx~ is a consequence of Proposition 7.4, and the case where r = cx~ of its proof. Indeed, given x, we construct a BMO-majorant for y = ess sup~o Ix (., o9) 1 in L P (dm), and then treat this majorant as a function of two variables. It is easy to find other examples of BMO-regular spaces on the basis of the same ideas. A less trivial example is the space L ~ ( d m , U ) , 0 < s < ~ (see [31]). While we could have cited this example (and, of course, T h e o r e m 7.7) when interpolating in (6.8), that proof can be based also on the duality L ~ ( d m ; s = L l(dm; gs'), for s ~> 1, where the latter space is BMO-regular by L e m m a 7.6. Thus, to interpolate in (6.8) we can refer to Corollary 7.8. E] Now we state our main interpolation result. We are assuming that X and Y are Banach lattices of measurable functions o n / - ' x S-2, and (when applicable) that X* and Y* are also lattices of measurable functions on/-" x S-2, all satisfying the conditions (i)-(iii) above. Also, the density of X n Y in X and Y is assumed when needed. THEOREM 7.7. If X and Y are BMO-regular, then
(XA, YA) is K-closed in (X, Y).
There is some evidence (see [27]) in favor of the conjecture that BMO-regularity is a self-dual property. However, this has not yet been verified in the general case. Thus, we combine L e m m a 7.1 with T h e o r e m 7.7 to obtain more information. COROLLARY 7.8. In any of the following three cases, (XA, YA) is K-closed in (X, Y), and ((X*)A, (Y*)A) is K-closed in (X*, Y*). (a) X and Y are BMO-regular, (b) X* and Y* are BMO-regular, (c) X and Y* are BMO-regular. Cases (a) and (b) of the corollary are direct consequences of T h e o r e m 7.7 and L e m m a 7.1. Case (c) is not needed in Section 6, so we leave it as an exercise. We pass to the proof of T h e o r e m 7.7. Let w0, wl be two weights whose logarithms are C - u n i f o r m l y in BMO.
T.W. Gamelin and S. V. Kislyakov
698
If f (., co) E N + for a.a. co and f = g + h with g E L~ h E LOC(Wl), then f = q)+ ~p with q) E H~ lp E H~ and I1~011~,~0 ~< C'llgll~,~o, II~rlloc,Wl ~< C'llhll~,~l, where C' is determined by C. LEMMA 7.9.
Clearly Theorem 7.7 is a consequence of this lemma. Given f = x + y with
f E XA -[- YA, X E X, y E Y, we find BMO-majorants for x and y in their respective spaces, and apply the l e m m a to these majorants as the weights. To prove L e m m a 7.9, we first assume that f E H~(wo) + H ~ ( w l ) . Then the statement to be proved is precisely the K-closedness of the couple (H~ H~ in (L~(wo),L~ By duality ( L e m m a 7.1), it suffices to check that the couple ( H I ( w o ) , H I ( w l ) ) is K-closed in (Ll(wo),Ll(wl)). So let z = a + b E Hl(wo) + H 1(wl), where a E L 1 (W0), b E L 1 (t/)l). We must replace a and b with functions roughly of the same size but analytic in the first variable. Let v be a BMO-majorant for b in the BMO-regular space L 1(Wl). Setting w -- wov, we apply L e m m a 7.2 and the remark after its proof to find a function k and a constant p < 1 such that w/C <. k <~Cw and [7--/kP(., co)l ~< CkP(.,co). Fixing an integer n > 1/p, we define oe - max{ 1,
kP + iT-/(kP) F = kPc~ + iT-/(kPot) '
(lal/v)l/n},
G -- 1 - (1 -
F') n.
We claim that z = (1 - G)z + Gz is the required decomposition. Indeed, the summands are analytic in the first variable, and it suffices to estimate the norms. Since 17-/(kP)l ~< CkP, we see that IFI ~< C1/~ <<.C1 and IGI ~< C2/~ n <<.C2, whence IGzl <<.C3lalot -n + C2lbl ~< C3v + C2lbl. By the choice of v, we obtain the required inequality IIGzlll,~l ~< C4llblll,wl. We estimate the quantity II(1 - G) z Ill, ~o ~< (1 + C2)Ila IIl, ~o + II(1 - G)b Ill, ~o- We have I1 - FI =
I(oe-
I~kp + iT-/(oekP) I
Since also IFI ~< C1, Ibl ~< v, and
t
~< (oe-
[1 - G[ [b[wo ~< C5 1 [1 /
1)+
17-/((oe- 1)kP)[
kP
w <~Ck, we see that /.
/.
/.
]
iT-/((ot- 1)kP)l
1)k p +
Finw <<.C 6 [ [1 - F]I/Pw
l t J
<<. c 7 ( f ( ~ _ l ) l / P w + f lT_t((~_l)kP)ll/P-kW) <~ c 8 ( f ( o t - 1 ) l / p w +
f 7-[((ot-1)kP)ll/P).
Since 7-/acts on L lip (dm d # ) and k <~Cw, it follows that the second integral in parentheses is dominated by the first. Now ot - 1 = 0 if loll ~< v and ot - 1 ~< (lal/v) 1In <<.(lal/v)P otherwise. Therefore
f (ot and we are done.
1)l/pw<.f (lal/v)w- f lalwo,
Uniform algebras as Banach spaces
699
It remains to get rid of the assumption f E H~C(wo) + H ~ ( w l ) . This is done by a standard approximation argument based on the Hardy space theory. Suppose only f E N +. For u = log Ifl define G(., co) = exp(u(., co) + i ~ u ( . , co)) and F = f / G , so that IFI = 1 a.e. Thus f = F G is the "inner-outer" factorization of f . Set uj = I f ] / x (jwo) and Gj(., co) = exp(uj(., co) + i ~ u j ( . , co)). Then IGjl <, IGI, and Gj --> G in measure. Set f j = F G j E H~ then f j --> f in measure, and Ifjl <~ ]fl ~< Igl + Ihl. By the first part of the proof, f j = ~oj + grj, where ~pj E H ~ ( w o ) , grj E H~ I]~ojIIoc,w0 <~ C'llgll~o,wo,and 1lTsjIloo,wl ~< C'llh II~o,Wl. For some subnet ofthe integers we have ~oj -+ ~o and 7tj --> ~ weak-star. Simultaneous convex combinations of the ~oj's and 7sj's can be chosen to converge a.e. to ~o and ~, and this guarantees that f : ~o+ ~ with the appropriate estimates for I~01 and I~l.
8. Bourgain projections Let w be a weight on the unit circle. We say that w admits an analytic projection if there is an operator Q that projects L p (w) onto H p (w) for all 1 < p < cx~ at once and, together with Q*, is of weak type (1, 1) relative to w,
{iQst >x}
7
cf tsl ,
cf
f E L 1(to),
where we denote w ( e ) - - fe w. Here the adjoint Q* is calculated relative to the duality (f, g) = f f ~ w dm. Operators of this sort served as the main technical tool in Bourgain's work extending Grothendieck's theorem to the disk algebra. Bourgain [9] proved that for every integrable weight u there exists a weight w admitting an analytic projection and satisfying w ~> u and f w ~< C f u, where C is a universal constant. The existence of such a projection does not imply the K-closedness in the scale H p (w) if the "extreme" exponents p -- 1 and p = cx~ are involved. However, it still implies certain "nice" interpolation properties of this scale, sufficient for instance for proving Lemma 6.4. Though here we have used different (simpler) techniques, Bourgain's projections remain interesting in themselves. We shall show that a weight w admits an analytic projection if and only if log w E BMO. Bourgain's majorization result quoted in the preceding paragraph then follows from the fact that L 1(d0) is BMO-regular. We need a technical notion. A two-tailed sequence of functions @j E H ~ - o c < j < e~, is called an analytic decomposition of unity subordinate to a weight w if there exists a constant c such that (i) Iq)jll/Sto ~ c2 j, --oc < j < oc, (ii) ~ I~Pjl 1/s2j <~cw, (iii) y~ Iqgjl 1/8 ~< c, (iv) ~ ~0j = 1. Roughly speaking, the functions ~j behave like the characteristic functions of the sets where 2 j-1 ~< w < 2 j . The exponent 1/8 is convenient technically but, in principle, may be replaced by any a E (0, 1]; see [28].
T. W. Gamelin and S. V. Kislyakov
700
THEOREM 8.1. For a weight w, the following conditions are equivalent: (1) log w E BMO, (2) there is an analytic decomposition o f unity subordinate to w, (3) w admits an analytic projection, (4) there is an operator Q projecting L p (w) onto H p (w) f o r two different values o f p. We focus on the implications (1) =~ (2) =~ (3). The implication (3) =~ (4) is trivial, and (4) =:~ (1) is proved in [32, Corollary 2.2]. To prove that (1) =~ (2), we need a lemma. LEMMA 8.2. Suppose u is a weight such that log u E BMO. Then log(1 A u) E BMO, and the BMO-norm of log(1 A u) is controlled by the BMO-norm of log u. PROOF. By Lemma 6.2, there exist C > 1, 0 < p < l, and f such that u / C <. f <. Cu and 17-/(fP)l ~< C f p. We put g - (1 4- f p ) l / p , then ]7-/(gP)l ~< Cg p and (u 4- 1)/C1 ~< g ~< C1 (u 4- 1). Thus the B M O - n o r m of log(u 4- 1) is controlled by the BMO-norm of log u, and consequently we have similar control over the BMO-norm of log(u/(u 4- 1)). Finally note that (1 A u ) / 2 <. u / ( u + 1) ~< 1 A u. Now to show (1) =~ (2), let log w E BMO. For any X > 0, we introduce two weights" u 0 - 1 A ( w / ~ ) 16, Ul -- 1 A (X/w) 8. Then IlloguoIISMO, IIlOgUlllSMO <~ C C(lllogwllsMO) by Lemma 8.2. Since 1 ~< u0 4- Ul, by Lemma 7.9 we find g E H ~ ( u o ) , h E H ~ ( u l ) such that 1 -- g + h, Igl ~< Cuo, Ilhll ~< C u l , where here and below all constants are determined by IIlog w IIBMo. We do this for each )v - 2 n, n E Z, and denote the resulting functions by gn and hn. Next we put qgn -- gn - gn+l m hn+l - hn, then IgOn[ ~< cmin{ (2-nw)16, (2nw_l)8}.
(8.1)
We claim that {qgn} is the required analytic decomposition of unity. Indeed, (iv) is clear; the convergence of the series a.e. easily follows from (8.1), and the sum telescopes. Again by (8.1), Iqgnll/Sw <. C2 n, which is (i). We verify (ii) (condition (iii) is proved similarly). Let ek -- {2k ~< w < 2k+1}, then, again by (8.1),
Z2n[qgnll/8~c~2n(~2-2n22kXek nEZ nEZ k~n
4-Z2n2-kXek). k>n
Changing the order of summation, we see that the latter expression is dominated by Y~k 2k Xek <. C w.
Now we sketch the proof of the implication (2) =:~ (3). Let {qgj} be an analytic decomposition of unity for w. We write qgj -- Oj lit8 with Oj inner and ~ j outer, and put
Qs - Z
4R(s 4t ,
jez
where R is the Riesz projection. Then Q is the required operator. We only check the weak type (1, 1) property of Q; the weak type (1.1) for Q* is similar, and the L P ( w ) -
Uniform algebras as Banach spaces
701
boundedness of Q for 1 < p < cx~ is simpler. Clearly the values of Q are analytic functions, and Q fixes analytic functions because ~ qgj = 1. D LEMMA 8.3. An operator T acting from a subset of L l(#) to measurable functions is of
weak type (1, 1) if and only if it satisfies the estimate
f
1/2 1/2 [Tfll/Zlgld# <~C l l f l t,(~)llgllL,(~)llglll~(~).
PROOF HINT. To prove the "if" part, take g = XE, where E C {[Tfl > Z} is an arbitrary set of finite measure. Now we check the estimate of Lemma 8.3 for Q, using the fact that it is true for R:
f
Iofll/2lg[w
<~
zf
I~jlzlg(fv/4)ll/21glw
J
c
e; f
lR(fJ)l'/21Ogl
J
1/2
1/2
~< C I I g l I ~ Z Z z J H f o 4 [L' IlOJgllL' J
~< CIIgll1/2l Z z J l I f O j l I L J
,
)1/2( ~ 2 J l l g O j l l L , )1/2
oc 1/2 1/2 <~ Cllglll/ZllfllLJ(w)llglltl(w).
J
[]
See [28] for more information on Bourgain projections. To illustrate their usefulness, we mention an application to conformal mapping. Let G be a Jordan domain with rectifiable boundary. The analogues for G of the classical HP-spaces are the spaces EP(G) consisting of the analytic functions f on G for which sup f• If(z)lPl dzl is finite, where {gn} is a fixed sequence of rectifiable curves that tend to OG in a natural sense. The domain G is a Smirnov domain if the derivative g l of the conformal mapping g of the unit disk onto G is outer. In this case we may identify the scale E p (G) with the scale H p (Ig'l) of weighted Hardy spaces on the disk. For a Smirnov domain G with conformal map g, it is important to know when log Igll is in BMO. See [40] for a detailed discussion. The following theorem is an immediate consequence of Theorem 8.1. THEOREM 8.4. The following statements are equivalent for the conformal mapping g of
the unit disk onto a Smirnov domain G. (1) log Ig'l e BMO. (2) There is an operatorprojecting LP(OG) onto EP(G) for all p E (1, cx~) and having weak type (1, 1). (3) There is an operatorprojecting LP(OG) onto EP(G) for two differentvalues of p.
702
T.W. Gamelin and S. V. Kislyakov
Theorem 8.4 together with a theorem of David [13] yields a proof of the (known) fact that if the arc-length measure on the boundary of G satisfies a Carleson condition, then log Ig'l ~ BMO. The Carleson condition is that the length of OG A Dr is bounded by cr for any small disk Dr of radius r. By David's theorem, this implies that the Cauchy integral over OG is a bounded operator on LP(OG) for 1 < p < e~. In particular, condition (3) above holds, and from (1) we obtain log Ig'l ~ BMO. Note that in general, if condition (1) is fulfilled, then the Bourgain projection in statement (2) has the form (8.2) and is not the Cauchy integral operator. Finally, we note that the theory discussed in Section 7 can be carried over, nearly word for word, to the Hardy spaces related to a weak-star Dirichlet algebra, as can the implications (1) ==~ (2) ==~ (3) in Theorem 8.1. See [48] for background on weak-star Dirichlet algebras.
9. Perturbation of uniform algebras Often a Banach function space remembers almost nothing about the set on which the functions are defined. For example, if p is fixed, the spaces LP(S2, #) for # separable and atomless are all isometric. The reason is that, up to isomorphism in a proper sense, there are no separable atomless measures other than Lebesgue measure on [0, 1]. Something similar occurs in the context of the spaces C(K). The celebrated Milyutin theorem (see [51, III.D, w asserts that for any uncountable compact metric space K, C(K) is linearly homeomorphic to C[0, 1]. On the other hand, if we do not change the norm of C(K) by too much, we keep K in sight. The Amir-Cambern theorem (see [25]) asserts that if T is a linear isomorphism of two C(K)-spaces such that IITll lIT -111 < 2, then the underlying compact spaces are homeomorphic. This leads us to consider linear isomorphisms of uniform algebras that are not too far from being isometries. For an exposition of this area, see [25]. The idea of nearness of two uniform algebras can be given several equivalent formulations, but we focus only on the Banach space notion of nearness to an isometry. We say that two uniform algebras are (1 + e)-isomorphic if there is a linear isomorphism T between them that satisfies IIT II IIT-11l < 1 + e, that is, if the Banach-Mazur distance between them is less than log(1 + e). Algebras that are 1-isomorphic in this sense are isometrically isomorphic as Banach spaces, and for these we have the following. THEOREM (Nagasawa [36]). If two uniform algebras A and B are isometrically isomorphic as Banach spaces, then they are isometrically isomorphic as uniform algebras. In particular, their maximal ideal spaces are then homeomorphic, as are their Shilov boundaries. As a consequence, if K1 and K2 are compact subsets of the complex plane such that the algebras A (K1) and A (K2) are linearly isometrically isomorphic, then there is a homeomorphism of K1 onto K2 that maps the interior of K1 conformally onto the interior of K2. On the other hand, deformation of the compact set K leads to linear isomorphisms of A ( K ) that are close to being isometries. For example, consider the scale of annuli Gr = {r < Izl < 1}, and the associated algebras A(Gr). Since these annuli are
Uniform algebras as Banach spaces
703
conformally distinct, no two of the algebras A(Gr) can be isometric. On the other hand, the linear operator T from Ar to As defined by
T
an Z n n=---~
m
Z n=--~
an
Zn
-k-
an Z n n--O
is a linear isomorphism of Ar onto As, and further IITII IIT-1 II ~ 1 as s ~ r. We say that a uniform algebra A is stable if there is e > 0 such that any uniform algebra B that is (1 + e) isomorphic to A is actually isometrically isomorphic to A. Thus the algebras C ( K ) are stable, while the annulus algebras are not. Rochberg [43] proved in 1972 that the disk algebra CA is stable, and he went on to study the perturbations of the algebras A ( K ) for K a finitely connected subset of the complex plane with smooth boundary, or more generally for K a finite bordered Riemann surface. The flavor of his work is given by the following result. THEOREM (Rochberg). Let K be a finite bordered Riemann surface. Then for e > 0 small, any uniform algebra B that is (1 + e)-isomorphic to A (K) has the form A (J) for a compact bordered Riemann surface J that is a deformation of K by a quasi-conformal homeomorphism with dilatation tending to 0 with e. For expository accounts of these results, see [44,45]. More recently, Jarosz [26] was able to prove that the nonseparable algebra H ~ ( A ) is also stable. Meanwhile there is currently no known compact set K in the complex plane with nonempty interior such that CA (K) can be shown to be linearly nonisomorphic to the disk algebra. Conformal mapping theory and the relation CA ~ (CA ~3 CA @ " ")co proved by Wojtaszczyk (see [51, III.E, w12]) suggest that a compact set K for which CA (K) (or P ( K ) , or R ( K ) ) is proper but is not isomorphic to the disk algebra should not be too simple (if it exists).
10. The dimension conjecture It is natural to ask how linear topological properties of an algebra of analytic functions of several complex variables reflect the geometry of the underlying domain. The oldest problem along these lines is to determine what effect the number of variables (dimension) has. DIMENSION CONJECTURE. If G1 and G2 are bounded domains in C n and C m respectively, with n ~: m, then the spaces A(G1) and A(G2) are not linearly homeomorphic, nor are H ~ ( G 1 ) and H ~ ( G 2 ) . We discuss briefly some results related to the dimension conjecture. The main references are [11] and [39, w 11]. It is most natural to examine the dimension conjecture first for the polydisks A n and the balls Bn. In [11 ] it was proved that A (A n) is not linearly homeomorphic to A (Am)
704
T. W. Gamelin and S. V. Kislyakov
if m r n. The invariant distinguishing the spaces (in fact, their duals) is the behavior of certain vector-valued multiindexed martingales on the measure space F ~ • ... x F ~ with natural filtration; the martingales in question must have some additional complex analytic structure. The method in [ 11 ] yields the following theorem. THEOREM 1 0.1. Let U1 . . . . . Un, V1 . . . . . Vm be strictly pseudoconvex domains with C2-smooth boundary (the dimension may vary f r o m one domain to another). I f m > n, then A ( V l x . . . x Vm)* does not e m b e d in A ( U ! x . . . x Un)* as a closed subspace. Theorem 10.1 includes the previously known result that the spaces A ( B m ) and A ( A n) are not linearly isomorphic for m, n ~> 2. It had been shown that A (An) * does not embed in a direct sum of an L 1-space and a separable space (see [39, w 11]), whereas A ( B m ) * is such a direct sum by Theorem 4.3. Very little is known beyond Theorem 10.1. Currently it is not even known whether the ball algebras A (Bm) are mutually nonisomorphic for m / > 2. They are all distinct from A ( B 1 ) -- A ( A ) = CA. The latter space is a subspace of C ( F ) with separable annihilator, whereas A ( B m ) for m ~> 2 does not embed in C ( K ) as a subspace with separable annihilator (see [39, w 11 ]). The series { H ~ ( A n ) } seems to be quite similar to {A(An)}; however, in general the method of [11] is not applicable to H ~ ( A n ) . It is known only that H ~ ( A ) differs from H ~ ( A n) for n ~> 2. Again, see [11,39] for proofs. For comparison, we describe the situation concerning the Hardy spaces H 1( F n) and H I ( O B n ) . The spaces of the first series are not isomorphic to one another (see [6,5]), whereas those of the second series are all isomorphic (see [52]). More recently, it was shown that for any strictly pseudoconvex domain with smooth boundary the corresponding space H 1 is isomorphic to the classical H 1 in the disk (see [2]). Finally, note that in contrast to the current state of affairs in one complex variable, it is possible to find in several complex variables many examples of nonisomorphic spaces A (G) where the underlying domains G have the same dimension. For instance, from Theorem 10.1 it follows that A(A 4) is not linearly isomorphic to A ( B 2 x B2).
References [1] H. Alexander and J. Wermer, Several Complex Variables and Banach Algebras, 3rd edn, Springer-Verlag (1998). [2] H. Arai, Degenerate elliptic operators, Hardy spaces and diffusions on strongly pseudoconvex domains, Tohoku Math. J. 46 (4) (1994), 469-498. [3] R.M. Aron, B.J. Cole and T.W. Gamelin, Weak-star continuous analytic functions, Canad. J. Math. 47 (1995), 673-683. [4] J. Bergh and J. Lrfstrrm, Interpolation Spaces, Springer-Verlag(1976). [5] J. Bourgain, The nonisomorphism of H 1-spaces in one and several variables, J. Funct. Anal. 46 (1) (1982), 45-57. [6] J. Bourgain, The nonisomorphism of H 1-spaces in a different number of variables, Bull. Soc. Math. Belg., Ser. B 35 (2) (1983), 127-136. [7] J. Bourgain, On weak completeness of the dual spaces of analytic and smooth functions, Bull. Soc. Math. Belg., Ser. B 35 (1) (1983), 111-118.
Uniform algebras as Banach spaces
705
[8] J. Bourgain, New Banach space properties of the disc algebra and H ~ , Acta Math. 152 (1984), 1-48; 387-391. [9] J. Bourgain, Bilinearforms on H ~ and bounded bianalyticfunctions, Trans. Amer. Math. Soc. 286 (1984), 313-337. [10] J. Bourgain, The Dunford-Pettis property for the ball-algebras, the polydisc-algebras and the Sobolev spaces, Studia Math. 77 (1984), 245-253. [11] J. Bourgain, The dimension conjecture for polydisk algebras, Israel J. Math. 48 (4) (1984), 289-304. [12] B.J. Cole and T.W. Gamelin, Tight uniform algebras and algebras of analytic functions, J. Funct. Anal. 46 (1982), 158-220. [13] G. David, Op~rateurs integraux singuliers sur certaines courbes du plan complexe, Ann. Sci. l~cole Norm. Sup. 17 (1) (1984), 157-189. [ 14] A.M. Davie, Bounded approximation of analytic functions, Proc. Amer. Math. Soc. 32 (1972), 128-133. [ 15] J. Diestel, H. Jarchow and A. Tonge, Absolutely Summing Operators, Cambridge University Press (1995). [16] S. Dineen, Complex Analysis on Infinite Dimensional Spaces, Springer-Verlag (1999). [ 17] EL. Duren, Theory of H P Spaces, Academic Press (1970). [ 18] T.W. Gamelin, Uniform algebras on plane sets, Approximation Theory, G.G. Lorentz, ed., Academic Press (1973), 101-149. [19] T.W. Gamelin, Uniform Algebras, 2nd edn, Chelsea (1984). [20] T.W. Gamelin, Analytic functions on Banach spaces, Complex Function Theory, Gauthier and Sabidussi, eds, Kluwer Academic (1994), 187-233. [21] D.J.H. Garling, On the dual of a proper uniform algebra, Bull. London Math. Soc. 21 (1989), 279-284. [22] J.B. Garnett, Bounded Analytic Functions, Academic Press (1981). [23] J. Guti6rrez, J. Jaramillo and J.L. Llavona, Polynomials and geometry of Banach spaces, Extracta Math. 10 (2) (1995), 227-228. [24] K. Hoffman, Banach Spaces of Analytic Functions, Prentice-Hall (1962). [25] K. Jarosz, Perturbations of Banach Algebras, Lecture Notes in Math. 1120, Springer-Verlag (1985). [26] K. Jarosz, H ~ ( D ) is stable, J. London Math. Soc. 37 (1988), 490-498. [27] N.J. Kalton, Complex interpolation ofHardy-type subspaces, Math. Nachr. 171 (1995), 227-228. [28] S.V. Kislyakov, Bourgain's analytic projection revisited, Proc. Amer. Math. Soc. 126 (11) (1988), 33073314. [29] S.V. Kislyakov, Proper uniform algebras are uncomplemented, Soviet Math. Dokl. 40 (3) (1990), 584-586. [30] S.V. Kislyakov, Absolutely summing operators on the disc algebra, St. Petersburg Math. J. 3 (4) (1992), 705-774. [31] S.V. Kislyakov, Interpolation of HP-spaces: some recent developments, Function Spaces, Interpolation Spaces, and Related Topics, Israel Math. Conf. Proceedings 13, Amer. Math. Soc., Providence, RI (1999). [32] S.V. Kislyakov and Q. Xu, Interpolation of weighted and vector-valued Hardy spaces, Trans. Amer. Math. Soc. 343 (3) (1994), 1-34. [33] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I, Ergebn. Math. Grenzgeb. 92, Springer-Verlag (1977). [34] H. Milne, Banach space properties of uniform algebras, Bull. London Math. Soc. 4 (1972), 323-326. [35] J. Mujica, Complex Analysis in Banach Spaces, North-Holland, Amsterdam (1986). [36] M. Nagasawa, Isomorphisms between commutative Banach algebras with application to rings of analytic functions, Kodai Math. Sem. Rep. 11 (1959), 182-188. [37] D. Oberlin, A Rudin-Carleson theorem for uniformly convergent Fourier series, Michigan Math. J. 27 (1980), 309-313. [38] G. Pisier, Interpolation between Hardy spaces and noncommutative generalizations. I, Pacific J. Math. 155 (1992), 341-368. [39] A. Peiczyfiski, Banach Spaces of Analytic Functions and Absolutely Summing Operators, Conf. Board of Math. Sciences, Vol. 30, Amer. Math. Soc. (1977). [40] Ch. Pommerenke, Boundary Behaviour of Conformal Maps, Springer-Verlag (1992). [41 ] IT Privalov, Boundary Properties of Analytic Functions, GITTL, Moscow, Leningrad 1950 (Russian). [42] R.M. Range, Holomorphic Functions and Integral Representations in Several Complex Variables, SpringerVerlag (1986). [43] R. Rochberg, The disk algebra is rigid, Proc. London Math. Soc. 39 (1979), 119-129.
706
T. W. Gamelin and S. V. Kislyakov
[44] R. Rochberg, Deformation theory for uniform algebras: an introduction, Proceedings of the Conference on Banach Algebras and Several Complex Variables (New Haven, Conn., 1983), Contemp. Math. 32, Amer. Math. Soc. (1984), 209-216. [45] R. Rochberg, Deformation theory for algebras of analytic functions, Deformation Theory of Algebras and Structures and Applications (I1Ciocco, 1986), NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci. 247, Kluwer (1988), 501-535. [461 S.E Saccone, The Petczyhski property for tight subspaces, J. Funct. Anal. 148 (1) (1997), 86-116. [47] S.F. Saccone, Function theory in spaces of uniformly convergent Fourier series, to appear. [48] T.P. Srinivasan and J.-K. Wang, Weak* Dirichlet algebras, Function Algebras, E Birtel, ed., Scott-Foresman (1966), 216-249. [49] E.L. Stout, The Theory of Uniform Algebras, Bogdon and Quigley (1971). [50] J. Wermer, Function rings and Riemann surfaces, Ann. Math. 67 (1958), 45-71. [51] P. Wojtaszczyk, Banach Spaces for Analysts, Cambridge University Press (1990). [52] T. Wolniewicz, On isomorphisms between Hardy spaces on complex balls, Ark. Math. 27 (1) (1989), 155168. [53] A. Zygmund, Trigonometric Series, Vol. I, II, Cambridge University Press (1959).
CHAPTER
17
Euclidean Structure in Finite Dimensional Normed Spaces
Apostolos A. Giannopoulos Department of Mathematics, University of Crete, Heraklion, Greece E-mail: apostolo @math. uch.gr
Vitali D. Milman* Department of Mathematics, Tel Aviv University, Tel Aviv, Israel E-mail: vitali@math, tau.ac, il
Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Classical inequalities and isotropic positions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Classical inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Extremal problems and isotropic positions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Background from classical convexity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Steiner's formula and Urysohn's inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Geometric inequalities of "hyperbolic" type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Continuous valuations on compact convex sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Dvoretzky's theorem and concentration of measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Concentration of measure on the sphere and a proof of Dvoretzky's theorem . . . . . . . . . . . . . 4.3. Probabilistic and global form of Dvoretzky's theorem . . . . . . . . . . . . . . . . . . . . . . . . . 4.4. Applications of the concentration of measure on the sphere . . . . . . . . . . . . . . . . . . . . . . 4.5. The concentration phenomenon: L6vy families . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6. Dvoretzky's theorem and duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.7. Isomorphic versions of Dvoretzky's theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. The low M*-estimate and the quotient of subspace theorem . . . . . . . . . . . . . . . . . . . . . . . . 5.1. The low M*-estimate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. The g-position . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
709 710 710 711 718 726 726 729 732 735 735 737 740 743 744 746 748 749 749 751
*The authors acknowledge the hospitality of the Erwin SchrSdinger International Institute for Mathematical Physics in Vienna, where this work has been completed. The second named author was supported in part by the Israel Science Foundation founded by the Academy of Sciences and Humanities. H A N D B O O K OF THE GEOMETRY OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 707
708
A.A. Giannopoulos and V.D. Milman
5.3. The quotient of subspace theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Variants and applications of the low M*-estimate . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Isomorphic symmetrization and applications to classical convexity . . . . . . . . . . . . . . . . . . . . . 6.1. Estimates on covering numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2. Isomorphic symmetrization and applications to classical convexity . . . . . . . . . . . . . . . . . . 7. Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1. The hyperplane conjecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2. Geometry of the B a n a c h - M a z u r compactum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3. Symmetrization and approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4. Quasi-convex bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5. Type and cotype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6. Nonlinear type theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
752 754 756 756 759 763 763 765 768 769 769 771 772
Euclidean structure in finite dimensional normed spaces
709
1. Introduction In this article we discuss results which stand between geometry, convex geometry, and functional analysis. We consider the family of n-dimensional normed spaces and study the asymptotic behavior of their parameters as the dimension n grows to infinity. Analogously, we study asymptotic phenomena for convex bodies in high dimensional spaces. This theory grew out of functional analysis. In fact, it may be viewed as the most recent one among many examples of directions in mathematics which were born inside this field during the twentieth century. Functional analysis was developed during the period between the World Wars by the Polish school of mathematics, an outstanding school with broad interests and connections. The influence of the ideas of functional analysis on mathematical physics, on differential equations, but also on classical analysis, was enormous. The great achievements and successful applications to other fields led to the creation of new directions (among them, algebraic analysis, noncommutative geometry and the modem theory of partial differential equations) which in a short time became autonomous and independent fields of mathematics. Thus, in the last decades of the twentieth century, geometric functional analysis and even more narrowly the study of the geometry of Banach spaces became the main line of research in what remained as "proper" functional analysis. The two central themes of this theory were infinite dimensional convex bodies and the linear structure of infinite dimensional normed spaces. Several questions in the direction of a structure theory for Banach spaces were asked and stayed open for many years. Some of them can be found in Banach's book. Their common feature was a search for simple building blocks inside an arbitrary Banach space. For example: does every Banach space contain an infinite unconditional basic sequence? Is every Banach space decomposable as a topological sum of two infinite dimensional subspaces? Is it true that every Banach space is isomorphic to its closed hyperplanes? Does every Banach space contain a subspace isomorphic to some gp or co ? This last question was answered in the negative by Tsirelson (1974) who gave an example of a reflexive space not containing any gp. Before Tsirelson's example, it had been realized by the second named author that the notion of the spectrum of a uniformly continuous function on the unit sphere of a normed space was related to this question and that the problem of distortion was a central geometric question for approaching the linear structure of the space. Although Tsirelson's example was a major breakthrough and introduced a completely new construction of norm, the search for simple linear structure continued to be the aim of most of the efforts in the geometry of Banach spaces. We now know that infinite dimensional Banach spaces have much more complicated structure than what was assumed (or hoped). All the questions above were answered in the negative in the middle of the 90s, starting with the works of Gowers and Maurey, Gowers, Odell and Schlumprecht. Actually, the line of thought related to Tsirelson's example and the concepts of spectrum and distortion were the most crucial for the recent developments. The systematic quantitative study of n-dimensional spaces with n tending to infinity started in the 60s, as an alternative approach to several unsolved problems of geometric functional analysis. This study led to a new and deep theory with many surprising consequences in both analysis and geometry. When viewed as part of functional analy-
710
A.A. Giannopoulos and V.D. Milman
sis, this theory is often called local theory (or asymptotic theory of finite dimensional normed spaces). However, it adopted a significant part of classical convexity theory and used many of its methods and techniques. Classical geometric inequalities such as the Brunn-Minkowski inequality, isoperimetric inequalities and many others were extensively used and established themselves as important technical tools in the development of local theory. Conversely, the study of geometric problems from a functional analysis point of view enriched classical convexity with a new approach and a variety of problems: The "isometric" problems which were typical in convex geometry were replaced by "isomorphic" ones, with the emphasis on the role of the dimension. This change led to a new intuition and revealed new concepts, the concentration phenomenon being one of them, with many applications in convexity and discrete mathematics. This natural melting of the two theories should perhaps correctly be called asymptotic (or convex) geometric analysis. This paper presents only some aspects of this asymptotic theory. We leave aside typecotype theory and other connections with probability theory, factorization results, covering and entropy (besides a few results we are going to use), connections with infinite dimension theory, random normed spaces, and so on. Other articles in this collection will cover these topics and complement these omissions. On the other hand, we feel it is necessary to give some background on convex geometry: This is done in Sections 2 and 3. The theory as we build it below "rotates" around different Euclidean structures associated with a given norm or convex body. This is in fact a reflection of different traces of hidden symmetries every high dimensional body possesses. To recover these symmetries is one of the goals of the theory. A new point which appears in this article is that all these Euclidean structures that are in use in local theory have precise geometric descriptions in terms of classical convexity theory: they may be viewed as "isotropic" ones. Traditional local theory concentrates its attention on the study of the structure of the subspaces and quotient spaces of the original space (the "local structure" of the space). The connection with classical convexity goes through the translation of these results to a "global" language, that is, to equivalent statements pertaining to the structure of the whole body or space. Such a comparison of "local" and "global" results is very useful, opens a new dimension in our study and will lead our presentation throughout the paper. We refer the reader to the books of Schneider [177] and of Burago and Zalgaller [43] for the classical convexity theory. Books mainly devoted to the local theory are the ones by: Milman and Schechtman [150], Pisier [164], Tomczak-Jaegermann [195].
2. Classical inequalities and isotropic positions 2.1. Notation 2.1.1. We study finite-dimensional real normed spaces X = (•n, I1" II). The unit ball K x of such a space is an origin-symmetric convex body in R n which we agree to call a body. There is a one to one correspondence between norms and bodies in Rn: If K is a body, then Ilxll = min{)~ > 0: x 9 ;~K} is a norm defining a space XK with K as its unit ball. In this way bodies arise naturally in functional analysis and will be our main object of study.
Euclidean structure in finite dimensional normed spaces
711
If K and T are bodies in ~n we can define a multiplicative distance d(K, T) by
d(K, T) -- inf{ab: a, b > O, K c bT, T cC_.aK}. The natural distance between the n-dimensional spaces XK and XT is the Banach-Mazur distance. Since we want to identify isometric spaces, we allow a linear transformation and set
d(XK, Xv) - inf{d(X, uT)" u E GLn }. In other words, d ( X x , XT) is the smallest positive number d for which we can find u ~ GLn such that K c_ uT c dK. We clearly have d(XK, XT) ~> 1 with equality if and only if XK and XT are isometric. Note the multiplicative triangle inequality d(X, Z) <~ d(X, Y)d(Y, Z) which holds true for every triple of n-dimensional spaces. 2.1.2. We assume that Rn is equipped with a Euclidean structure (.,-) and denote the corresponding Euclidean norm by l" I. Dn is the Euclidean unit ball and S n-1 is the unit sphere. We also write I" I for the volume (Lebesgue measure) in R ' , and # for the Haar probability measure on the orthogonal group O (n). Let G.,k denote the Grassmannian of all k-dimensional subspaces of R ~. Then, O(n) equips G,,k with a Haar probability measure v,,k satisfying
Vn,k(A)- #{u E O(n)" uEk E A} for every Borel subset A of Gn,k and every fixed element Ek of Gn,k. The rotationally invariant probability measure on S n- 1 will be denoted by a. 2.1.3. Duality plays an important role in the theory. If K is a body in R n, its polar body is defined by
K~
{y E R n" I(x, y)l <~ l forallx 6 K}.
That is, IlYlIKo - - m a x x e K I(x, Y)I. Note that XKO = X ~ " K ~ is the unit ball of the dual space of X. It is easy to check that d(X, Y) -- d(X*, Y*).
2.2. Classical inequalities
(a) The Brunn-Minkowski inequality. Let K and T be two convex bodies in ~n. If K 4- T denotes the Minkowski sum {x 4- y" x E K, y E T} of K and T, the Brunn-Minkowski inequality states that IK 4- TI 1/" >1 IKI ~/" 4- ITI l/',
(1)
with equality if and only if K and T are homothetical. Actually, the same inequality holds for arbitrary nonempty compact subsets of R n.
A.A. Giannopoulos and V.D. Milman
712
One can rewrite (1) in the following form: For every ~. E (0, 1), I~.K -I- (1 - X)TI 1/n >1 X I K I 1/n -t- (1 - ~,)ITI ~/n.
(2)
Then, the arithmetic-geometric means inequality gives a dimension free version: I~.K -t-(1 - ~ . ) T I > IKIXITI ~-~.
(3)
There are several proofs of the Brunn-Minkowski inequality, all of them related to important ideas. We shall sketch only two lines of proof. The first (historically as well) proof is based on the Brunn concavity principle: Let K be a convex body in ~n and F be a k-dimensional subspace. Then, the function f " F • --+ R defined by f (x) -- [K A (F + x)[ 1/~ is concave on its support.
The proof is by symmetrization. Recall that the Steiner symmetrization of K in the direction of 0 E S n-1 is the convex body So (K) consisting of all points of the form x + )v0, where x is in the projection Po ( K ) of K onto 0 • and [)v[ ~< 89• length(x + 1R0) N K. Steiner symmetrization preserves convexity: in fact, this is the Brunn concavity principle for k = 1. The proof is elementary and essentially two dimensional. A well known fact which goes back to Steiner and Schwarz but was later rigorously proved in [45] (see [43]) is that for every convex body K one can find a sequence of successive Steiner symmetrizations in directions 0 E F so that the resulting convex body K has the following property: K A ( F + x) is a ball with radius r ( x ) , and [K A (F + x)l -- [K A (F + x)l for every x ~ F • Convexity of K implies that r is concave on its support, and this shows that f is also concave. 71 The Brunn concavity principle implies the Brunn-Minkowski inequality. If K, T are convex bodies in IRn, we define K1 = K • {0}, T1 -- T • {1 } in IRn+l and consider their convex hull L. If L(t) -- {x ~ IRn" (x, t) ~ L}, t ~ IR, we easily check that L(0) -- K, L(1) = T, and L(1/2) = K+T Then, the Brunn concavity principle for F -- ]1~n shows that 2 " K+T 2
1/n
1
1
> -~IKI ~/~ + -~ITI~/n
[3
(4)
A second proof of the Brunn-Minkowski inequality may be given via the KnOthe map: Assume that K and T are open convex bodies. Then, there exists a one to one and onto map cp:K --+ T with the following properties: (i) q~ is triangular: the ith coordinate function of 4~ depends only on Xl . . . . . xi. That is, ~ ( x l . . . . . x~) - ( ~ l ( x l ) , ~ : ( x l , x2) . . . . . ~ ( x l . . . . . x~)).
(5)
Euclidean structure in finite dimensional normed spaces
713
0q~i (ii) The partial derivatives 7Z/are nonnegative on K, and the determinant of the Jacobian of 4) is constant. More precisely, for every x E K [-I Od/)i (det J~)(x) -- i=1
~-X/(x) -
ITI
(6)
Igl
The map 05 is called the Kn6the map from K onto T. Its existence was established in [103] (see also [ 150, Appendix I]). Observe that each choice of coordinate system in ~n produces a different Kn6the map from K onto T. It is clear that (I + 4~)(K) _ K + T, therefore we can estimate IK + T] using the arithmetic-geometric means inequality as follows:
IK+TI
~>f(/
+4~)(K)
-
dX--fKldetJ~+r + Oxi/dx>
(1
+ detJ2/n) n dx
i=1
=
1+
]K]
K J/n
= (Igll/n + Irll/n) n. This proves the Brunn-Minkowski inequality.
(7) D
Alternatively, instead of the Kn6the map one may use the Brenier map O:K -+ T, where K and T are open convex bodies. This is also a one to one, onto and "ratio of volumes" preserving map (i.e., its Jacobian has constant determinant), with the following property: There is a convex function f c C2(K) defined on K such that gr = V f . A remarkable property of the Brenier map is that it is uniquely determined. Existence and uniqueness of the Brenier map were proved in [42] (see also [ 126] for a different proof and important generalizations). It is clear that the Jacobian Jo = Hess f is a symmetric positive definite matrix. Again we have (I + 7r)(K) ___K + T, hence
IK-+-TI >1~ldetJ/+g~(x)ldx- s det(I + H e s s f ) d x /7
(8)
=s t=l
where/~i (X) are the nonnegative eigenvalues of Hessf. Moreover, by the ratio of volumes preserving property of gr, we have Ui%l )~i(x) --]TI/IK] for every x 6 K. Therefore, the
A.A. Giannopoulos and V.D. Milman
714
arithmetic-geometric means inequality gives
IK + TI >1 fK
( In ] 1+
)U(X)
1/n)ndx =
(IKI 1/n + IZl~/n) n
Fq
(9)
i=1
This proof has the advantage of providing a description for the equality cases: either K or T must be a point, or K must be homothetical to T. Let us describe here the generalization of Brenier's work due to McCann: Let #, v be probability measures on ~n such that # is absolutely continuous with respect to Lebesgue measure. Then, there exists a convex function f such that V f:It~ n --+ R n is defined # almost everywhere, and v(A) = # ( ( V f ) - l (A))) for every Borel subset A of IRn ( V f pushes forward/z to v). If both/z, v are absolutely continuous with respect to Lebesgue measure, then the Brenier map V f has an inverse (V f ) - 1 which is defined v-almost everywhere and is also a Brenier map, pushing forward v to #. A regularity result of Caffarelli [44] (see [5]) states that if T is a convex bounded open set, f is a probability density on IRn, and g is a probability density on T such that (i) f is locally bounded and bounded away from zero on compact sets, and (ii) there exist Cl, C2 > 0 such that Cl ~ g(y) ~ C2 for every y e T, then, the Brenier map V f : (IRn, f dx) --+ (IRn , g dx) is continuous and belongs locally to the H61der class C c~ for some ot > 0. The following recent result [5] makes use of all this information: FACT 1. Let K1 and K2 be open convex bounded subsets of 1Rn with IKll = IK2I = 1.
There exists a Cl-diffeomorphism r : K1 --+ K2 which is volume preserving and satisfies K1 +)vK2 = {x + X ~ ( x ) : x E K1},
& > 0.
(10)
PROOF. Let p be any smooth strictly positive density on ~n. Consider the Brenier maps
~i = V fi " (]t~n, p d x ) --->.(Ki, dx),
i = 1, 2.
(11)
Caffarelli's result shows that they are C 1-smooth. We now apply the following theorem of Gromov [82] (for a proof, see also [5]): FACT 2. (i) Let f : ]~n __~ ]I~ be a C2-smooth convex function with strictly positive Hessian. Then, the image of the gradient map Im V f is an open convex set. (ii) If f l, f2 are two such functions, then I m ( V f l + V f2) = I m ( V f l ) + Im(Vf2). It follows that, for every )~ > 0, K1 + )vK2 -- { V f l (x) + ~Vf2(x)" x E I[~n }.
(12)
Then, one can check that the map r = ~P2 o (l~rl) - 1 :K1 -+ K2 satisfies all the conditions of Fact 1.
Euclidean structure in finite dimensional normed spaces
715
The existence of a volume preserving l/r" K1 --+ K2 such that (I + ~ ) (K1) = K1 -~- K2 covers a "weak point" of the Kn6the map and should have important applications to convex geometry. We discuss some of them in Section 3.2. (b) Consequences o f the Brunn-Minkowski inequality ( b l ) The isoperimetric inequality f o r convex bodies. body K is defined by
O(K)-
lim
[K + e D n ] -
s--+O+
IK]
e
The surface area O(K) of a convex
.
(13)
It is a well-known fact that among all convex bodies of a given volume the ball has minimal
surface area. This is an immediate consequence of the Brunn-Minkowski inequality: If K is a convex body in R" with ]K] -- ]r Dn ], then for every e > 0 (14)
IK + eDnl '/n >1 IKI '/n + s l O , I~/" = (r + s)ID~I ~/~. It follows that the surface area O(K) of K satisfies 0 ( K ) --
lim s--+O+
[K +
eDn[-
[K[ /> lira
e
s-+O+
(r + e) n - r n
ID, I
= nlDnll/nlK[(n-l)/n
(15)
with equality if K -- r Dn. The question of uniqueness in the equality case is more delicate. (b2) The spherical isoperimetric inequality. Consider the unit sphere S n-1 with the geodesic distance p and the rotationally invariant probability measure or. For every Borel subset A of S " - 1 and for every s > 0, we define the s-extension of A" Ae -- {x ~ S n-l" p(x, A) <<,s}.
(16)
Then, the isoperimetric inequality for the sphere is the following statement: A m o n g all Borel subsets A o f S n-1 with given measure ~ ~ (0, 1), a spherical cap B ( x , r) o f radius r > 0 such that cr(B(x, r)) -- ~ has minimal s-extension f o r every s > O. This means that if A _c S " - I and (r(A) = cr(B(x0, r)) for some xo E S n-1 and r > 0, then a ( A s ) >~ cr(B(xo, r + e))
(17)
for every s > 0. Since the a - m e a s u r e of a cap is easily computable, one can give a lower bound for the measure of the s-extension of any subset of the sphere. We are mainly interested in the case or(A) = 1/2, and a straightforward computation (see [61]) shows the following:
A.A. Giannopoulos and V.D. Milman
716 THEOREM
2.2.1. If A is a Borel subset of S n+l and or(A) - 1/2, then (18)
cr(Ae) >~ 1 - x / ~ e x p ( - e 2 n / 2 ) f o r every e > O.
(The constant ~/-~/8 may be replaced by a sequence of constants an tending to 1/2 as n -----~ o ~ . )
The spherical isoperimetric inequality can be proved by spherical symmetrization techniques (see [ 176] or [61 ]). However, it was recently observed [ 11 ] that one can give a very simple proof of an estimate analogous to (18) using the Brunn-Minkowski inequality. The key point is the following lemma: LEMMA. Consider the probability measure lz(A) -- IAI/IDnl on the Euclidean unit ball Dn. If A, B are subsets of Dn with # ( A ) >~ or, # ( B ) ~ el, and if p ( A , B) -- inf{la - bl" a A, b E B} = p > O, then
oe <<.exp(--p2n/8). (In other words, if two disjoint subsets of Dn have positive distance p, then at least one of them must have small volume (depending on p) when the dimension n is high.) PROOF. We may assume that A and B are closed. By the Brunn-Minkowski inequality, # (-&~-) ~> or. On the other hand, the parallelogram law shows that if a 6 A, b 6 B then la -+- bl 2 -- 2lal 2 -+- 21bl 2 - la - bl 2 ~ 4 - p 2 . p2 It follows that 3 _ ~ c_ (1 - 7-) 1/2 Dn, hence
lZ ( A +
) ~ < (1 - p 2 ) n/
<~ exp(--p2n/8).
IS]
PROOF OF THEOREM 2.2.1 (with weaker constants). Assume that A c_ S n-1 with or(A) 1/2. Let e > 0 and define B.~-- (Ae) c c_ S n-1 . We fix )v e (0, 1) and consider the subsets A - U{tA" )~ <<.t <<.1} and B -- [~{tB" )~ <<.t <<.1} of Dn. These are disjoint with distance _~ )~e. The Lemma shows that # ( B ) ~< exp(-c)~ZeZn/8), and since # ( B ) -- (1 - )~n)~(B) we obtain o-(Ae) ~> 1 -
1 - 1)~--------~exp(_c)2e2n/8
)
We conclude the proof by choosing suitable )~ 6 (0, 1).
(19) []
Euclidean structure in finite dimensional normed spaces
717
(b3) C. Borell's lemma and Khinchine type inequalities. Let # be a Borel probability measure on R n. We say that # is log-concave if whenever A, B are Borel subsets of R n and )~ E (0, 1) we have # ( h A + (1 - )0B) ~> # ( A ) X # ( B ) 1-x.
(20)
The following lemma of C. Borell [25] holds for all log-concave probability measures: LEMMA. Let # be a log-concave Borel probability measure on R n, and A be a symmetric convex subset o f R n. If # ( A ) -- 0 > 1/2, then f o r every t ~ 1 we have t+l
(21) PROOF. Immediate by the log-concavity of #, after one observes that
Rn\A D -t+l
2
t-1 (Nn\ta) +--a. t+l
[-1
(22)
Let K be a convex body in R n. By the Brunn-Minkowski inequality we see that the measure #K defined by #K (A) = ]A A g ] / ] g ] is a log-concave probability measure. In this context, Borell's lemma tells us that if A is convex symmetric and if A N K contains more than half of the volume of K, then the proportion of K which stays outside t A decreases exponentially in t as t ~ + e c in a rate independent of the convex body K and the dimension n. This observation has important applications to the study of linear functions f (x) = (x, y), y 6 R n, defined on convex bodies. Let us denote by ]]f]]p the Lp n o r m with respect to the probability measure # K. Then, for every linear function f : K --~ R we have ]]fl]q~llfl]p~cpllfllq,
(23)
O < q < p,
where Cp are universal constants depending only on p. The left hand side inequality is just H61der's inequality, while the right hand side (in the case 1 ~< q < p) is a consequence of Borell's lemma (see [86]). One writes
[K[
[ f (x)[P dx --
p t P - l l z K ( { X E K: If(x)] ~> t})dt
(24)
and estimates # K ( { x E K: If(x)[ ~> t}) for large values of t using Borell's lemma with say A -- {x ~ Rn: If(x)[ ~< 3[[fllq}. The dependence of Cp on p is linear as p --+ ec. This is equivalent to the fact that the L 0~ (K) norm of f
IIf lIL e', (K) = i n f , ~, > 0: IX[
e x p ( l f ( x ) [ / X ) ~< 2
/
(25)
718
A.A. Giannopoulos and V.D. Milman
is equivalent to II f Ill. The question to determine the cases where c(p) ~_ ~ as p --+ ec in (23) is very important for the theory. This is certainly true for some bodies (e.g., the cube), but the example of the cross-polytope shows that it is not always so. Inverse H61der inequalities of this type are very similar in nature to the classical Khinchine inequality (and are sometimes called Khinchine type inequalities). In fact, the argument described above may be used to give proofs of the Kahane-Khinchine inequality (see [150, Appendix III]). Khinchine type inequalities do not hold only for linear functions. For example, Bourgain [28] has shown that if f :K --+ R is a polynomial of degree m, then Ilfllp ~< c(p, m)llfll2
(26)
for every p > 2, where c(p, m) depends only on p and the degree m of f (for this purpose, the Brunn-Minkowski inequality was not enough, and a suitable direct use of the Kn6the map was necessary). It was also recently proved [108] that (23) holds true for any norm f on R n. Finally the interval of values of p and q in (23) can be extended to ( - 1, +cx~) (see [ 146] for linear functions, [88] for norms).
2.3. Extremal problems and isotropic positions In the study of finite dimensional normed spaces one often faces the problem of choosing a suitable Euclidean structure related to the question in hand. In geometric language, we are given the body K in R n and want to find a specific Euclidean norm in ]~n which is naturally connected with our question about K. An equivalent (and sometimes more convenient) approach is the following: we fix the Euclidean structure in R n, and given K we ask for a suitable "position" u K of K, where u is a linear isomorphism of R n . That is, instead of keeping the body fixed and choosing the "right ellipsoid" we fix the Euclidean norm and choose the "right position" of the body. Most of the times the starting point is a question of the following type: we are given a functional f on convex bodies and a convex body K and we ask for the maximum or minimum of f (u K) over all volume preserving transformations u. We shall describe some very important positions of K which solve such extremal problems. What is interesting is that there is a simple variational method which leads to a description of the solution, and that in most cases the resulting position of K is isotropic. Moreover, isotropic conditions are closely related to the Brascamp-Lieb inequality [41] and its reverse [20], a fact that was discovered and used by K. Ball in the case of John's representation of the identity. For more information on this very important connection, see the article [ 18] in this collection.
(a) John's position. A classical result of E John [94] states that d(X, g~) <, ~ for every n-dimensional normed space X. This estimate is a by-product of the study of the following extremal problem: Let K be a body in R n. Maximize ]detu] over all u "g.~ --+ XK with ]lull - 1.
Euclidean structure in finite dimensional normed spaces
719
If u0 is a solution of this problem, then uoDn is the ellipsoid of maximal volume which is inscribed in K. Existence and uniqueness of such an ellipsoid are easy to check. An equivalent formulation of the problem is the following: Let K be a body in R n. Minimize tions u.
I l u ' ~ ~ XK II over all volume preserving transforma-
We assume that the identity map I is a solution of this problem, and normalize so that
II/ C
II -
- min{ll u "C
I1" Idet
u I --
1 }.
(1)
This means that the Euclidean unit ball Dn is the maximal volume ellipsoid of K. We shall use a simple variational argument [68] to give necessary conditions on K" THEOREM 2.3.1. Let Dn be the maximal volume ellipsoid o f K. Then, f o r every T E L ( N n , R n) we can find a c o n t a c t p o i n t x o f K and Dn (i.e., I x l - Ilxll- 1) such that tr T (x, T x ) ) ~ .
(2)
PROOF. We may assume that K is smooth enough. Let S E L ( R n , It{n). We first claim that tr S IISx II/>
(3)
n
for some contact point x of K and D~. Let e > 0 be small enough. From (1) we have tr S III + ~SII ~ [ d e f t / + sS)] l / n = 1 -Jr-~-- - } - 0 @ 2 ) .
(4)
n
Let xe E S n-1 be such that IIx~ -+- eax~ll -- III -+- call. Since D~ ___ K, we have IIx~ll ~ 1. Then, the triangle inequality for II- II shows that
IlSx~ II/>
tr S n
+ O(e).
(5)
We can find x E S n-1 and a sequence em ---> 0 such that xe m --> x. By (5) we obviously have IISx II/> ~.t~s Also, Ilxll - lim IIx~m + emSx~m II - IIIII - 1. This proves (3). Now, let T E L ( R n, R n) and write S - I + e T , e > 0. We can find xe such that IIx~ II Ixe I - 1 and
IIx~ + e Txe II/>
tr(I + eT) n
tr T -- 1 + e ~ . n
(6)
A.A. Giannopoulos and V.D. Milman
720
Since IIx~ + eTx~ll -- 1 + e(Vllx, II, Txe) -+- O(62), we obtain (Vllx~ II, Txe) ) ~ + O(e). Choosing again em -+ 0 such that xe m --->x 9 S n- 1, we readily see that x is a contact point of K and Dn, and trT
(7)
But, V Ilx II is the point on the boundary of K ~ at which the outer unit normal is parallel to x (see [177, pp. 44]). Since x is a contact point of K and Dn, we must have V llx II - x. This proves the theorem. D As a consequence of T h e o r e m 2.3.1 we get John's upper bound for d (X, s THEOREM 2.3.2. Let X be an n-dimensional normed space. Then, d(X, s
~ ~/-n.
PROOF. By the definition of the B a n a c h - M a z u r distance we may clearly assume that the unit ball K of X satisfies the assumptions of T h e o r e m 2.3.1. In particular, ]ix ]l ~< Ix] for every x 9 R n . Let x 9 ]~n and consider the map T y = (y, x ) x . T h e o r e m 2.3.1 gives us a contact point z of K and Dn such that
(z, Tz) ~>
tr T
=
H
Ix 12
.
(8)
/7
On the other hand, (z, Tz) = {z,x) 2 ~ Ilzll 2, IIx II2 = Ilx II2 ,
(9)
since one can check that Ilzll, -- 1. Therefore, Ilxll ~ Ixl ~ ~/-~llx II. This shows that Dn K C_ ~/~Dn. D REMARK. The estimate given by John's theorem is sharp. If that d (X, s - v/ft.
X =
s or s
one can check
T h e o r e m 2.3.1 gives very precise information on the distribution of contact points of K and Dn on the sphere S n-1 , which can be put in a quantitative form: THEOREM 2.3.3 (Dvoretzky-Rogers lemma [54]). Let Dn be the maximal volume ellipsoid o f K. Then, there exist pairwise orthogonal vectors Yl . . . . . Yn in IRn such that
n-i
+ 1)1/2 /,/
]]Jill ~ l y i l - 1,
i -- 1 . . . . . n.
(10)
Euclidean structure in finite dimensional normed spaces
721
PROOF. We define the yi's inductively. The first vector Yl can be any contact point of K and Dn. A s s u m e that yl . . . . . yi-1 have been defined. Let Fi --span{yl . . . . . Yi-| }. Then, t r ( P F l ) -- n -- i + 1 and using T h e o r e m 2.3.1 we can find a contact point xi for which
n-i+1 ~ - - . n
[PFixXi[ 2 -- (Xi, PF.•
(11)
We set yi - PF~Xi/IPF~Xil. Then,
<xi, yi}
1 - - lyil ~ Ily/ll-
Ily/ll 9 Ix/ll,/> (n--i+l) 1/2
= [PFi•
>/
9
n
D
(12)
Finally, a separation argument and T h e o r e m 2.3.1 give us John's representation of the identity: THEOREM 2.3.4. Let Dn be the maximal volume ellipsoid of K. There exist contactpoints X l . . . . . Xm of K and Dn, and positive real numbers )~1. . . . . )~m such that m
I -- Z )~iXi | Xi. i=1 PROOF. Consider the convex hull C of all operators x @ x, where x is a contact point of K and Dn. We need to prove that I / n E C. If this is not the case, there exists T 6 L (R n , R/7) such that
(13)
(T, I / n ) > (x @ x, T)
for every contact point x. But, (T, 1/n) -- tr__TTand (x | x T) -- (x Tx) This would contradict T h e o r e m 2.3.1. D DEFINITION. A Borel m e a s u r e / z on S n-1 is called isotropic if
is
(X, 0} 2 d/z(x) -- # ( s n - 1 ) n-I
for every 0
n
6 S n-1 .
John's representation of the identity implies that m
Z
i=1
)~i (Xi, 0} 2 -- 1
(14)
722
A.A. Giannopoulos and V.D. Milman
for every 0
6 S n-1 .
This means that if we consider the measure v on S n-j which gives
m a s s )~i at the point xi, i = 1 . . . . . m, then v is isotropic. In this sense, John's position is
an isotropic position. One can actually prove that the existence of an isotropic measure supported by the contact points of K and Dn characterizes John's position in the following sense (see [ 16,68]):
"Assume that Dn is contained in the body K. Then, Dn is the maximal volume ellipsoid of K if and only if there exists an isotropic measure v supported by the contact points of K and Dn." NOTE. The argument given for the proof of Theorem 2.3.1 can be applied in a more general context: If K and L are (not necessarily symmetric) convex bodies in R n, we say that L is of maximal volume in K if L c_ K and, for every w E ~n and T E SLn, the affine image w + T (L) of L is not contained in the interior of K. Then, one has a description of this maximal volume position, which generalizes John's representation of the identity" THEOREM 2.3.5. Let L be of maximal volume in K. For every z E int(L), we can find contactpoints Vl . . . . . Vm of K - z and L - z, contactpoints ul . . . . . Um of (K - z) ~ and (L - z) ~ andpositive reals )~l . . . . . )~m, such that ~-~ ~.juj - - 0 , (uj, vj) -- 1, and m
I -- Z
~,jUj (~ Vj.
j=l
This was observed by Milman in the symmetric case with z - 0 (see [195, Theorem 14.5]). For the extension to the nonsymmetric case see [70], where it is also shown that under mild conditions on K and L there exists an optimal choice of the "center" z so that, setting z = 0, we simultaneously have y~ ) ~ j U j - - ~ ~ , j l ) j - - 0 in the statement above.
(b) Isotropic position - Hyperplane conjecture. A notion coming from classical mechanics is that of the Binet ellipsoid of a body K (actually, of any compact set with positive Lebesgue measure). The norm of this ellipsoid E B ( K ) is defined by IIx II2Es~K~ = Igl 1 f g I<x, y)l 2 dy.
(15)
The Legendre ellipsoid EL (K) of K is defined by
fE
L(K)
(X, y)2 dy -- fK (x, y)2 dy
(16)
for every x 6 R n, and satisfies (see [147])
Es(K)
-
-
(n + 2)I/2IEL(K)1-1 (EL(K)) ~
(17)
Euclidean structure in finite dimensional normed spaces
723
That is, EL (K) has the same moments of inertia as K with respect to the axes. A body K is said to be in isotropic position if [KI = 1 and its Legendre ellipsoid EL (K) (equivalently, its Binet ellipsoid E s ( K ) ) is homothetical to Dn. This means that there exists a constant L x such that
fg(y,
O)2dy--L2
(18)
for every 0 6 S n- 1 (K has the same moment of inertia in every direction 0). It is not hard to see that every body K has a position u K which is isotropic. Moreover, this position is uniquely determined up to an orthogonal transformation. Therefore, LK is an affine invariant which is called the isotropic constant of K. An alternative way to see this isotropic position in the spirit of our present discussion is to consider the following minimization problem: Let K be a body in R n. Minimize f , i( Ix 12dx over all volume preserving transformations u. Then, we have the following theorem [147]: THEOREM 2.3.6. Let K be a body in R n with IKI = 1. The identity map minimizes fuK Ix 12dx over all volume preserving transformations u if and only if K is isotropic.
Moreover, this isotropic position is unique up to orthogonal transformations. PROOF. We shall use the same variational argument as for John's position. Let T L(R n, R n) and e > 0 be small enough. Then, u = (I + eT)/[det(I + eT)] 1In is volume preserving, and since f , x Ix I2 dx ~> fK Ix I2 dx we get
lx + ETxl 2 dx >~ [det(I + sT)] 2In fK Ixl2 dx.
(19)
But, Ix + eTxl 2 - Ixl 2 + 2e(x, Tx) + 0(82) and [det(l + sT)] 2/n -- 1 + 2e t_~_ _+_0(82). Therefore, (19) implies
fK (x, Tx) dx >~ trT fK Ix 12dx.
(20)
n
By symmetry we see that
fK (x, Tx) dx -- trT fK Ix 12dx
(21)
n
for every T E L (R n , R n). This is equivalent to
fK
(X, O)2 d x
1 n
fK IxlZdx'
0 E S n-1
(22)
A.A. Giannopoulos and V.D. Milman
724
Conversely, if K is isotropic and if T is any volume preserving transformation, then
fTK lXl2 dx -- fK lTXl2 dX -- fK(X' T*Tx) dx ---- tr(T*T) fK lXl2 dx >~fK lXl2
(23)
which shows that K solves our minimization problem. We can have equality in (23) if and only if T ~ O (n). D It is easily proved that L/~ ~> Lon ~> c > 0 for every body K in ~;~n, where c > 0 is an absolute constant. An important open question having its origin in [26] is the following: PROBLEM. Does there exist an absolute constant C > 0 such that LK ~< C for every body K? A simple argument based on John's theorem shows that L/~ ~< c~/-n for every body K. Uniform boundedness of L/~ is known for some classes of bodies: unit balls of spaces with a 1-unconditional basis, zonoids and their polars, etc. For partial answers to the question, see [ 13,47,48,98,99,105,147]. The best known general upper estimate is due to Bourgain [28]: L x <~ c~Cfflog n for every body K in R n . In the Appendix we give a brief presentation of Bourgain's result. The problem we have just stated has many equivalent reformulations, which are deeply connected with problems from classical convexity. For a detailed discussion, see [ 147]. An interesting property of the isotropic position is that if K is isotropic then all central sections K A 0 • 0 ~ S n - 1 are equivalent up to an absolute constant. This comes from the fact that
fI~ (x, 0) 2 dx
= L 2 "~ [K A10-1-[2,
0 6 s n _ 1,
(24)
a consequence of the log-concavity of #K. This was first observed in [91 ]. Then, uniform boundedness of L/~ is equivalent to the statement that an isotropic body has all its (n - 1)dimensional central sections bounded below by an absolute constant. This is equivalent to the following: HYPERPLANE CONJECTURE. Is it true that a body K of volume 1 must have an (n - 1)dimensional central section with volume bounded below by an absolute constant?
(c) Minimal surface position.
Let K be a convex body in ~;~n with normalized volume IK[ = 1. We now consider the following minimization problem:
Find the minimum of
O(u K) over all volume preserving transformations u.
This minimum is attained for some u0 and will be denoted by Or (the minimal surface invariant of K). We say that K has minimal surface if O(K) - OKIK[ (n-1)/n.
Euclidean structure in finite dimensional normed spaces
725
Recall that the area measure CrK of K is defined on S "-I and corresponds to the usual surface measure on K via the Gauss map" For every Borel A c_ S n- 1, we have
CrK(A) -- v({x E bd(K)" the outer normal to K at x is in A}),
(25)
where v is the (n - 1)-dimensional surface measure on K. We obviously have O(K) =
CYK(Sn-1). A characterization of the minimal surface position through the area measure was given by Petty [159]: THEOREM 2.3.7. Let K be a convex body in R n with [ K I - 1. Then, O(K) - - O K if and only if aK is isotropic. Moreover, this minimal surface position is unique up to orthogonal transformations. The proof makes use of the same variational argument. The basic observation is that if u is any volume preserving transformation, then
O((u-')*K)-
fs,, l luxl K(dX .
(26)
Ball [ 15] has proved that the minimal surface invariant OK is maximal when K is a cube in the symmetric case, and when K is a simplex in the general case. It follows that OK ~< 2n for every body K in R n. For applications of the minimal surface position to the study of hyperplane projections of convex bodies, see [69] (also, [14] for an approach through the notion of volume ratio).
(d) Minimal mean width position.
Let K be a convex body in R n . The mean width of K
is defined by
w(K) -- 2 ji,,_, hK(u)cr(du),
(27)
where hK(x) = maxyeK(X, y) is the support function of K. We say that K has minimal mean width if w ( T K ) ~ w(K) for every volume preserving linear transformation T of R n . Our standard variational argument gives the following characterization of the minimal mean width position: PROPOSITION 2.3.8. A smooth body K in R n has minimal mean width if and only if fs n-l
( V h K Tu) ( u cr(du) ) , -
trT w(K) n 2
(28)
for every linear transformation T. Moreover, this minimal mean width position is uniquely determined up to orthogonal transformations.
726
A.A. Giannopoulos and V.D. Milman
Consider the measure wK o n
S n-
1 with density h/( with respect to a. If we define
Ix(O) = fsn-1 (VhK (u), O)(u, O) cr (du),
0 ~ S n- 1,
(29)
an application of Green's formula shows that
w(K)
+ IK(O) -- (n + 1) fgn_ 1 h g ( u ) ( u , 0) 2 o"(du).
(30)
Combining this identity with Proposition 2.3.8, we obtain an isotropic characterization of the minimal mean width position (see [68])" THEOREM 2.3.9. A convex body K in IRn has minimal mean width if and only if wK is isotropic. Moreover, the position is uniquely determined up to orthogonal transformations. NOTE. It is natural to ask for an upper bound for the minimal width parameter, if we restrict ourselves to bodies of fixed volume. It is known that every body K has a linear image K with IK I-- IDn I such that
(31)
w(F, ) <<.c l o g ( 2 d ( X K , g~)) ~< clog(2n),
where c > 0 is an absolute constant. This statement follows from an inequality of Pisier [ 161 ] after work of Lewis [ 111 ], Figiel and Tomczak-Jaegermann [60], and plays a central role in the theory. We shall use the minimal mean width position and come back to the estimate (31) in Section 5.
3. Background from classical convexity 3.1. Steiner's formula and Urysohn's inequality 3.1.1. Let/Cn denote the set of all nonempty, compact convex subsets of R n. We may view/Cn as a convex cone under Minkowski addition and multiplication by nonnegative real numbers. Minkowski's theorem (and the definition of the mixed volumes) asserts that if K1 . . . . . Km ~ 1Cn, m ~ N, then the volume of tl K1 + . . . + tmKm is a homogeneous polynomial of degree n in ti >/0 (see [43,177]). That is, ItlKl + ' "
+ tmKml -
Zin~m V(Kil . . . . .
Kin)til ""tin,
1~
where the coefficients V (Kil . . . . . Kin) are chosen to be invariant under permutations of their arguments. The coefficient V (K1 . . . . . Kn) is called the mixed volume of K1 . . . . . Kn.
Euclidean structure in finite dimensional normed spaces
727
Steiner's formula, which was already considered in 1840, may be seen as a special case of Minkowski's theorem. The volume of K + t Dn, t > 0, can be expanded as a polynomial in t:
]g-~-tOn[--~(~)wi(g) ti,
(1)
i=0
where Wi (K) = V (K; n - i, D,; i) is the i th Quermassintegral of K. It is easy to see that the surface area of K is given by
O(K) = n W I ( K ) .
(2)
Kubota's integral formula
ID,I
Wi (K) -- ID~-i I,-i
fG
IP~KI,-i dvn,n-i(~) ....
(3)
i
applied for i = n - 1 shows that IDnl
Wn-l(K) = ~w(K). 2
(4)
3.1.2. The Alexandrov-Fenchel inequalities constitute a far reaching generalization of the Brunn-Minkowski inequality and its consequences: If K, L, K3 . . . . . Kn ~ 1Cn, then
V ( K , L, K3 . . . . . Kn) 2 >~ V ( K , K, K3 . . . . . K n ) V ( L , L, K3 . . . . . Kn).
(5)
The proof is due to Alexandrov [6,7] (Fenchel sketched an alternative proof, see [58]). From (5) one can recover the Brunn-Minkowski inequality as well as the following generalization for the quermassintegrals:
W i ( K -+- L) l/(n-i) > / W i ( K ) 1~(n-i) -k- Wi(L) 1~(n-i),
i --0 ..... n-
1
(6)
for any pair of convex bodies in R n . If we take L = t Dn, t > 0, then Steiner's formula and the Brunn-Minkowski inequality give
~(~) wi(g) t i =
i=0
IDnl
IOnl
\ \ [--~[
----~(~)( Igli--0]---~nl)(n-i)/n t i
+ t (7)
A.A. Giannopoulosand V.D.Milman
728
for every t > 0. Since the first and the last term are equal on both sides of this inequality, we must have
W( ~n (- K ~ ) /~n(lIDKnI l)
(8)
which is the isoperimetric inequality for convex bodies, and
w(K)__2Wn-I(K)>~2(IKI)I/n IDnl
-~
Urysohn's inequality. Both inequalities
which is
,
(9)
inequalities are special cases of the set of Alexandrov
(Wi(K)) 1~(n-i) (Wj(K)) 1~(n-j) [Dn[ 3.1.3.
>/
[Dn[
, n > i > j >~O.
(10)
Let K be a body in R n . We define
M*(K)
= f s ,,-, I l x l l , o ( d x ) - -w(K) -~.
(11)
The Blaschke-Santal6 inequality asserts that the volume product [KIIK~ is maximized over all symmetric convex bodies in R n exactly when K is an ellipsoid:
IKIIK~ ~< IDnl 2.
(12)
A proof of this fact via Steiner symmetrization was given in [ 12] (see also [ 130,131 ] where the nonsymmetric case is treated). H61der's inequality and polar integration show that
1
M*(K) <~
(L n-1
Ilxll."
=
~
.
(13)
Combining with (12) and applying (13) for K instead of K ~ we obtain
1
(Ig[)
M(K) <~\ ]---~n[
'/~
<~M*(K),
(14)
that is, Urysohn's inequality.
Ui E
3.1.4. A third proof of Urysohn's inequality can be given as follows: Let O(n), i = 1 . . . . . m, and 0ti > 0 w i t h ff-~im__~1 0ti - - 1. It is easily checked that M* (Zi=lm otiui(g)) -M* (K). It follows that
M*(f ~
(n)
u(K)d#(u))=
M*(K).
(15)
Euclidean structure in finite dimensional normed spaces
729
But, T - f o ( n ) u ( K ) d # ( u ) is a ball of radius (ITI/IDnl) l/n, and the B r u n n - M i n k o w s k i inequality implies that [T[ ~> IK[. Therefore,
1/n (16) 3.1.5. For any (n - 1)-tuple C - - K 1 . . . . . Kn-1 E K~n, the Riesz representation theorem shows the existence of a Borel measure S(C, .) on the unit sphere S n-1 such that
V ( L , K1 . . . . . K n - 1 ) - - n
hL(u) d S ( C , u )
(17)
n-I
for every L 6 /Cn. If K E K~n, the j t h area measure of K is defined by S j ( K , .) = S ( K ; j, Dn; n - j - 1, .), j = 0, 1 . . . . . n - 1. It follows that the quermassintegrals W i ( K ) can be written in the form
Wi(K)-
-
hK(u) d S n - i - l ( K , u ) ,
n
i - - O , 1. . . . . n -
1,
(18)
n- 1
or, alternatively,
W i ( K ) -- -
dSn-i(K,u),
n
i = 1 . . . . . n.
(19)
n-I
If we assume that hi( is twice continuously differentiable, then S j ( K , .) has a continuous density sj (K, u), the j t h elementary symmetric function of the eigenvalues of the Hessian of hi( a t u . In the spirit of 2.3, we say that a body K minimizes Wi if W i ( K ) <<,W i ( T K ) for every volume preserving linear transformation T of R n . The cases i = 1 and i = n - 1 correspond to the minimal surface area and minimal mean width respectively. For every i = 1 . . . . . n - 1 one can prove that, if K minimizes Wi then S n - i ( K , ") is isotropic (see [68], where other necessary isotropic conditions are also given).
3.2. Geometric inequalities of "hyperbolic" type The Alexandrov-Fenchel inequalities are the most advanced representatives of a series of very important inequalities. They should perhaps be called "hyperbolic" inequalities in contrast to the more often used in analysis "elliptic" inequalities: Cauchy-Schwarz, H61der, and their consequences (various triangle inequalities). A consequence of "hyperbolic" inequalities is concavity of some important quantities. 3.2.1. Let us start this short review by recalling some old and classical, but not well remembered, inequalities due to Newton. Let x l . . . . . Xn be real numbers. We define the elementary symmetric functions eo(xl . . . . . xn) = 1, and
ei(xl . . . . . Xn)
-
-
Z l~jl<'"<ji<~n
XjlXj2 "" "Xji ,
1 <<,i <<,n.
(1)
730
A.A. Giannopoulos and V.D. Milman
In particular, e l ( x 1 . . . . . Xn) -- ~-~in___lXi, en(Xl . . . . . Xn) -- 1-Iin__l Xi. We then consider the normalized functions 1 E i ( x l . . . . . Xn) -- 7h-~ei(xl . . . . . Xn).
(2)
t i) Newton proved that, for k -- 1 . . . . . n - 1,
(3)
Ek2 (x 1. . . . . Xn) ~ Ek-1 (Xl , . . ., Xn) Ek+ 1(x I ,. .., Xn),
with equality if and only if all the Xi'S a r e equal. An immediate corollary of (3), observed by Newton's student Maclaurin, is the string of inequalities E1 (xl . . . . . Xn) > / E 21/2 (Xl . . . . . Xn) > / " " > / ~Fn1 / n ( X l , . . . , Xn),
(4)
which holds true for any n-tuple (x 1. . . . . Xn) of positive reals. Note the similarity between (3), (4) and the Alexandrov-Fenchel and Alexandrov inequalities 3.1.2(5) and (10) respectively. To prove (3) we consider the polynomial
P ( x ) --
(x - xi) -i--1
(-1) j
E j ( x l . . . . . Xn)X n - j ,
(5)
(--1) j
Ej(xl ..... Xn)tn-Jr j
(6)
j--0
or in homogeneous form,
Q(t, r) - r n P
= j=O
Since P has only real roots, the same is true for the derivatives of P (with respect to t or r ) of any order. If we differentiate (6) (n - k - 1)-times with respect to t and then (k - 1)-times with respect to r, we obtain the polynomial n!
2 Ek-1 (Xl,.
Xn)t 2 -- n ! E k ( x l . . . .
n!
Xn)t• -k- -~ Ek+l (Xl . . . .
Xn)r 2,
(7)
which has two real roots for fixed r = 1. This is exactly Newton's inequality (3). We refer to [ 167] for a very nice different proof and generalizations. 3.2.2. Let us now turn to a multidimensional, but still numerical, analogue of Newton's inequalities. Consider the space Sn of real symmetric n • n matrices. We polarize the function A --+ det A to obtain the symmetric multilinear form
1 D(A1 . . . . . A~) -- S;
~ eE{O, 1}n
(-1
)n+Eei
det
(~-~
eiAi
)
(8)
Euclidean structure in finite dimensional normed spaces
731
where Ai E Sn. Then, if t| . . . . . tm > 0 and A I . . . . . Am E Sn, the determinant of tl A1 + 9.. + tm Am is a homogeneous polynomial of degree n in ti" det(tl A1 + . . . + tm Am) --
n!D(Ail . . . . . Ain)til ""tin.
(9)
l<~il<~...<~i,,<~m The coefficient D ( A | . . . . . An) is called the mixed discriminant of A1 . . . . . An. The fact that the polynomial P (t) = det(A + t I) has only real roots for any A ~ Sn plays the central role in the proof of a number of very interesting inequalities connecting mixed discriminants, which are quite similar to Newton's inequalities. They were first discovered by Alexandrov [7] in one of his approaches to what is now called Alexandrov-Fenchel inequalities. Today, they are part of a more general theory (see, e.g., [93]). We mention some of them: If Ai, i = 1 . . . . . n, are positive, then n
(10)
D ( A 1 , A2 . . . . . An) >~ F I [ d e t A ] | / n
i=1 Also, the following concavity principle (reverse triangle inequality) is true: The function [detA] l/n is concave in the positive cone of Sn. This is in fact easy to demonstrate directly. We want to show that, if A|, A2 are positive then [det(A| + A2)] 1/n >/[detA|] |/n + [detA2] 1/n.
(11)
We may bring two positive matrices to diagonal form without changing their determinants. Then, we should show that for )~i, #i > 0,
(n
)in (H)ljn (H)
1 7 ()~i + #i ) i=1
|/n
~
)~i i=1
_qt_
#i
(12)
i=1
which is a consequence of the arithmetic-geometric means inequality. 3.2.3. We now return to convex sets. The results of Sections 3.2.1 and 3.2.2 have their analogues in this setting, but the parallel results for mixed volumes are much more difficult and look unrelated. Even the fact that the volume of tl K1 + . . . § tm Km is a homogeneous polynomial in ti ~ 0 is a nontrivial statement, while the parallel result for determinants follows by definition. To see the connection between the two theories we follow [5]. Consider n fixed convex open bounded bodies Ki with normalized volume IKi] = 1. As in Section 2.2(a), consider the Brenier maps
~i "( I~n, Yn) ~ gi,
(13)
where Yn is the standard Gauss|an probability density on R n. We have ~ i -- V j~, where fi are convex functions on IRn. By Caffarelli's regularity result, all the 7ti's are smooth
A.A. Giannopoulos and V.D. Milman
732
maps. Then, Fact 2 from 2.2(a) shows that the image of (R n, Yn) by ~ ti ~i is the interior of y~ ti Ki. Since each Oi is a measure preserving map, we have d e t ( 02fi ) OxkOxl
(x) -- y~(x),
i = 1 . . . . . n.
(14)
It follows that
• i=1
ti Ki
-- f R n d e t ( ~ t i ( i=l
--
t i , " " tin
02fi ) ) dx OxkOxl D
-~xk O~ . . . . .
02fin(X)) -O-~k-O~
dx .
(15)
il ..... in--1
In particular, we recover Minkowski's theorem on polynomiality of ] y~ ti Ki 1, and see the connection between the mixed discriminants D(Hessfil . . . . . Hess fin) and the mixed volumes
V (Kil . . . . . Kin ) -- f~n D (Hess j~, (x) . . . . . Hess fin (x)) dx.
(16)
The Alexandrov-Fenchel inequalities do not follow from the corresponding mixed discriminant inequalities, but the deep connection between the two theories is obvious. Also, some particular cases are indeed simple consequences. For example, in [5] it is proved (as a consequence of (16)) that
V(K1 . . . . . Kn) ~ ~-I [Kill/n
(17)
i=1
3.3. Continuous valuations on compact convex sets
(a) Polynomial valuations. We denote by K~n the set of all nonempty compact convex subsets of ~n and write L for a finite dimensional vector space over IR or C. A function go : K2n --+ L is called a valuation, if go(K 1 U K2) -+- go(K1 A K2) = go(K1) + go(K2) whenever K1, K2 E K~n are such that K1 U K2 E ~ n . We shall consider only continuous valuations: valuations which are continuous with respect to the Hausdorff metric. The notion of valuation may be viewed as a generalization of the notion of measure defined only on the class of compact convex sets. Mixed volumes provide a first important example of valuations. A valuation go: K~n ~ L is called polynomial of degree at most I if go(K + x) is a polynomial in x of degree at most I for every K E K~n. The following theorem of Khovanskii and Pukhlikov [ 102] generalizes Minkowski's theorem on mixed volumes (see also [ 127,2]):
Euclidean structure in finite dimensional normed spaces
733
THEOREM 3.3.1. Let qg" ]~n ---> L be a continuous valuation, which is polynomial of degree atmost I. Then, if K1 . . . . . Km ~ 1Cn, cp(tlK1 4 - ' " 4- tmKm) is apolynomial in tj >/0 of degree at most n 4- I. Let K -- (K1 . . . . . Ks) be an s-tuple of compact convex sets in ~n, and F" ]1~n ~ C be a continuous function. Alesker studied the Minkowski operator M E. which maps F to M~. F'R~_ --+ C with
(M~F)(/Z1 . . . . . Xs)
--
f~_~4.<s,~.iKi
F(x)dx.
Let ,A(C n) be the Frechet space of entire functions of n variables and C r (R n) be the Frechet space of r-times differentiable functions on R n, with the topology of uniform convergence on compact sets. The following facts are established in [1 ]: (i) If F 6 .A(Cn), then M y F has a unique extension to an entire function on C S, and the operator ME, ",A(C n) --> A ( C '~) is continuous. It follows that if F is a polynomial of degree d then M~. F is a polynomial of degree at most d 4- n. (ii) If F 6 C r (Rn), then M~. F 6 C~ (R~_), and M E. is a continuous operator. Moreover, continuity of the map K w-> M E. with respect to the Hausdorff metric is established.
(b) Translation invariant valuations. A valuation of degree 0 is simply translation invariant. If ~o(uK) = ~p(K) for every K 6/Cn and every u E SO(n), we say that ~0 is SO(n)invariant. Hadwiger [89] characterized the translation and SO(n) invariant valuations as follows (see also [ 101 ] for a simpler proof): THEOREM 3.3.2. A valuation ~p is translation and SO(n)-invariant if and only if there
exist constants ci, i -- 0 . . . . . n, such that
~p(K) -- ~ ci Wi (K) i=0
(1)
for every K c 1Cn. After Hadwiger's classical result, two natural questions arise: to characterize translation invariant valuations without any assumption on rotations, and to characterize O(n) or SO(n) invariant valuations without any assumption on translations. Both questions are of obvious interest in translative integral geometry and in the asymptotic theory of finite dimensional normed spaces respectively (consider, for example, the valuation cp(K) -- fK Ixl2 dx which was discussed in 2.3(b)). It is a conjecture of McMullen [128] that every continuous translation invariant valuation can be approximated (in a certain sense) by linear combinations of mixed volumes. This is known to be true in dimension n ~< 3. In [127,128] it is proved that every transn lation invariant valuation ~o can be uniquely expressed as a sum ~0 - ~-~i=0 ~0i, where ~i
A.A. Giannopoulos and V.D. Milman
734
are translation invariant continuous valuations satisfying (,/9i (t K) = t i ~o(K) (homogeneous of degree i). Moreover, in the case L = R, homogeneous valuations ~oi as above can be described in some cases: ~00 is always a constant, ~On is always a multiple of volume, ~0n-1 is always of the form
q)n-1 (K) = fs.-' f (u) dSn-1 (K, u),
(2)
where f : S n-1 --+ R is a continuous function (which can be chosen to be orthogonal to every linear functional, and then it is uniquely determined). Under the additional assumption that ~o is simple (~0(K) = 0 if dim K < n), a recent theorem of Schneider [ 178] completely describes ~0: THEOREM 3.3.3. Every simple, continuous translation invariant valuation q) : ]~n ~ I[{
has the form q)(K) -- clKI § Ln-1 f (u) dSn-1 (K, u),
(3)
where f : S n- 1 ~ • is a continuous odd function. REMARK. McMullen's conjecture was recently proved by Alesker [3] in dimension n = 4.
Added in proofs.
Even more recently, Alesker [4] gave a description of translation invariant valuations on convex sets, which in particular confirms McMullen's conjecture in all dimensions.
(c) Rotation invariant valuations. Alesker [2] has recently obtained a characterization of O (n) (respectively SO(n)) invariant continuous valuations. The first main point is that every such valuation can be approximated uniformly on the compact subsets of K~n by continuous polynomial O(n) (or SO(n)) invariant valuations. Then, one can describe polynomial rotation invariant valuations in a concrete way. To this end, let us introduce some specific examples of such valuations. We write v for the (n - 1)-dimensional surface measure on K and n(x) for the outer normal at bd(K) (this is uniquely determined v-almost everywhere). If p, q are nonnegative integers, we consider a valuation ~p,q:]~n --~ ]I~ with
~p,q(K) -- fb
(x,n(x))Plxl 2q dr(x).
(4)
d(K)
All ~p,q are continuous, polynomial of degree at most p + 2q § n, and O(n)-invariant. Theorem 3.3.1 shows that, for every K 9 1On, ~p,q(K § eDn) is a polynomial in e ~> 0, therefore it can be written in the form
~p,q(K§
p+2q+n Z ~r(i)p,q(K)ei" i=0
(5)
Euclidean structure in finite dimensional normed spaces
735
~lr(i) are continuous, polynomial and O(n)-invariant. These particular valuations sufAll wp,q fice for a description of all rotation invariant polynomial valuations [2]:
THEOREM 3.3.4. I f n >~ 3, then every SO(n)-invariant continuous polynomial valuation (i) ~o" lCn --+ R is a linear combination o f the wp,q. Since ,tr (i) are O(n)-invariant, Theorem 3.3.4 describes O(n)-invariant valuations as wp,q well. The case n = 2 is also completely described in [2] (and the same statements hold true if • is replaced by C).
4. Dvoretzky's theorem and concentration of measure
4.1. Introduction A version of the Dvoretzky-Rogers lemma [54] asserts that for every body K whose maximal volume ellipsoid is Dn, there exist k ~_ x/-ff and a k-dimensional subspace Ek of ]1~n such that Dk ___K N Ek ___2 Qk, where Dk denotes the Euclidean ball in Ek and Qk the unit cube in Ek (for an appropriately chosen coordinate system). Inspired by this, Grothendieck asked whether Qk can be replaced by Dk in this statement. He did not specify what the dependence of k on n might be, asking just that k should increase to infinity with n. A short time after, Dvoretzky [52,53] proved Grothendieck's conjecture: THEOREM 4.1.1. Let e > 0 and k be a positive integer. There exists N = N ( k , e) with the following property: Whenever X is a normed space o f dimension n ~ N we can find a k-dimensional subspace Ek o f X with d(Ek, s <~ 1 + e. Geometrically speaking, every high-dimensional body has central sections of high dimension which are almost ellipsoidal. The dependence of N ( k , e) on k and e became a very important question, and Dvoretzky's theorem took a much more precise quantitative form: THEOREM 4.1.2. Let X be an n-dimensional normed space and e > O. There exist an integer k ~ ce 2 logn and a k-dimensional subspace Ek o f X which satisfies d ( E k , s 2) l+e. This means that Theorem 4.1.1 holds true with N ( k , e) = e x p ( c e - 2 k ) . Dvoretzky's original proof gave an estimate N ( k , e) = e x p ( c e - Z k 2 logk). Later, Milman [132] established the estimate N ( k , e) = exp(ce-2l log elk) with a different approach. The logarithmic in e term was removed by Gordon [75], and then by Schechtman [174]. Other proofs and extensions of Dvoretzky's theorem in different directions were given in [59,180,107] (see also the surveys [ 112,114,143]). The logarithmic dependence of k on n is best possible for small values of e. One can see this by analyzing the example of s Every k-dimensional central section of Qn is a polytope with at most 2n facets. If we assume that we can find a subspace Ek of s with
A.A. Giannopoulos and V.D. Milman
736
d(Ek, g~) ~< 1 + e, then there exists a polytope Pk in IRk with m ~< 2n facets satisfying Dk ___ Pk g (1 + e)Dk. The hyperplanes supporting the facets of Pk create m spherical caps J1 . . . . . Jm on (1 -+- 6)8 k-1 such that (1 + e)S k-1 g LJim=lJi. O n the other hand, since Dk g Pk, if we assume that e is small, then each Ji has angular radius of the order of v/-e-. An elementary computation shows that the normalized measure of such a cap does not e x c e e d (c6) (k-1)/2. Therefore, we must have 2n f> (C8) -(k-1)/2 w h i c h shows that
k <~clogn/log(1/e).
(1)
The same argument shows that if P is a symmetric polytope and f (P) is the number of its facets, then k <<.c(e) log f (P). The right dependence of N(k, e) on e for a fixed (even small) positive integer k is not clear. It seems reasonable that ~ is the worst case and that the computation we have just made gives the correct order: Q U E S T I O N 4.1.3. Can we take
N(k, e) -- c(k)8 -(k-l)~2 in Theorem 4.1.1?
Using ideas from the theory of irregularities of distribution, Bourgain and Lindenstrauss [31] have shown that the choice N(k, e) = c(k)e-(k-1)/2l log el is possible for spaces X with a 1-symmetric basis. There are numerous connections of this question with other branches of mathematics (algebraic topology, number theory, harmonic analysis). For instance, an affirmative answer to Question 4.1.3 would be a consequence of the following hypothesis of Knaster: Let f : S k- 1 __+ R be a continuous function and x l . . . . . xk be points on S k-1 . Does there exist a rotation u such that f is constant on the set {uxi: i <~k}? This hypothesis has been settled only in special cases (see [138] for a discussion of this and other problems related to Question 4.1.3). NOTE. Bourgain and Szarek [35] proved a stronger form of the Dvoretzky-Rogers lemma: If Dn is the ellipsoid of minimal volume containing K, then for every 6 6 (0, 1) one can choose xl . . . . . Xm, m >~ (1 - 8)n, among the contact points of K and D, such that for every choice of scalars (ti)i <.m, m
fQi)
t2 i=1
<~ ~ t i x i i=1
~tiXi
~ltil.
i=1
i=1
(2)
This is a Dvoretzky-Rogers lemma for arbitrary proportion of the dimension. It can also be stated as a factorization result: For any n-dimensional normed space X and any 6 6 (0, 1), one can find m ~> (1 - 6)n and two operators or" g~n __+ X, r " X --+ s such that the identity id2,~ .g~n __+ g ~ can be written as id2,~ - fl oot and II~ll II/~11 ~< 1/f(6). For an extension to the nonsymmetric case see [ 116]. Using this result, Bourgain and Szarek answered in the negative the question of uniqueness, up to a constant, of the centre of the Banach-Mazur compactum, and gave the first nontrivial estimate o(n) for the Banach-Mazur distance from an n-dimensional space X
Euclidean structure in finite dimensional normed spaces
737
to g n . It is now known [186,63] that (2) holds true with f ( 6 ) -- c3. The question about the best possible exponent of 3 in the D v o r e t z k y - R o g e r s factorization is also open. By [63,169] it must lie between 1/2 and 1. In the Appendix we give a brief account on these and other questions related to the geometry of the B a n a c h - M a z u r compactum.
4.2. Concentration of measure on the sphere and a pr oof of Dvoretzky's theorem We shall outline the approach of [ 132] to Dvoretzky's theorem. The method uses the concentration of measure on the sphere and was further developed in [61 ]. We need to introduce the average parameter
M-
M ( X K ) --
fs"-'
(1)
Ilxll a ( d x ) ,
the average on the sphere S " - l of the norm that K induces on R n . REMARKS ON M. (i) It is clear from the definition that M depends not only on the body K but also on the Euclidean structure we have chosen in R ~ If we assume that allxl ~< Ilxll ~< blxl and that a, b > 0 are the smallest constants for which this is true for all x E R ~, then we have the trivial bounds a1 ~< M ~< b. (ii) For every p > 0 we define 9
Ilxll p a ( d x )
Mp -- M p ( X K ) -n-
(2)
I
In this notation M -- M1 and as a consequence of the K a h a n e - K h i n c h i n e inequality one can check that M1 "~ M2 independently from the dimension and the norm. It can be actually shown [118] that, for every 1 ~< p <~ n,
max{, Cl /
max/2l c2 + /
(3)
where c l, c2 > 0 are absolute constants. (iii) Let g l . . . . . gn be independent standard Gaussian r a n d o m variables on some probability space S-2 and {e'1. . . . . e'n } be any orthonormal basis in R" . Integration in polar coordinates establishes the identity
2
17
Z
gi (co)el
1/2 -- v/-nM2.
(4)
i=1
Using the s y m m e t r y of the gi's and the triangle inequality for II 9 II we get
k y'~ gi (co)e I
•
i--1
i=1
gi (co)e I d~o,
(5)
A.A. Giannopoulos and V.D. Milman
738
for every 1 ~< k ~< n, and combining with the previous observations we have
M ( E k ) <~Cv/-~/kM
(6)
for every k-dimensional subspace Ek of XK. The main step for our proof of Theorem 4.1.2 will be the following [132]: THEOREM 4.2.1. Let X be an n-dimensional normed space satisfying al lxl ~ Ilxll ~ blxl. For every e ~ (0, 1) there exist k >~ce2n(M/b) 2 and a k-dimensional subspace Eg of R n such that
~Llxl l+e
~ Ilxll ~ (1 +E)Llxl,
x6E~.
The constant L appearing in the statement above is the L~vy mean (or median) of the function f ( x ) = Ilxll on a n - l . This is the unique real number L = L f for which 1 o'({x" f ( x ) > ~ L } ) ~ > ~
and
o-({x" f ( x ) ~ L } ) ~ > ~ .
1
A few observations arise directly from this statement: Assume that x ~ S n- 1 has maximal norm IIx II = b. Consider the one-dimensional subspace E1 spanned by x. We have b = M(E1) <~c~/~M, and this shows that n ( M / b ) 2 ~> c > 0 for every norm. This is of course not enough for a proof of Dvoretzky's theorem. On the other hand, recall that M >~ 1/a. By Theorem 4.2.1, every X has a subspace of dimension k >~ ce2n/(ab) 2 on which I1" II is (1 + e)-equivalent to the Euclidean norm. Since we can choose a linear transformation of K x so that ab <~d(X, s we obtain the following corollary [ 132]: COROLLARY 4.2.2. For every n-dimensional space X and every e ~ (0, 1) we can find a subspace Ek of X with dimE~ = k ~ ce2n/d2(X, s such that d(Ek, s <~ 1 + e. This already shows that spaces with small Banach-Mazur distance from s have Euclidean sections of dimension much larger than log n (even proportional to n). However, since John's theorem is sharp this observation is not enough for the general case. The proof of Theorem 4.2.1 is based on the concentration of measure on the sphere. Recall that as a consequence of the spherical isoperimetric inequality we have the following fact:
If A c_C_S n-1 and or(A) - 1, then cr(Ae)/> 1 - Cl exp(-c2e2n). This inequality explains the term "concentration of measure": However small e > 0 may be, the measure of the set outside a "strip" of width e around the boundary of any subset of the sphere of half measure is less than 2cl exp(-c2e2n), which decreases exponentially
Euclidean structure in finite dimensional normed spaces
739
fast to 0 as the dimension n grows to infinity. This surprising fact was observed and used by L6vy [ 110]" Let f be a continuous function on the sphere. By coy(.) we denote the modulus of continuity of f :
corn(t)- max{If(x)- f(Y)I"
p ( x , y) <<.t, x, y ~ s ' - ' }.
Consider the L6vy mean L f of f . It is not hard to see that
{x" f - - Lf}e -- ({x" f >~ Lf})e A ({x" f <~ L f}) . Since If(x) - L f l <~ Of(E) o n {x: f the following direct consequence:
=
Lf}e, the spherical isoperimetric inequality has
FACT 1. For every continuous function f : S n-1 --+ IR and every e > O,
s
If(x)- LmI
Cl e x p ( - c 2 e 2 n ) .
(7)
If the modulus of continuity of f behaves well, then Fact 1 implies strong concentration of the values of f around its median. Moreover, from a set of big measure on which f is almost constant we can extract a subspace of high dimension, on the sphere of which f is almost constant: FACT 2. Let f : S n- 1 ~ IR be a continuous function and 6, 0 > O. There exists a subspace F o f N n with d i m F = k ~> c62n/log(3/O) such that
If(x) - L f l ~ cof(~) 4- cof(O) f o r every x ~ S ( F )
:=
S n-1 (-'1F.
PROOF. Fix k < n (to be determined) and Fk E Gn,k. A standard argument shows that there exists a 0-net N" of S(Fk) with cardinality IArl ~< (1 + 2)k ~< exp(k log(3/0)). If x 6 N', then
.(u O(n)If(ux)-
>
Cl exp(-c262n).
(8)
Therefore, if ci IN'I e x p ( - c 2 6 2 n ) < 1 then most u E O(n) satisfy
If (ux) - L f l <~ coS(a)
(9)
for every x 6 N'. It follows that If(x) - L f l ~< COf(~) nt- Of(O) for every x ~ S(uFk). A simple computation shows that the necessary condition will be satisfied for some k ~> c~2n/log(3/O). D For the proof of Theorem 4.2.1 we are going to apply this fact to the norm f (x) = Ilx II. In this case, one can say even more (see [150, pp. 12]):
A.A. Giannopoulos and V.D. Milman
740
FACT 3. Let X -- (JRn, I1" II) and assume that Ilxll ~ blxl. For every e ~ (0, 1) there exists a subspace Ek with dimEk -- k >>,( ( c e 2 ) / l o g ( 1 / e ) ) n ( L f / b ) 2 such that
~Lflxl l+s
~ Ilxll ~ (1 + e ) L f l x l
f o r every x ~ Ek. The proof of Theorem 4.2.1 is now complete. We just have to observe that if f (x) =
Ilxll on S n-l, then L f ~ M. By Markov's inequality, cr(x" f ( x ) >>,2M) ~< 1/2 and this shows that L f <~ 2M. It can be checked that L f >~ c M as well, where c > 0 is an absolute constant [150]. It follows that we can have almost spherical sections of dimension k ~> ( ( c e 2 ) / l o g ( 1 / e ) ) n ( M / b ) 2 in Theorem 4.2.1. In order to remove the logarithmic in e term, one needs to put additional effort (see [75,174]). From Theorem 4.2.1 we may deduce Dvoretzky's theorem (Theorem 4.1.2): For every n-dimensional space X and any e 6 (0, 1) there exists a subspace Ek of X with dim Ek = k ~ ce 2 logn, such that d(Ek, g~w <~ 1 + e. PROOF. We may assume that Dn is the maximal volume ellipsoid of K x . Then, Ilxll on R n and in view of Theorem 4.2.1 we only need to show that M 2 1> c log n / n . is a consequence of the Dvoretzky-Rogers lemma: There exists an orthonormal yl . . . . . Yn in ]Rn with Ilyi II >1 ((n - i + 1)/n)l/2. In particular, Ilyi II ~> 1/2, i - 1 . . . . . From the equivalence of M1 and M2 we see that
• gi (co)yi
M ~ ~
This basis n/4.
n/4
~ gi (o))yi dw
dco~>
i=1
~
~ Ixl
i=1
i<~n/amaX [[gi(w)Yilldw >>,- -c~
c" Clog n ,
fo i<~n/4max[gi (o9)1 dw ~> ~
(10)
where we have used the fact (see, e.g., [109, pp. 79]) that if gl . . . . . gm are independent standard Gaussian random variables on s then max/<~m [gi[ ~ x/log m. D
f~
4.3. Probabilistic and global form of Dvoretzky's theorem The proof of Theorem 4.2.1, being probabilistic in nature, gives that a subspace Ek of X with dim Ek = [ce2n(M/b) 2] is (1 + e)-Euclidean with high probability. This leads to the definition of the following characteristic of X" DEFINITION. Let X be an n-dimensional normed space. We set k ( X ) to be the largest positive integer k ~< n for which
(
Prob Ek ~ Gn k: _MIx[ <<,Ilxll ~ 2Mlxl, x E Ek '
)
k
~> 1 - ~ . n+k
(1)
Euclidean structure in finite dimensional normed spaces
741
In other words, k ( X ) is the largest possible dimension k ~< n for which the majority of kdimensional subspaces of X are 4-Euclidean. Note that the presence of M in the definition corresponds to the right normalization, since the average of M ( E k ) o v e r Gn,k is equal to M for all 1 ~< k ~< n. Theorem 4.2.1 implies that k ( X ) >~ c n ( M / b ) 2. What is surprisingly simple is the observation [152] that an inverse inequality holds true. The estimate in Theorem 4.2.1 is sharp in full generality: THEOREM 4.3.1. k ( X ) ~ 4 n ( M / b ) 2. PROOF. Fix orthogonal subspaces E 1, . . ., E t of dimension k ( X ) such that ]Kn - ~-~I= 1 E i (there is no big loss in assuming that k ( X ) divides n). By the definition of k ( X ) , most orthogonal images of each E i are 4-Euclidean, so we can find u 9 O (n) such that 1
- M l x l ~ Ilxll ~ 2MIxl, 2
x 9 uE i
(2)
for every i -- 1 . . . . . t. Every x e IKn can be written in the form x - Y~.I=1Xi' where Xi 9 u E i . Since the xi's are orthogonal, we get t
Ilxll ~
2M
~ Ixil ~ 2M~/71xl.
(3)
i--1
This means that b ~< 2Mx/t, and since t - n / k ( X ) we see that k ( X ) <. 4 n ( M / b ) 2.
[]
In other words, the following asymptotic formula holds true: THEOREM 4.3.2. Let X be an n-dimensional normed space. Then,
k(X)'~n(M/b)
2.
Dvoretzky's theorem gives information about the central sections of a body, or equivalently, about the local structure of the corresponding normed space. By a global result we mean a statement about the full body or space. In order to describe the global version of Dvoretzky's theorem, we need to introduce a new quantity: DEFINITION. Let X -- (R n, I1 II), We define t (X) to be the smallest positive integer t for which there exist Ul . . . . . u t 9 0 (n) such that
1Mixl ~
1
t
~ i=1
for every x 9 Rn.
lluixll ~ 2MIxl
A.A. Giannopoulos and V.D. Milman
742
Geometrically speaking, t ( X ) is the smallest integer t for which there exist rotations vl . . . . . vt such that the average Minkowski sum of vi K ~ is 4-Euclidean. Once again, the presence of M in the definition corresponds to the correct normalization. It is proved in [37] that t ( X ) <<.c ( b / M ) 2 (we postpone a proof of this fact until Section 4.5). It was recently observed in [ 152] that a reverse inequality is true in full generality: THEOREM 4.3.3. t ( X ) >>.l ( b / M ) 2 . For the proof of this assertion we shall make use of the following lemma: LEMMA. L e t x l . . . . . xt E S n-1. There exists y ~ S n-1 such that Y~=I [(y, xi)l ~ ~/-i. PROOF. We consider all vectors of the form z(e) - - Z i = tI E i X i , where 6i = -4-1. If z = z(g) has maximal length among them, the parallelogram law shows that Izl f> ~/7. Also, t
~
t
[(z, xi)l ~ ~ ( z ,
i=1
"~tXi) = [zl 2 ~
Izl~/t,
(4)
i=1
Choosing y = z/Iz[ we conclude the proof. P R O O F OF T H E O R E M 4.3.3. Assume that we can find t orthogonal transformations
Ul . . . . . ut such that 71 ~ i =t 1 [luixl[ <~ 2Mlxl for every x ~ R n We find xo e S n-1 with Ilxoll = b (minimal distance from the origin). It is clear that 1 -Ilxoll,llxo[I = bllxoll,. We set x i - - u i- l xo and use the lemma to find y 6 S n-1 such that Y ~ it= I I(y, xi)l >>-47. Then, we have t
t
,/7 ~<~I{Y, ui-~xo)l- ~ i=1
t
](ui y, xo)] <~ Ilxol], ~
i=1
Ilui y]] <~
2Mt
i=1
b
(5)
This shows that 4t ~> ( b / M ) 2. Combining Theorem 4.3.3 with the upper bound for t (X) we obtain a second asymptotic formula: THEOREM 4.3.4. For every finite dimensional normed space X we have
t ( X ) ~ " ( b / M ) 2. Theorems 4.3.2 and 4.3.4 give a very precise asymptotic relation between a local and a global parameter of X [152]" FACT. There exists an absolute constant c > 0 such that 1 - n <~k ( X ) t ( X ) <. cn C
f o r every n-dimensional normed space X.
Euclidean structure in finite dimensional normed spaces
743
4.4. Applications of the concentration of measure on the sphere We used the concentration of measure on S n- J for the proof of Dvoretzky's theorem. The same principle applies in very different situations. We shall demonstrate this by two more examples.
(a) Banach-Mazur distance. Recall that by John's theorem d(X, g.~) <<.x/~ for every n-dimensional space X. Then, the multiplicative triangle inequality for d shows that d(X, Y) <~n for every pair of spaces X and Y. On the other hand, Gluskin [71] has proved that the diameter of the Banach-Mazur compactum is roughly equal to n: There exists an absolute constant c > 0 such that for every n we can find two n-dimensional spaces Xn, Yn with d (Xn, Yn) ~ cn. The spaces Xn, Yn in Gluskin's example are random and of the same nature: random symmetric polytopes with om vertices. We shall show that spaces whose unit balls are geometrically quite different objects have "small" distance [49]: THEOREM 4.4.1. Let X and Y be two n-dimensional normed spaces such that #Extr(Kx) <~ n c~ and #Extr(Kr.,) <<.n ~ for some or, fl > O, where #Extr(.) denotes the number of extreme points. Then,
d(X, Y) ~ Cv/O~ + 3v/n logn. (In other words, if a body has few extreme points and a second body has few facets, then their distance is not more than x/n log n.) PROOF. We may assume that ~1 Dn c__ K x c_ Dn C Ky C_ x/-ffDn. Then, Ky, c_ Dn. If
u E O(n), it is clear that ]]u -1 :Y --+ XI] ~< n. We are going to show that ]]u:X --+ YI] is small for a random u. We estimate the norm of u as follows: I]u:X--~ YII = sup Iluxllr = xEKx
max
max
x6Extr(Kx) y*6Extr(Kr, )
I(ux, y*)l.
Observe that if x E Extr(Kx) and y* E E x t r ( K r , ) , then ux, y* E Dn. It follows that
#(u E O(n)" ](ux, y*)] ~> e) ~< cexp(--e2n/2). Therefore, if cn ~+3 e x p ( - e 2 n / 2 ) < 1, we can find u E O(n) such that ]]u:X --+ YI] ~< e. Solving for e we see that we can choose
e "~ v/c~ + fl v/iog n / n . Hence, there exists u E O(n) for which
d(X, Y) <~ ]]u "X ---> VII ][u-" V ---> X][ <~ Cv/Ot + flv/n logn.
A.A. Giannopoulos and V.D. Milman
744
(b) Random projections. L
Let 1 <~ k ~< n, and E
~ Gn,k. A simple computation shows that
IPE(x)l 2 a ( d x ) -- k n-1
n
and since PE is a 1-Lipschitz function, concentration of measure on the sphere shows that
a(x E sn-l" l lPE(x)l - x / ~ n I > e) ~< Cl exp(--c282n) for every e > 0. Double integration and the choice e = 8~/-k-/n show that for any fixed subset {yl . . . . . YN } of S n-1 and any 8 6 (0, 1) we have
Vn,k(EeGn,k" ( 1 - - 6 ) v ~ / n < l P E ( Y j ) l < ( l + 6 )
kx/~, j~
> 1 - cl g exp(-c262k). If N ~< c]-1 exp(c282k), then we can find a k-dimensional subspace E such that IPE(Yj)I v/-k-/n for every j ~< N. It can be also arranged so that the distances of the yj's will shrink in a uniform way under PE (this application comes from [95]).
4.5. The concentration phenomenon: Ldvy families The concentration of measure on the sphere is just an example of the concentration phenomenon of invariant measures on high-dimensional structures. Assume that (X, d,/~) is a compact metric space with metric d and diameter diam(X) ~> 1, which is also equipped with a Borel probability measure #. We then define the concentration function (or "isoperimetric constant") of X by or(X; e) - 1 - inf{#(Ae): A Borel subset of X, #(A) ~> 89}, where Ae = {x 6 X: d(x, A) ~< e} is the e-extension of A. As a consequence of the isoperimetric inequality on S n+l we saw that
of(sn+l; 8)~
~/~78exp(-e2n/2),
an estimate which was crucial for the proof of Dvoretzky's theorem and the applications in Section 4.4. L6vy [ 110] was the first to observe the role of the dimension in this particular example. For this reason, a family (Xn, dn, #n) of metric probability spaces is called a normal L&y family with constants (cl, c2) (see [85] and [10]) if
ol(Xn, e) ~ Cl exp(--c282n), or, more generally, a Ldvyfamily if for every e > 0 a(X,;e) ~ 0
Euclidean structure in finite dimensional normed spaces
745
as n --+ cx~. There are many examples of L6vy families which have been discovered and used for local theory purposes. In most cases, new and very interesting techniques were invented in order to estimate the concentration function c~(X; e). We list some of them (and refer the reader to [175] in this volume for more information; see also [83,84] for a development in a different direction): (1) The family of the orthogonal groups (SO(n), Pn, lZn) equipped with the HilbertSchmidt metric and the Haar probability measure is a L6vy family with constants C1 = ~ / ~ / 8 and c2 = 1/8. (2) The family X~ -- I--[ 9 9m" S ~ with the natural Riemannian metric and the product probability measure is a L6vy family with constants cl = ~/-~-/8 and c2 = 1/2. (3) All homogeneous spaces of SO(n) inherit the property of forming L6vy families. In particular, any family of Stiefel manifolds W~,k,, or any family of Grassmann manifolds G~,k, is a L6vy family with the same constants as in (1). (All these examples of normal L6vy families come from [85].) (4) The space F n = { - 1 , 1}~ with the normalized Hamming distance d(r/, r/') = #{i ~< n: r/i ~ rll }/n and the normalized counting measure is a L6vy family with constants cl - 1/2 and c2 = 2. This follows from an isoperimetric inequality of Harper [90], and was first put in such form and used in [9]. (5) The group/7~ of permutations of {1 . . . . . n } with the normalized Hamming distance d(o-, r) = #{i ~< n: or(i) -~ r(i)}/n and the normalized counting measure satisfies oe(/7~; e) ~< 2exp(-e2n/64). This was proved by Maurey [122] with a martingale method, which was further developed in [172]. We shall give two more examples of situations where L6vy families are used. In particular, we shall complete the proof of the global form of Dvoretzky's theorem using the concentration phenomenon for products of spheres.
(a) A topological application. Let 1 ~< k ~< n and Vk = {(~,x): ~ E Gn,k, x E S(~)} be the canonical sphere bundle over Gn,k. Assume that f :S n-l --+ R is a Lipschitz function with the following property: For every ~ E Gn,k we can find x E S(~) such that f (x) = O. One can easily check that Vk is a homogeneous space of SO(n) whose concentration function satisfies oe(Vk; ~) ~< x / - ~ e x p ( - - e 2 n / 8 ) . A standard argument shows that given 6 > 0, if k <<.c62n/log(3/3) then we can find a subspace ~ E Gn,k and a f-net A/" of S(~) such that f ( x ) = 0 for every x E N'. Assuming that the Lipschitz constant of f is not large, we get [85]:
There exists ~ E Gn,k such that If(x)l ~ c6 for every x E S(~). (b) Global form of Dvoretzky's theorem. Recall that t (X) is the least positive integer for 1 t which there exist Ul . . . . . Ut E O(n) such that 89 ~< T Z i - 1 ]]uixl[ <~2Mix] for every x E • n.
A.A. Giannopoulos and V.D. Milman
746
We saw that 4t (X) ~> (b/M) 2. We shall now prove the reverse inequality (which is stated in Theorem 4.3.4) following [ 118]: Consider the space ,~t __ {~ __ (XI . . . . . Xt)" Xi E S n-1 }.
Define f ( s
1
t
-- 7 ~--~'~i:1 [Ixi II- Then, for every s ~ e ~t we have:
1
If()~) -- f ( Y ) l ~ 7 /=1
[[xi
-
Yi
[I ~<
i=1
Ilxi - Yi I12
~< ~ P ( s
Y).
The concentration estimate for products of spheres gives P r o b (] 7l / ~ _ 1 ]]xill - L f
> 6L f ) ~ exp(-c62tL2fn/b 2)
for every 6 e (0, 1). Equivalently, if x E S n-1 then
(1 - 6 ) L f
<<.t
Iluix[I~ (1 + 6 ) L f i=l
(Ui)i~t
in a subset of [O(n)] t of measure greater than 1 - exp(-c62tL2fn/b2). If .Af is a 6-net for S n-l, we can find Ul, . . . , ut E O(n) such that 71 y~ Iluixll "~ L f for all x e .M', provided that n/log(3/8) <~ c62tL2fn/b2. We choose 6 > 0 small enough so that for all
successive approximation will give 71 ~ Iluix II "~ L f for all x e S n- 1 and we verify that the condition will be satisfied for some t <~c I( b / L f ) 2. Since M "~ L f up to a multiplicative constant, the proof is complete, f-q
4.6. Dvoretzky's theorem and duality 4.6.1. Recall that if X = (R n, I1" II) is a normed space, then the dual norm is defined by Ilxll. - sup{l(x, Y)I" IlYll ~< 1}. It is clear that ~lxl ~< Ilxll. ~< alxl, hence if we define k* = k(X*) and M* = M(X*) then Theorem 4.3.2 shows that
k* ~- n(M*/a) 2. On the other hand, it is a trivial consequence of the Cauchy-Schwarz inequality that
MM* >1 )
n-1
n-1
Ilxlll/2llxll 1/2,
a(dx)
I(x,x)l 1/2a(dx)
-- 1.
(1)
Euclidean structure in finite dimensional normed spaces
747
Multiplying the estimates for k and k* we obtain
kk* ~ cn 2
(MM*) 2 >/cn2/(ab) 2 (ab) 2
(2)
Since we can always assume that ab ~< x/if, we have proved: THEOREM 4.6.1 ([61]). Let X be an n-dimensional normed space. Then,
k ( X ) k ( X * ) >~cn. This already shows that for every pair (X, X*) at least one of the quantities k, k* is greater than cv/-ff. Recall that for X - ~ we have k ( s ~ log n, therefore k(s ~> c n / l o g n - almost proportional to n. In fact, a direct computation shows that M(s b(s "~ ~/-ff, therefore k(s "~ n. Although d ( X , s is the maximal possible, ~ has Euclidean sections of dimension proportional to n. 4.6.2. Let k - min{k, k*}. Since Dvoretzky's theorem holds for random subspaces of the appropriate dimension, we can find a subspace E 6 G n,-~ on which we have
1
- M l x l <~ Ilxll ~< 2MIxl, 2
1M,
2
9
Ixl ~< Ilxll ~< 2M*lxl
(3)
simultaneously. This implies that IIPE'X ----->Ell ~ 4 M M * . We see this as follows: let x 6 R n. Then,
IPE(x)I 2 - ( P E ( x ) , x ) <~ IIeE(x)ll,llxll ~ 2M*IPE(x)I Ilxll,
(4)
since PE (x) 6 E. For the same reason,
IIPE(X)II ~ 2MIPE(x)I ~ 4MM*IIxlI.
(5)
kk* "~ n 2 (MM*)2 IIPE II2 (ab)2 >/cn 2 (ab)2 '
(6)
But then,
which is a strengthening of Theorem 4.6.1 [61 ]. In the example of X - s we know that k "~ log n, therefore our estimate shows that for a random subspace E (log n) of dimension roughly equal to log n we must have k(s
logn ~ cnllPEflogn)l] 2.
On the other hand, the norm of a random projection of s x/log n, so we get the correct estimate k(s >~cn.
of rank log n is known to exceed
A.A. Giannopoulos and V.D. Milman
748
4.6.3. Another example where the preceding computation gives precise information on several parameters of X is the case X -- g np, 1 < p < 2. Let q be the conjugate exponent of p. We need the following result [23] (see also [150, pp. 22]): THEOREM 4.6.2. k(gq) ~< c(q)n 2/q. l
2
It is a simple consequence of H61der's inequality that (ab) 2 ~< n - ~ for X = g np. Our computation in 4.6.2 and Theorem 4.6.2 show that if k -- min{k(gp), k(gq) }, then
c(q)nZ/qk(g,p) >/n
1+ 2
-4 I[PE(k)ll 2.
(7)
Since k(gp) ~< n (!), we immediately get:
THEOREM 4.6.3. Let 1 < p < 2 and q be its conjugate exponent. Then,
k(ep) ~" n,
k(eq) '~ .v/--qn2/q,
d(g.np, g.~) -- d(gq, s
1
1
~" n 2
q
4.6.4. A combinatorial application. We saw that the log n estimate in Dvoretzky's theorem is optimal by studying the example of ~ . The argument we used for the cube shows something more general: Let P be a symmetric polytope, and denote its number of facets by f ( P ) and its number of vertices by v(P). Then, k ~< log f ( P ) and since v(P) = f ( P ~ we also get k* ~< log v(P). We have seen that kk* >~cn, and this proves the following fact [61]: THEOREM 4.6.4. Let P be a symmetric polytope in ]1~n. Then, log f (P) log v (P) ~ cn.
4.7. Isomorphic versions of Dvoretzky's theorem 4.7.1. Bounded volume ratio. quantity
v r ( K ) = inf,
-~
Let K be a body in R n . The volume ratio of K is the
. EcK
,
where the inf is over all ellipsoids contained in K. It is easily checked that v r ( K ) is an affine invariant. We shall show that if a body K has small volume ratio, then the space X/( has subspaces F of dimension proportional to n which are "well-isomorphic" to ~dimF. THEOREM 4.7.5. Let K be a body in 1Rn with v r ( K ) = A. Then, for every k <~ n there exists a k-dimensional subspace F of X K such that
d ( F , g~) ~ (cA) ~-k .
Euclidean structure in finite dimensi, mal normed spaces
749
PROOF. We may assume that Dn is the maximal volume ellipsoid of K. Then, Ilxll Ix l for every x 6 R n. Given k ~< n, double integration shows that there exists F E Gn,k satisfying
fs
Ilxll -n o-~(dx) ~< v r ( K ) n -- A n.
(1)
n-lnF
Then, Markov's inequality shows that for any r > O, crk{x 6 S n - 1 N F: Ilxll < r} ~< (rA) n. If we consider just one point x in S n- 1 n F, then the r/2 neighbourhood of x with respect to I" I has trk measure greater than (cr) k, for some absolute constant c > O. This means that if (rA) n < (cr) k then the set {x 6 S n-1 n F: Ilxll >/r} is an r/2 net for S n-1 n F: if y E S n-1 n F, we can find x with Ix - Yl ~< r/2 and Ilxll/> r, and the triangle inequality shows that
Ilyll/> Ilxll- IIx - yll
> r -Ix
- yl >1 r/2.
(2)
This shows that d(F, g.k2) <~ 2/r. Analyzing the necessary condition on r we obtain
d(F, s
<~ (cA) ~-~-k.
D
(3)
Theorem 4.7.2 has its origin in the work of Kashin [ 100], who proved that there exist c(c~)-Euclidean subspaces of g~ of dimension [an], for every ot 6 (0, 1). Szarek [181] realized the fact that bounded volume ratio is responsible for this property of s while the notion of volume ratio was formally introduced somewhat later in [ 187]. 4.7.2.
A natural question related to Dvoretzky's theorem is to give an estimate for max
dim X=n
min{d(F, s
F C X, d i m F -- k}.
for each 1 <~ k ~< n. Such an "isomorphic" version was proved by Milman and Schechtman [ 151 ] who showed the following: THEOREM 4.7.6. There exists an absolute constant C > 0 such that, for every n and every k ~> C log n, every n-dimensional normed space X contains a k-dimensional subspace F for which
d(F, s
<~ C v / k / l o g ( n / k ).
For an extension to the nonsymmetric case, see [87,78].
5. The low M*-estimate and the quotient of subspace theorem 5.1. The low M*-estimate Dvoretzky's theorem gives very strong information about the Euclidean structure of kdimensional subspaces of an arbitrary n-dimensional space when their dimension k is up
A.A. Giannopoulos and V.D. Milman
750
to the order of log n. In some cases one can find Euclidean subspaces of dimension even proportional to n, but no "proportional theory" can be expected in such a strong sense. However, surprisingly enough, there is nontrivial Euclidean structure in subspaces of dimension )~n, )~ 6 (0, 1), even for )~ very close to 1. The first step in this direction is the low M*-estimate: THEOREM 5.1.1. There exists a function f ' ( 0 , 1) -+ R + such that for every A. E (0, 1) and every n-dimensional normed space X, a random subspace E E Gn,[Zn] satisfies f()~)lxl M*
~ Ilxll
x E E,
(1)
where c > 0 is an absolute constant.
Theorem 5.1.1 was originally proved in [133] and a second proof using the isoperimetric inequality on S n-1 was given in [134], where (1) was shown to hold with f()~) ) c(1 - )~) for some absolute constant c > 0 (and with an estimate f(~.) >~ 1 - )~ + o()~) as ~. --+ 0+). This was later improved to f()~) ~> c~/1 - )~ in [156] (see also [140] for a different proof with this best possible ~/1 - ) ~ dependence). Finally, it was proved in [76] that one can have 1
Geometrically speaking, Theorem 5.1.1 says that for a random ~.n-dimensional section of K x we have M ~
Kx n E c ~Dn - f ()~)
N E,
(3)
that is, the diameter of a random section of a body of dimension proportional to n is controlled by the mean width M* of the body (a random section does not feel the diameter a of K x but the radius M* which is roughly the level r at which half of the supporting hyperplanes of r Dn cut the body K x ) . The dual formulation of the theorem has an interesting geometric interpretation. A random )~n-dimensional projection of K x contains a ball of radius of the order of 1/M. More precisely, for a random E E Gn,xn we have P E ( K x ) D_
f()~) Dn N E. M
(4)
We shall present the proof from [ 134] which gives linear dependence in ~. and is based on the isoperimetric inequality for s n - l : PROOF OF THE LOW M*-ESTIMATE. Consider the set A -- {y E S n-l" obviously have a ( A ) ~> 1/2.
Ilyll, ~ 2M*}. We
Euclidean structure in finite dimensional n o r m e d spaces
751
CLAIM. For every ~. E (0, 1) there exists a s u b s p a c e E o f d i m e n s i o n k = [kn] such that (5)
E N S n - 1 c_ A ( ~ _ 5 ) , where 6 ~ c(1 - )~).
PROOF OF THE CLAIM. W e have o'(Arr/4) ~ 1 - - c ~ ' f f f o / 4 s i n n - Z t dt, and double integration through G n , k s h o w s that a random E E G n , k satisfies
~rk(nrr/4 n E ) ~ 1 - c x / n
f
Tr/4
sin n-2 t dt.
J0
(6)
On the other hand, for every x E S n - 1 n E we have
f0
sin k-2 t dt.
(7)
This means that if sin k-2 t dt ~
sin n-2 t dt,
(8)
Jr _ a) ~ 0 , and hence x E AT_~ then Arc~4 n B ( x , -~ ~ . Analyzing the sufficient condition (8) we see that we can choose 6 ~> c(1 - Z). D We complete the proof of T h e o r e m 5.1.1 as follows: Let x E S n-1 N E. There exists y E A such that sin6 ~< I(x, y)l ~< Ilyll,llxll ~ 2M*llxll,
(9)
2 ~> Ct (1 - )~), the theorem follows. and since sin 6 >~ if6
5.2. The g.-position
Let X be an n-dimensional normed space. Figiel and Tomczak-Jaegermann [60] defined the t~-norm of T 6 L (g~, X) by
e ( T ) -- ~
n-I
IITyll 2 cr (dy)
.
(1)
Alternatively, if {e j} is any orthonormal basis in ]~n, and if gl . . . . . gn are independent standard Gaussian random variables on some probability space $2, we have
g~(T)--
(11 II) E
giT(ei)
i=1
where E denotes expectation.
,
(2)
A.A. Giannopoulos and V.D. Milman
752
Let now RadnX be the subspace of L2(I-2, X) consisting of functions of the form
Y~4n_=lgi (og)Xi where xi 9 X (the notation here is perhaps not canonical, but convenient). The natural projection from L2(s
X) onto RadnX is defined by
(3)
Radn f = ~ ( fs2 gi f ) g i" i=1
We write IIRadn IIx for the norm of this projection as an operator in L2(S2, X). The dual norm e* is defined on L(X, ~ ) by g*(S) - sup{tr(ST)" T e L(e'~, X), e(T) <~1}.
(4)
From a general result of Lewis [111] it follows that for some T e L(g~, X) one has g(T)g* (T -1) -- n. Using this fact, Figiel and Tomczak-Jaegermann proved that for every n-dimensional space X there exists T'g~ --+ X such that
e(r)e((r-1) *) ~ nllRadnllX.
(5)
The norm of the projection Radn was estimated by Pisier [ 161 ]: For every n-dimensional space X, ]]Radn ]Ix ~< clog[d(X, g~) + 1].
(6)
This implies that for every X = (R n , I1" II) we can define a Euclidean structure (., .) (called the e-structure) on R n, for which
M(X)M*(X) ~ clog[d(X, ~ ) + 1].
(7)
Equivalently, we can state the following theorem: THEOREM 5.2.1. Let K be a symmetric convex body in IRn. There exists a position K of
K for which
M(YM*(Y
clog[d(XK,
+
(8)
where c > 0 is an absolute constant. Pisier's argument uses symmetry in an essential way, therefore one cannot transfer directly this line of thinking to the nonsymmetric case. For recent progress on the nonsymmetric M M*-estimate, see Appendix 7.2.
5.3. The quotient of subspace theorem The quotient of subspace theorem [135] states that by performing two operations on an n-dimensional space, taking first a subspace and then a quotient of it, we can always
Euclidean structure in finite dimensional normed spaces
753
arrive at a new space of dimension proportional to n which is (independently of n) close to Euclidean:
THEOREM 5.3.1 (Milman). Let X be an n-dimensional normed space and c~ E [ 89 1).
Then, there exist subspaces E D F of X for which k - d i m ( E / F ) > an,
d ( E / F , e~) <. c(1 - o~)-1110g(1 - c~)l.
(1)
1 Geometrically, this means that for every body K in ~;~n and any ot E [ ~, 1), we can find subspaces G C E with dim G ~> an and an ellipsoid s such that
C C PG(K M E) C c(1
-
ot) -1
Ilog(1 - oe)lC.
(2)
The proof of the theorem is based on the low M*-estimate and an iteration procedure which makes essential use of the e-position. PROOF. We may assume that Kx is in g-position: then, by Theorem 5.2.1 we have M ( X ) M * ( X ) <~clog[d(X, g~) + 1]. Step 1: Let )v E (0, 1). We shall show that there exist a subspace E of X, dim E ~> )vn, and a subspace F of E*, d i m F -- k ~> ,ken, such that d ( F , t~k) ~< c(1 - X)-l log[d(X, e~) + 1]. The proof of this fact is a double application of the low M*-estimate. By Theorem 5.1.1, a random )vn-dimensional subspace E of X satisfies Clx/1 --,k
M*(X)
Ixl ~ Ilxll ~ blxl,
x E E.
(3)
Moreover, since (3) holds for a random E E Gn,zn, we may also assume that M ( E ) <~ c2 M (X). Therefore, repeating the same argument for E*, we may find a subspace F of E* with dim F = k ~> ~.2n and c3~/1--~.
M(X)
Ixl ~<
Cl~/1--~.
M*(E*)
Ixl <~ IlxllE* ~<
M*(X) Cl~/1 - ) v
Ixl
(4)
for every x E F. Since K x is in g-position, we obtain
d(F, g~) <~c4(1
- ~)-1
M ( X ) M * ( X ) <~c(1 - )v)-' log[d(X, g~) -I- 1].
(5)
Step 2: Denote by QS(X) the class of all quotient spaces of a subspace of X, and define a function f : ( 0 , 1) -+ IR+ by f ( o t ) - inf{d(F, e~)" F E QS(X), d i m F > a n } .
(6)
Then, what we have really proved in Step 1 is the estimate f()v2ez) ~ c(1 -- X) -1 log f(ot).
(7)
A.A. Giannopoulos and V.D. Milman
754
An iteration lemma (see [135] or [164, pp. 130]) allows us to conclude that f (or) ~< c(1 -
oe)-I I log(1
- ot) l.
5.4. Variants and applications of the low M*-estimate 1. An almost direct consequence of the low M*-estimate is the existence of a function f : (0, 1) --+ R + with the following property [142]:
If K is a body in R n and if X ~ (0, 1), then a random Xn-dimensional section K N F of K satisfies diam(K N F) ~ 2r, where r is the solution of the equation (1)
M * ( K n r D n ) = f(X)r.
One can choose f0~) = (1 - e)~/1 - )~ for any e 6 (0, 1), and then (1) is satisfied for all F in a subset of Gn,[)~n] of measure greater than 1 - C l e x p ( - c z e 2 ( 1 - X)n). 2. Let t (r) = t (XK; r) be the greatest integer k for which a random subspace F ~ Gn,k satisfies diam(K n F) ~< 2r. The following linear duality relation was proved in [141]:
If t* (r) = t(X*; r), then for any ~ > O and any r > Owe have t(r) + t * ( ~ r )
(2)
~ ( 1 - ~)n - C,
where C > 0 is an absolute constant. This surprisingly precise connection between the structure of proportional sections of a body and its polar is also expressed as follows [67]:
Let ~ > 0 and k, 1 be integers with k + I <, (1 - ~ )n. Then, for every body K in R n we have fG
M * ( K N F) dvn,k(F) fG n,k
M* (K ~ n F') dvn,t(F') <, C
(3)
n,l
where C > 0 is an absolute constant. 3. An estimate dual to (1) was established in [65]. There exists a second function g ' ( 0 , 1) --+ R such that: for every body K in ~n and every X 6 ( 89 1), a random Xndimensional section K N F of K satisfies diam(K n F) ~> 2r, where r is the solution of the equation
M * ( K n r D n ) = g(X)r.
(4)
Euclidean structure in finite dimensional normed spaces
755
This double sided estimate provided by (1) and (4) may be viewed as an (incomplete) asymptotic formula for the diameter of random proportional sections of K, which is of interest from the computational geometry point of view since the function r --+ M* (K A r Dn) is easily computable. 4. The diameter of proportional dimensional sections of K is connected with the following global parameter of K: For every integer t >~ 2 we define r t ( K ) to be the smallest r > 0 for which there exist rotations ul . . . . . ut such that u l ( K ) A . . . A u t ( K ) ___r D n . If R t ( K ) is the smallest R > 0 for which most of the [ n / t ] - d i m e n s i o n a l sections of K satisfy diam(K A F) ~< 2R, then it is proved in [142] that rzt (K) ~< x/tRt (K). The fact that a reverse comparison of these two parameters is possible was established in [66]: There exists an absolute constant C > 1 such that Rc,(K)
(5)
~ Ctrt(K)
for every t ~> 2. 5. Fix an orthonormal basis {el . . . . . en } of R n . Then, for every non empty ~r ___{ 1 . . . . . n } we define the coordinate subspace It~~ - span{e j" j e ~r }. We are often interested in analogues of the low M*-estimate with the additional restriction that the subspace E should be a coordinate subspace of a given proportional dimension (see [63] for applications to Dvoretzky-Rogers factorization questions). Such estimates are sometimes possible [64]: If K is an ellipsoid in R n, then for every )~ e (0, 1) we can find cr ___ {1 . . . . . n } of cardinality ]or I ~> (1 - )~)n such that [)~/log(1/)O] 1/2 PR~, ( K ) D_
MK
(6)
D n AIR '~ .
Analogues of this hold true if the volume ratio of K or the cotype-2 constant of X/< is small. Finally, let us mention that Bourgain's solution of the A (p) problem [27] (see also [ 189] and [29]) is closely related to the following "coordinate" result: Let (4~i)i ~
(7)
for every choice of reals (ti)iecr. We refer the reader to the article [97] in this collection for the results of Bourgain and Tzafriri on restricted invertibility, which are closely related to the above.
A.A. Giannopoulos and V.D. Milman
756
6. Isomorphic symmetrization and applications to classical convexity 6.1. Estimates on covering numbers Let K1 and K2 be convex bodies in •n. The covering number N ( K 1 , K2) of K1 by K2 is the least positive integer N for which there exist Xl . . . . . xu E ]1~n s u c h that N
(1)
K1 C_ U ( x i -F K 2 ) . i=1
We shall formulate and sketch the proofs of a few important results on covering numbers which we need in the next section. The well known Sudakov inequality [179] estimates N ( K , t Dn): THEOREM 6.1.1. Let K be a body in R n. Then, N ( K , tDn) <~ e x p ( c n ( M * / t ) 2)
(2)
f o r every t > O, where c > 0 is an absolute constant. The dual Sudakov inequality, proved by Pajor and Tomczak-Jaegermann [ 155], gives an upper bound for N (Dn, t K): THEOREM 6.1.2. Let K be a symmetric convex body in •n. Then, N ( D n , t K ) ~ e x p ( c n ( M / t ) 2)
(3)
f o r every t > O, where c > 0 is an absolute constant. We shall give a simple proof of Theorem 6.1.2 which is due to Talagrand (see [ 109, pp. 82]). PROOF OF THEOREM 6.1.2. We consider the standard Gaussian probability measure Yn on R n, with density dyn - (2;r) -n/2 e x p ( - I x l 2 / 2 ) d x . A direct computation shows that Markov's inequality shows that
f Ilxll din(X) - otnM, where an/V/-n --+ 1 as n ~ ec. 1
yn(X" Ilxll ~ 2Morn) >~ -~.
(4)
Let {Xl . . . . . XN} be a subset of Dn which is maximal under the requirement that Ilxit xj II t> t, i 7~ j. Then, the sets xi + -2 K have disjoint interiors. The same holds true for the
Euclidean structure in finite dimensional normed spaces sets Yi -+-2Morn K, Yi
-
757
(4Morn/ t)xi . Therefore,
-
N
Z
Yn (Yi nt- 2Morn K)
(5)
<~ 1.
i--1
Using the convexity of e -s, the symmetry of K and (4), we can then estimate Yn (Yi -k2M~n K) from below as follows: 1 exp(_(4Motn/t)e) ?'n(Yi + 2M~n K) >/-~
(6)
Now, (5) shows that
(7)
N <. 2exp((4Motn/t)2), and since C~n ~ ~
we conclude the proof.
Sudakov's inequality (Theorem 6.1.1) can be deduced from Theorem 6.1.2 with a duality argument of Tomczak-Jaegermann [ 194]" Let A -- sup t (log N(Dn, tK~
(8)
t>0
t2
We check that 2K A (-y K o) _ t Dn for every t > 0, and this implies that
N(K tDn) <. N(K, 2K N ( T12 K ~
t2 K o N(K ~ -g )
<. N(K, 2tDn)N(Dn, gt K~
(9)
t(logN(K, tDn))1/2 < t(logN(K, 2tDn))1/2 + 8A,
(10)
This shows that
from which we easily get
supt(logN(K, tDn)) 1/2 <. 16A.
(11)
t>0
This is equivalent to the assertion of Theorem 6.1.1 (just observe that M*(K) - M(K~
A weaker version of Sudakov's inequality can be proved if we use Urysohn's inequality: For every body K and any t > 0, we have
N(K, tDn) ~ exp(2nM*/t).
(12)
758
A.A. Giannopoulos and V.D. Milman
PROOF. Consider a set {xl XN} C K which is maximal under the requirement int(xi + ~t Dn) n int(xj + t Dn) --0. Then, . . . . .
+ ~Dn[ N ( K , tDn) <. N <. [K ~t .... I~Onl
=
( ~ ) n IK + ~Dn[ t IDnl '
(13)
and Urysohn's inequality shows that N ( K , tDn) ~ =
t
(M*(K + (t/2)Dn)
=
( 12+M * ) ~<. exp(2nM*/t). t
M* + -~ D
(14)
Using the covering numbers one can compare volumes of convex bodies in various situations. A main ingredient of the proof of the lemmas below (which may be found in [139]) is the Brunn-Minkowski inequality: LEMMA 1. Let K, T, and P be symmetric convex bodies in R n. Then, IK n (T + x ) + PI ~< IK n T + PI
(15)
for every x E ]~n. PROOF. Let Tx = K n (T + x) + P. We easily check that Tx + T-x c__2T0, and then apply the Brunn-Minkowski inequality. [2] LEMMA 2. Let K and P be symmetric convex bodies in ~n. If t > O, then IK + PI <~N(K, tDn)l(K n t D n ) + PI.
(16)
PROOF. If K c Ui<~N K n (Xi + tDn), then K + P c__ Ui<~N[(Xi -Jr-tDn) n K + P]. We compare volumes using the information from Lemma 1. D LEMMA 3. Let K and L be symmetric convex bodies in R n. Assume that L c_ b K for some b ~ 1. Then,
(
N co(KUL),
(')) 1+-
K
<.2bnN(L,K).
(17)
n
Using Lemma 3 with L = 71Dn and combining with Lemma 2, we have: LEMMA 4. Let K and P be symmetric convex bodies in ]Rn. Assume that Dn ~ tbK for some t > O. Then, Ico(K u (1/t)Dn) +
PI <~2ebnN(Dn,
t K ) I K + PI.
(18)
Euclidean structure in finite dimensional normed spaces
759
6.2. Isomorphic symmetrization and applications to classical convexity The functional analytic approach and the methods of local theory lead to new isomorphic geometric inequalities. In this way, the ideas we described in previous sections find applications to classical convexity theory in R n . We shall describe two results in this direction: 6.2.1. The inverse Blaschke-Santal6 inequality [34]. There exists an absolute constant c > 0 such that
0
~ ID.IID,,I
~<1
(1)
f o r every body in R n.
The inequality on the right is the Blaschke-Santal6 inequality: the volume product s ( K ) = [KIIK~ is maximized (among symmetric convex bodies) exactly when K is an ellipsoid. A well-known conjecture of Mahler states that s ( K ) >~ 4 ~/n! for every K. This has been verified for some classes of bodies, e.g., zonoids and 1-unconditional bodies (see [165,129,171,79]). The left handside inequality comes from [34] and answers the question of Mahler in the asymptotic sense: For every body K, the affine invariant s (K) 1/,1 is of the order of 1/ n. 6.2.2. The inverse Brunn-Minkowski inequality [136]. There exists an absolute constant C > 0 with the following property" For every body K in IRn there exists an ellipsoid MK such that ]K] -- [MK[ and f o r every body T in IR'~
1 IMK + T[ ~/n ~ IK 4- TI ~/~ ~ CIMK + TI ~/"
C
(2)
This implies that for every body K in IRn there exists a position K -- u K(K) of volume [KI - ]K] such that the following reverse Brunn-Minkowski inequality holds true: "If K I and K2 are bodies in R n, then
ItlK1 + t2/~211/n ~ C(tl]Kl] 1In -I- t2]/~2[l/n),
(3)
f o r all tl , t2 > O, where C > 0 is an absolute constant"
The ellipsoid M x in 6.2.2 is called an M-ellipsoid for K. Analogously, the body K -u K ( K ) is called an M-position of K (and then, one may take M~, = pDn). The symmetry of K is not really needed in 6.2.1 and 6.2.2 (see, e.g., [148]). Both results were originally proved by a dimension descending procedure which was based on the quotient of subspace theorem. We shall present a second approach, which appeared in [139] and introduced an "isomorphic symmetrization" technique. This is a symmetrization scheme which is in many ways different from the classical symmetrizations. In each step, none of the natural parameters of the body is being preserved, but the
760
A.A. Giannopoulos and V.D. Milman
ones which are of interest remain under control. After a finite number of steps, the body has come close to an ellipsoid and this is sufficient for our purposes, but there is no natural notion of convergence to an ellipsoid. 6.2.3. Remarks.
Applying (2) for T -- MK we get
IK + MKI ~/n ~ CIKI ~/n.
(4)
This is equivalent to Theorem 6.2.2 and to each one of the following statements: (i) There exists a constant C > 0 such that for every body K we can find an ellipsoid MK with I M K I - IKI and
N ( K , MK) <, exp(Cn). (ii) There exists a constant C > 0 such that for every body K we can find an ellipsoid MK with [MK[ - [K[ and
N ( M K , K) <~exp(Cn). We can also pass to polars and show that for every body T in 1
C
~,~n,
IMp. -I- TI ~/~ <. IK ~ -+- TI 1/n <~CIM~ -I- TI ~/~
Since the M-position is isomorphically defined, one may ask for stronger regularity on the covering numbers estimates (i) and (ii)" Pisier proved (see [ 164, Chapter 7]) that, for every a > 1/2 and every body K there exists an affine image K of K which satisfies IKI - [Dn I and
max{N(K, tDn),N(Dn,tK),N(K~176
<<.exp(c(a)nt -1/a)
for every t ~> 1, where c ( a ) is a constant depending only on a, with c ( a ) = O ((a -
1)-1/2)
as a --+ 1. We then say that K is in M-position of order a (a-regular in the terminology of [ 164]). PROOF OF THE THEOREMS. Since s ( K ) is an affine invariant, we may assume that K is in a position such that M ( K ) M * ( K ) <~ clog[d(XK, s + 1]. We may also normalize so that M ( K ) -- 1. We define
)~1 -- M * ( K ) a l ,
!
)~l -- M ( K ) a l ,
(5)
for some a l > 1, and consider the new body
[
1 ]9
K 1 -- co (K A ~ 1Dn) tO -fillDn
(6)
Euclidean structure in finite dimensional n o r m e d spaces
761
Using Sudakov's inequality and L e m m a 2 with P = {0}, we see that IKll ~> IK A~.ID.I ~> IKI/N(K,)~ID,)
>~IKlexp(-cn/a2),
(7)
while using the dual Sudakov inequality and Lemma 3 we get
( 1 ) <~2e-~l nN(Dn, )~1K)IKI b
IK~I~ co K U ~--~1Dn
~< exp(cn/a2).
(8)
The same computation can be applied to K~, and this shows that ( K | ) <~exp(cn/a 2). exp(-cn/a2) <~ss(K)
(9)
We continue in the same way. We now know that d(XK~, g~) <~ M(K)M*(K)a 2 and, since s(K|) is an affine invariant, we may assume that M(KI)M*(KI) <~clog[d(XK~, g.~) + 1] and M (K1) = 1. We then define )~2 -- M * ( K 1 ) a 2 ,
(10)
~/2 -- M ( K I ) a 2 ,
and consider the body K2 - - c o [ ( K l O )~2Dn) U 1-,2-Dn].Estimating volumes, we see that
exp(-cn/a 2) <~s(K2) <~exp(cn/a2)"
(11)
s(K1)
We iterate this scheme, choosing al - logn, a2 -- l o g l o g n . . . . . at -- log(t)n - the titerated logarithm of n, and stop the procedure at the first t for which at < 2. It is easy to check that d(XK,, g~) <~C, therefore 1
C
<~s(Kt) 1In <~C.
(12)
On the other hand, combining our volume estimates we see that
cl ~< exp - c
((1
+--.+a2
1))
1))
~< exp c ~212+..-+a-- ~
~<
(13)
,
which proves Theorem 6.1.1 since the series
s(Kt) |/n s(K)l/n
1 1 a---~ -+-""" -q- a-~ -+- " " " remains bounded by an /
absolute constant.
D
762
A.A. Giannopoulos and V.D. Milman
The proof of Theorem 6.2.2 follows the same pattern. In each step, we verify that for every convex body T IKs + T] <~exp(cn/a2) ' e x p ( - c n / a 2) <~ IK s - , + T[
(14)
and the same holds true for K s. At the tth step, we arrive at a body Kt which is Cisomorphic to an ellipsoid M, and (14) shows that IKtl 1/n ~ IKI 1/n up to an absolute constant. If we define MI( -- p M where p > 0 is such that IMKI -- IKI, then p ~ 1 and using (14) we conclude the proof. NOTE. The existence of the M-ellipsoid MK of K in the nonsymmetric case was established in [148]. The key lemma is the observation that if 0 is the centroid of the convex body K, then [K A ( - K)[ ~> 2 -n [K[. We close this section with a few geometric consequences of the M-position" (1) Every body K has a position K with the following property: there exist u, v ~ SO(n) such that if we set P = K + u ( K ) and Q - P~ + v(P~ then Q is equivalent to a Euclidean ball up to an absolute constant. Actually, this statement is satisfied for a random pair (u, v) ~ SO(n) x SO(n). This double operation may be called isomorphic Euclidean regularization. Compare with the following examples" If K is the unit cube, then P is already equivalent to a ball for most u ~ SO(n) (this follows from [100], see 4.7.1). If K is the unit ball of ~ , the second operation is certainly needed. A closely related result from [142] is the following isomorphic inequality connecting K with K ~ Let
pt(K)--max
{
p > 0: pDn C u i ( K ) , Ui E O(n) t i=1
} .
Then, there exists an absolute constant c > 0 such that p 2 ( K ) p 3 ( K ~ >/c for every body K in I~n. Observe that Kashin's result is a consequence of this fact: if K is the cube, then p3(K ~ ~< c/~/-ff. Therefore, K + u ( K ) D c,,/-ffDn for some u ~ O(n). It is not clear if two rotations of K ~ suffice for a similar statement. (2) One may use the M-position in order to obtain a random version of the quotient of subspace theorem: If K is in M-position, then using Remark 6.2.3(i) we see that every )~n-dimensional projection P E ( K ) of K has finite volume ratio (which depends on ~). We can therefore apply Theorem 4.7.2 to conclude that a random )~2n-dimensional section P F ( K ) N E of P F ( K ) has distance depending only on )~ from the corresponding Euclidean ball.
Euclidean structure in finite dimensional normed spaces
763
7. Appendix 7.1. The hyperplane conjecture In Section 2.3 we saw that every body in IK" has an isotropic position K with IKI = 1, which satisfies
fK(
x, O)2 dx -- L2
(1)
for every 0 E S "-1 . This position is uniquely determined up to orthogonal transformations, and the affine invariant LK is called the isotropic constant of K. It is an open problem whether there exists an absolute constant C > 0 such that L K ~< C for every body K. Let K be a body in R n . Using Theorem 2.3.6, one can easily check that
nL 2 <~ Idetul
luKll§ ~
f~:
lux
12
dx
(2)
for every invertible linear transformation u. For the same reason,
nL2o <~
Idet(u-1)* .... I fK [(u - l ) * (x)l 2 dx. I(u-l),(go)ll+ 88 o
We may choose u ' X K -+ g~ such that d(X1,:, g~) --Ilull that
(3)
Ilu -111.
Then, (2) and (3) imply
n2L2 L2o d2(X , e )(lugl I(u-')*(K~ -2In,
(4)
and an application of the inverse Santal6 inequality shows that
(5) Therefore, duality gives the following first estimates on the isotropic constant: THEOREM 7.1.1. Let K be a body in N n. Then, LK ~ cd(XK, g~) ~ cx/n. Moreover, either L K ~ c ~/-ff or L Ko ~ c ~/~. Bourgain [28] has proved that LK <, c~/-dlogn, where c > 0 is an absolute constant, for every body K. We shall give a proof of this fact following Dar's presentation in [46]. Recall that for every 0 6 S n-1 and p > 1 we have
1 fx ) 1/p -~l I(x, 0)1 p dx <~cp -~1
](x, 0)l dx,
(6)
764
A.A. Giannopoulos and V.D. Milman
where c > 0 is an absolute constant. This is a consequence of Borell's lemma (see Section 2.2). It follows from Section 2.2 (23) that if K is isotropic, then ~
e x p ( l ( x , O ) l / c L K ) d x <~ 2,
(7)
for every 0 E S n-1 , where c > 0 is an absolute constant. We shall use this information in the following form: LEMMA 1. Let K be an isotropic body. I f N is a finite subset o f S n-1 , then
f/
max I(x, 0)l dx ~< c L K log INI.
(8)
O6N
Starting with an isotropic body K, we see from Theorem 2.3.6 that Ixl 2dx =
n g 2 <, - -
n
IITxlIK~dx --
(x, r x ) dx
(9)
yETKmaxI(x, Y)I dx
for every symmetric, positive-definite volume preserving transformation T of ]1~ n . In order to estimate this last integral, we first reduce the problem to a discrete one using the DudleyFernique decomposition: LEMMA 2. Let A be a body in •n, and R be its diameter. For every r and j = 1 . . . . . r, we can find finite subsets Nj o f A with log lNjl ~< c n ( w ( A ) 2 J / R ) 2 with the following property: every x E A can be written in the form X ~-Zl -+-'''-'F-Zr @Wr,
where zj E Z j - (Nj - N j - 1 ) M ( 3 R / 2 J ) D n and l/Or E ( R / 2 r ) D n (we set No -- {o}).
The proof of this decomposition is simple. The estimate on the cardinality of Nj comes from Sudakov's inequality (Theorem 6.1.1). We now choose T in (9) so that A = T K will have minimal mean width: Theorem 5.2.1 allows us to assume that w ( T K ) <<,cv/-fflogn. From Lemma 2, we see that for every x E K,
max I(y,x)l ~
yETK
2..,maxl(z,x)l + j=l
zEZj
~3R j=l
- m a x I(~,, x)l +
2J zEZj
max
wE(R/2r)Dn
R -~
Ixl
'
I(w,x)l
(10)
Euclidean structure in finite dimensional normed spaces where ~ -- z/Jzi 6 S "-1 9Now, Lemma 1 and the estimate o n
/K
max J{~, x)i dx ~< cLKloglZjl <~cnLK
~czj
INjl
765
imply that
( w(TK)2J ) 2 R
(11)
for every j - 1 . . . . . r. Going back to (9), we conclude that
nL 2 <<.cLK
nw2(TK)--~ + ~ 7 ~ j----1
C'LK nw2(TK)--R + ~-Tw/n ,
(12)
and the optimal choice for r gives
nL 2 <~c~/-nw(TK)v/-nLK.
(13)
Since w(TK) <~cv/-filogn, the proof is complete. THEOREM 7.1.2. For every body K in R n we have LK ~ c~/-nlogn. REMARK. The same holds true for nonsymmetric convex bodies as well (see [158]).
7.2. Geometry of the Banach-Mazur compactum 1. Consider the set B, of all equivalence classes of n-dimensional normed spaces X -(Rn, [l" ll), where X is equivalent to X ~ if and only if X and X' are isometric. Then,/3, becomes a compact metric space with the metric log d, where d is the Banach-Mazur distance (the Banach-Mazur compactum). There are many interesting questions about the structure of the Banach-Mazur compactum, and most of them remain open. Below, we describe some fundamental results and problems in this area. The interested reader will find more information in the book [ 195] and the surveys [74,184]. 2. John's theorem shows that d(X, Y) <~n for every X, Y 6 / 3 , . Therefore, diam(B,) ~< n. The natural question of the exact order of diam(B,) remained open for many years and was finally answered by Gluskin [71]: diam(Bn) ~> cn. Gluskin does not describe a pair X, Y E 13n with d(X, Y) >~cn explicitly (in fact, there is no concrete example of spaces with distance of order greater than ~/-fi). The idea of the proof is probabilistic" a random T ' g ~ ~ g~ satisfies I]T II IIT-Ill ~> cn, and this suggests that by "spoiling" g~ it is possible to obtain X and Y with distance cn. The spaces which were used in [71] have as their unit ball a body ofthe form K = co{-+-ei,-+-xj: 1 <~j <~2n}, where {ei } is the standard orthonormal basis of R" and the xj's are chosen uniformly and
A.A. Giannopoulos and V.D. Milman
766
independently from the unit sphere S n - 1 . A random pair of such spaces has the desired property. This method of considering random spaces proved to be very fruitful in problems where one needed to establish "pathological behavior". We mention Szarek's finite dimensional analogue of Enflo's example [56] of a space failing the approximation property: there exist n-dimensional normed spaces whose basis constant is of the order of ~ [182]. See also [72,120] and subsequent work of Szarek and Mankiewicz where random spaces play a central role. The article [121 ] in this collection covers this topic. 3. Another natural question about the geometry of the Banach-Mazur compactum is that of the uniqueness of its center: If dimX = n and d(X, Y) <~c~ff for every Y ~ Bn, is it then true that X is "close" (depending on c) to g~ ? This question was answered in the negative by Bourgain and Szarek [35]" Let X0 = ~ @ e /17 - - S , where s = [n/2]. Then, d(Xo, Y) <~ c~/-ff for every Y E Bn (and, clearly, d(Xo, g~) >~c'~/-ff). The proof of the fact that X0 is an asymptotic center of the compactum is based on the proportional version of the DvoretzkyRogers lemma (see Section 4.1). 4. Fix X 6/3n. Then, one can define the radius of/3n with respect to X by R(X) -max{d(X, Y): Y ~/3n}. Many problems of obvious geometric interest arise if one wants to give the order of the radius with respect to important concrete centers. For example, the problem of the distance to the cube R(e~r remains open. It is known that R(g~a) ~< cn 5/6 (see [35,186] and [62]). On the other hand, Szarek has proved [183] that R(g n ) >~ c v/-fflog n, therefore ~ and ~ are not asymptotic centers of the compactum (these are actually the only concrete examples of spaces for which this property has been established). 5. If we restrict ourselves to subclasses of/3,,, then the diameter may be significantly smaller than n: Let An be the family of all 1-symmetric spaces. Tomczak-Jaegermann [ 192] (see also [73]) proved that d(X, Y) <, c ~ whenever X, Y E .An. This result is clearly optimal: recall that d ( ~ , ~ ) = v/ft. The analogous problem for the family of 1-unconditional spaces remains open. Lindenstrauss and Szankowski [115] have shown that in this case d (X, Y) <, cn ~, where ol is a constant close to 2/3. It is conjectured that the right order is close to v/-fi. The diameter of other subclasses of/3n was estimated with the method of random orthogonal factorizations. The idea (which has its origin in work of Tomczak-Jaegermann [ 190], and was later developed and used by Benyamini and Gordon [22]) is to use the average of IITIIx_+YIIT -~ IIY-+x with respect to the probability Haar measure on SO(n) as an upper bound for d(X, Y). Using this method one can prove a general inequality in terms of the type-2 constants of the spaces [22,49]:
a(x, Y) c,/a[T2(X) + T2(Y*)] for every X, Y ~ 13n. This was further improved by Bourgain and Milman [33] to
a x,
<. c(a(v, e' )r2 x + a(x,
Euclidean structure in finite dimensional normed spaces
767
In [33] it is also shown that d (X, X*) ~< c(log n) • n 5/6 for every X 6 B~. All these results indicate that the distance between spaces whose unit balls are "quite different" should be significantly smaller than diam(Bn). 6. The Banach-Mazur distance d(K, L) between two not necessarily symmetric convex bodies K and L is the smallest d > 0 for which there exist Z l, Z2 E ]]~n and T ~ GLn such that K - zi c_ T ( L - z2) c d ( K - zi). The question of the maximal distance between nonsymmetric bodies is open. John's theorem implies that d ( K , L) <~ n 2. Better estimates were obtained with the method of random orthogonal factorizations and recent progress on the nonsymmetric analogue of the M M*-estimate (Theorem 5.2.1). In [19] it was proved that every convex body K has an affine image K| such that M (K|) M* (K1) ~ Cx//-n, a bound which was improved to cn 1/3 log ~ n, fl > 0 in [170]. Using this fact, Rudelson showed that d ( K , L) ~ cn 4/3 log ~ n for any K, L E/Cn. See also recent work of Litvak and Tomczak-Jaegermann [116] for related estimates in the nonsymmetric case. 7. Milman and Wolfson [ 153] studied spaces X whose distance from g~ is extremal. They showed that if d ( X , ~ ) - x/n, then X has a k-dimensional subspace F with k >i c log n which is isometric to g~. The example of X -- ~ shows that this estimate is exact. An isomorphic version of this result is also possible [153]: If d ( X , ~ ) >~ otx/-fi for some c~ E (0, 1), then X has a k-dimensional subspace F (with k - - h ( n ) -+ ec as n ~ cx~) which satisfies d(F, ~ ) <<,c(ot), where c(ot) depends only on c~. The original estimate for k in [153] was later improved to k >~ Cl (or)logn through work of Kashin, Bourgain and Tomczak-Jaegermann (see [195, Section 31] for details). An extension of this fact appears in [160]" Recall that a Banach space X contains g.~ 's uniformly if X contains a sequence of subspaces Fn, n ~ N with d (Fn, ~ ) <~ C. Then, the following are equivalent: (i) X does not contain gnl' S uniformly. (ii) sup{d(F, ~)" F C X, dim F -- n} -- o(x/n). (iii) There exists a sequence otn = o(v/-ff) with the following property: If F is an ndimensional subspace of X, there exists a projection P :X ~ F with IIP [I ~< otn. In the nonsymmetric case the extremal distance to the ball is n. Palmon [157] showed that d (K, Dn) = n if and only if K is a simplex. 8. Tomczak-Jaegermann [193] defined the weak distance wd (X, Y) of two n-dimensional normed spaces X and Y by w d ( X , Y) = max{q(X, Y), q(Y, X)}, where q ( X , Y) -- i n f J ~ IIS(~o)ll IIT(~o)ll dco, and the inf is taken over all measure spaces S2 and all maps T :S2 --+ L ( X , Y), S:F2 --+ L(Y, X) such that fs2 S(co) o T(co)dco = idx. It is not hard to check that w d ( X , Y) <<, d ( X , Y) and that with high probability the weak distance between two Gluskin spaces is bounded by cx/-n. In fact, Rudelson [168] has proved that w d ( X , Y) <, cn |3/14 log 15/7 n for all X, Y c Bn. It is conjectured that the weak distance in Bn is always bounded by cx/ft.
768
A.A. Giannopoulos and V.D. Milman
7.3. Symmetrization and approximation Symmetrization procedures play an important role in classical convexity. The question of how many successive syrnmetrizations of a certain type are needed in order to obtain from a given body K a body K which is close to a ball was extensively studied with the methods of local theory. This study led to the surprising fact that only few such operations suffice: Let K 6 ]~n and u 6 S n-1 . Consider the reflection Zru with respect to the hyperplane orthogonal to u. The Minkowski symmetrization of K with respect to u is the convex body !2 ( K + ZruK) Observe that this operation is linear and preserves mean width. A random Minkowski symmetrization of K is a body 7ru K, where u is chosen randomly o n S n-1 with respect to the probability measure cr. In [37] it was proved that for every s > 0 there exists no(s) such that for every n ~> no and K 6/Cn, if we perform N = Cn log n + c(s)n independent random Minkowski symmetrizations on K we receive a convex body K such that ( 1 - s ) w ( K ) D n C K C (1 + s ) w ( K ) D n with probability greater than 1 - e x p ( - c l (s)n). The method of proof is closely related to the concentration phenomenon for SO(n). The same question for Steiner symmetrization was studied in [38]. Mani [119] has proved that, starting with a body K 6/Cn, if we choose an infinite random sequence of directions uj ~ S n-1 and apply successive Steiner symmetrizations (ruj of K in these directions, then we almost surely get a sequence of convex bodies converging to a ball. The number of steps needed in order to bring K at a fixed distance from a ball is much smaller [38]: If K 6/Cn with IKI -- [Dnl, we can find N <~cln logn and Ul . . . . . uU ~ S n-1 such that D
c 2 1D n C__(~UN
0''"
0
(ru~)(K) cc_c2Dn,
(1)
where cl, C2 > 0 are absolute constants. It is not clear what the bound f (n, s) on N would be if we wanted to replace c2 by 1 - s, s 6 (0, 1). The proof of (1) is based on the previous result about Minkowski symmetrizations. Results of the same nature concern questions about approximation of convex bodies by Minkowski sums. The global form of Dvoretzky's theorem is an isomorphic statement of this type. Recall that a zonotope is a Minkowski sum of line segments, and a zonoid is a body in R n which is the Hausdorff limit of a sequence of zonotopes. A body is a zonoid if and only if its polar body is the unit ball of an n-dimensional subspace of L 1(0, 1) (for this and other characterizations of zonoids, see [24]). The unit ball of ep is a zonoid if and only if 2 ~< p ~< c~ (see [51 ]). In particular, the Euclidean unit ball Dn can be approximated arbitrarily well by sums of segments. The question of how many segments are needed in order to come (1 + s)-close to Dn is equivalent to the problem of embedding t~ into ~ . From the results in [61] it follows that N <~ c(s)n segments are enough. In [39] it was shown that the same bound on N allows us to choose the segments having the same length. The linear dependence of N on n is
Euclidean structure in finite dimensional normed spaces
769
optimal, but the best possible answer if we view N as a function of both n and e is not known (see [30,32,39,113,196]). If we replace the ball Dn by an arbitrary zonoid Z, then the same approximation problem is equivalent to the question of embedding an n-dimensional subspace of L 1(0, 1) into ~ U. Bourgain, Lindenstrauss and Milman [39] proved, by an adaptation of the empirical distribution method of Schechtman [173], that for every e 9 (0, 1) there exist N <~ce-Zn logn and segments I1 IN such that (1 - e)Z C Y~ Ij C (1 + e)Z. Moreover, if the norm of Z is strictly convex then N can be chosen to be of the order of n up to a factor which depends on e and the modulus of convexity of II" ]lz. Later, Talagrand [188] showed (with a considerably simpler approach) that one can have N ~< cllRadn 112e-2n. For more information on this topic, we refer the reader to the surveys [112,114] and [97]. . . . . .
7.4. Quasi-convex bodies Many of the results that we presented about symmetric convex bodies can be extended to a much wider class of bodies. We have already discussed extensions of the main facts to the nonsymmetric convex case. We now briefly discuss extensions to the class of quasi-convex bodies. Recall that a star body K is called quasi-convex if K + K C cK for some constant c > 0. Equivalently, if the gauge f of K satisfies (i) f ( x ) > 0 if x ~ 0, (ii) f()~x) = [,klf(x) for any x 9 R n, and (iii) f 9 C (oe), i.e., there exists c~ 9 (0, 1] such that
off(x) <~ ( f , f)(x)" - - i n f { f ( x | ) + f(x2), xl + x 2 - - x } ,
x 9 n.
A body K is called p-convex, p 9 (0, 1), if for any x, y 9 K and )~, # > 0 with )~P + #P = 1 we have )~x + #y 9 K. Every p-convex body K is quasi-convex, and K + K C 21/p K. Conversely, for every quasi-convex body K (with constant C) we can find a q-convex body KI such that K C Kl C 2K, where 21/q = 2C (see [166]). Most of the basic results we described in the previous sections were extended to this case. Versions of the Dvoretzky-Rogers lemma and Dvoretzky's theorem were proved by Dilworth [50]. For the low M*-estimate and the quotient of subspace theorem in the quasi-convex setting, see [117] and [77] respectively (see also [144] for an isomorphic Euclidean regularization result and the random version of the QS-theorem). The reverse Brunn-Minkowski inequality is shown in [21]. For results on existence of M-ellipsoids, entropy estimates and asymptotic formulas, see [117,118] and [148]. In most of the cases, the tools which were available from the convex case were not enough, and new techniques had to be invented: some of them provided interesting alternative proofs of the known "convex results".
7.5. Type and cotype The notions of type and cotype were introduced by Hoffmann-JCrgensen [92] in connection with limit theorems for independent Banach space valued random variables. Their
A.A. Giannopoulosand V.D.Milman
770
importance for the study of geometric properties of Banach spaces was realized through the work of Maurey and Pisier (see the article [123] in this collection for a discussion of the development of this theory). Given an n-dimensional normed space X, and 1 ~< p ~< 2 (2 ~< q < cx~, respectively), the type-p (cotype-q) constant Tp(X) (Cq(X)) of X is the smallest T > 0 (C > 0) such that: for every m E N and X l . . . . . Xm E X,
(So
m
Zri(t)xi
dt
<<,T
i=l
i=1
Ilxi IIp
i=1
m
Z Ilxi Ilq
(5)l,, 2)j2)
~ C
Zri(t)xi
respectively.
i=l
Results of Tomczak-Jaegermann ([191] when p = q = 2), K6nig ([104] for any p and q not equal to 2, up to constants depending on p, q) and Szarek [ 185] show that in order to determine the (Gaussian) type-p or cotype-q constants of X up to an absolute constant, it is enough to consider n vectors. In the Rademacher case, the definite answer is not yet known. It is clear that Tz(g~) = C 2 ( ~ ) - 1 and, conversely, Kwapiefi [106] proved that
d(X, ~) ~ C2(X)T2(X). Letkp(X; e), 1 ~ p ~ cx~,be the largest integer k ~< n for which gpk is (1 + e)-isomorphic to a subspace of X (in this terminology, k(X) = k2(X; 4)). The following results show how type and cotype enter in the study of the linear structure of a space: (i) In [61] it is shown that k2(X) ~> cn/C2(X) and k2(X) ~> cn2/q/c2(x). This gives another proof of the facts k2(gp) ~> cn, 1 ~< p ~< 2, and k2(gq) ~ n 2/q, q ~ 2. (ii) In [161] it is proved that kp(X; e) ~ c(p, e)Tp(X) q, where 1 < p < 2 and 1/p + 1/q -- 1. This generalizes the estimate kp(g~; e) ~ c(p, e)n, 1 ~< p ~< 2, of Johnson and Schechtman [96]. (iii) A quantitative version of Krivine's theorem [ 10] states that, for every A ~> e,
kp(X; e) >/c(8, A)[kp(X; A)] cl(e/A)p. Gowers [80,81 ] obtained related estimates on the length of (1 + e)-symmetric basic sequences in X. (iv) In [124] it is shown that if no cotype-q constant of X is bounded by a number independent of n, then X contains (1 § e)-isomorphic copies of g k for large k. Alon and Milman [8], using combinatorial methods, provided a quantitative form of this fact: kz(X; 1)k~(X; 1) ~> exp(cx/logn). Bourgain and Milman [34] proved that vr(Kx) <~f(C2(X)). Thus, spaces with bounded cotype-2 constant satisfy all consequences of bounded volume ratio (this had been independently observed, see, e.g., [61,55]). Milman and Pisier [149] introduced the class of spaces with the weak cotype 2 property: X is weak cotype 2 if there exists 6 > 0 such that kz(E) ~> 6dimE for every E C X. One can then prove that vr(E) ~ C(~) for every E C X [149].
771
E u c l i d e a n structure in finite d i m e n s i o n a l n o r m e d s p a c e s
In Section 6.2 we saw that every n-dimensional normed space X has a subspace E with d i m E ~> n/2 such that vr(KE,) <, C. This suffices for a proof of the quotient of subspace theorem. However, the following question remains open: does every X contain a subspace E with d i m E i> n/2 such that C2(E*) ~ C? This problem is related to many open questions in the local theory (for a discussion see [137,145]). Finally, let us mention the connection between Gaussian and Rademacher averages [ 125]: Let X be an n-dimensional normed space, and {xj } be a finite sequence in X. Then, do)) 1/2
Erj(t)xj J
J 2
<~ c ( l + l o g n ) l / 2 ( f o l
1/2
Zrj(t)xj J
If X has bounded cotype-q constant Cq (X) for some q ~> 2, then the constant in the right hand side inequality may be replaced by cx/~C q (X).
7.6. Nonlinear type theory Let (T, d) be a metric space, and F n = { - 1, 1}n with the normalized counting measure #n. An n-dimensional cube in T is a function f : F n --+ T. For any such f and i 6 {1 . . . . . n}, we define (Ai f ) ( s ) = d ( f ( s l . . . . .
Ei . . . . .
En),
f (sl .... , -el . . . . . Sn)).
A metric space (T, d) has metric type p, 1 ~< p ~< 2, if there exists a constant C > 0 such that, for every n 6 N and every f : F n --+ T we have
j--1 Every metric space has type 1, and if 1 ~< p i ~< p2 ~ 2, metric type p2 implies metric type p i. Let 4~ : (Tl, di ) --+ (T2, d2) be a map between metric spaces. The Lipschitz norm of 4~ is defined by
I1~ [[Lip = sup
t#,~
d2(4~ (t), 4~(s)) dl(t,s)
Let Fp be the space F n equipped with the metric induced by ~np. We say that a metric space (T, d) contains F~'s (1 + s)-uniformly if for every n ~ N there exist a subset Tn C T and a bijection Ckn "Fp -+ Tn such that 114~[[Lipl[~bnl []Lip ~ 1 + s. Bourgain, Milman and Wolfson [40] (see also [ 154]) proved the following:
772
A.A. Giannopoulos a n d V.D. M i l m a n
THEOREM 7.6.1. A metric space (T, d) has metric type p f o r some p > 1 if and only if there exists s > 0 such that T does not contain F~ 's (1 + s)-uniformly. A natural question which arises is to compare the notions of metric type and type in the case where T is a normed space. An answer to this question was given in [40], see also [163]:
THEOREM 7.6.2. Let X be a Banach space and let 1 < p < 2. (i) If X has type (respectively, metric type) p, then X has metric type (respectively, type) Pl f o r all 1 <<.Pl < P.
(ii) X contains F~ 's uniformly if and only if X contains g'ln's uniformly. We refer the interested reader to [40,163] for the proofs of these facts, and a comparison with another notion of metric type which was earlier proposed by Enflo [57]. In [40] and [36] one can find a generalization of Dvoretzky's theorem for metric spaces: For every e > 0 there exists a constant c(e) > 0 with the following property: every metric space T of cardinality N contains a subspace S with cardinality at least c(e) log N such that for some C ~2 with ISl - ISl we can find a bijection ~b" S --~ S with IIr -111Lip ~< 1 4- e (this means that S is (1 + e)-isomorphic to a subset of a Hilbert space). Let us finally mention an interesting connection between nonlinear problems and a more advanced form of type and cotype, the so-called Markov type and cotype which was introduced and studied by Ball [17].
References [1] S. Alesker, Integrals of smooth and analytic functions over Minkowski's sums of convex sets, Convex Geometric Analysis, MSRI Publications, Vol. 34 (1998), 1-16. [2] S. Alesker, Continuous rotation invariant valuations on convex sets, Ann. of Math. 149 (1999), 977-1005. [3] S. Alesker, On P. McMullen's conjecture on translation invariant valuations in It~4, Adv. Math. (to appear). [4] S. Alesker, Description of translation invariant valuations on convex sets with solution of P. McMullen's conjecture, Geom. Funct. Anal. (to appear). [5] S. Alesker, S. Dar and V.D. Milman, A remarkable measure preserving diffeomorphism between two convex bodies in R n, Geom. Dedicata 74 (1999), 201-212. [6] A.D. Alexandrov, On the theory of mixed volumes of convex bodies, II: New inequalities between mixed volumes and their applications, Mat. Sb. N.S. 2 (1937), 1205-1238 (in Russian). [7] A.D. Alexandrov, On the theory of mixed volumes of convex bodies, IV." Mixed discriminants and mixed volumes, Mat. Sb. N.S. 3 (1938), 227-251 (in Russian). [8] N. Alon and V.D. Milman, Embedding of g k in finite-dimensional Banach spaces, Israel J. Math. 45 (1983), 265-280. [9] D. Amir and V.D. Milman, Unconditional and symmetric sets in n-dimensional normed spaces, Israel J. Math. 37 (1980), 3-20. [10] D. Amir and V.D. Milman, A quantitative finite-dimensional Krivine theorem, Israel J. Math. 50 (1985), 1-12. [ 11 ] J. Arias-de-Reyna, K. Ball and R. Villa, Concentration of the distance infinite dimensional normed spaces, Mathematika 45 (1998), 245-252. [12] K.M. Ball, Isometric problems in ep and sections of convex sets, Ph.D. Dissertation, Trinity College, Cambridge (1986).
Euclidean structure in finite dimensional normed spaces
773
[13] K.M. Ball, Normed spaces with a weak Gordon-Lewis property, Lecture Notes in Math. 1470, Springer, Berlin (1991), 36-47. [14] K.M. Ball, Shadows ofconvex bodies, Trans. Amer. Math. Soc. 327 (1991), 891-901. [15] K.M. Ball, Volume ratios and a reverse isoperimetric inequality, J. London Math. Soc. (2) 44 (1991), 351-359. [16] K.M. Ball, Ellipsoids ofmaximal volume in convex bodies, Geom. Dedicata 41 (1992), 241-250. [17] K.M. Ball, Markov chains, Riesz transforms and Lipschitz maps, Geom. Funct. Anal. 2 (1992), 137-172. [ 18] K.M. Ball, Convex geometry and functional analysis, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 161-194. [ 19] W. Banaszczyk, A. Litvak, A. Pajor and S.J. Szarek, The flatness theorem for nonsymmetric convex bodies via the local theory of Banach spaces, Math. Oper. Res. 24 (1999), 728-750. [20] F. Barthe, On a reverse form of the Brascamp-Lieb inequality, Invent. Math. 134 (1998), 335-361. [21 ] J. Bastero, J. Bernu6s and A. Pena, An extension of Milman's reverse Brunn-Minkowski inequality, Geom. Funct. Anal. 5 (1995), 572-581. [22] Y. Benyamini and Y. Gordon, Random factorization of operators between Banach spaces, J. d'Analyse Math. 39 (1981), 45-74. [23] G. Bennett, L.E. Dor, V. Goodman, W.B. Johnson and C.M. Newman, On uncomplemented subspaces of L p, 1 < p < 2, Israel J. Math. 26 (1977), 178-187. [24] E.D. Bolker, A class of convex bodies, Trans. Amer. Math. Soc. 145 (1969), 323-346. [25] C. Borell, The Brunn-Minkowski inequality in Gauss space, Inventiones Math. 30 (1975), 207-216. [26] J. Bourgain, On high dimensional maximal functions associated to convex bodies, Amer. J. Math. 108 (1986), 1467-1476. [27] J. Bourgain, Bounded orthogonal sets and the A(p)-setproblem, Acta Math. 162 (1989), 227-246. [28] J. Bourgain, On the distribution of polynomials on high dimensional convex sets, Lecture Notes in Math. 1469, Springer, Berlin (1991), 127-137. [29] J. Bourgain, A p-sets in analysis: Results, problems and related aspects, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 195-232. [30] J. Bourgain and J. Lindenstrauss, Distribution of points on spheres and approximation by zonotopes, Israel J. Math. 64 (1988), 25-31. [31] J. Bourgain and J. Lindenstrauss, Almost Euclidean sections in spaces with a symmetric basis, Lecture Notes in Math. 1376 (1989), 278-288. [32] J. Bourgain and J. Lindenstrauss, Approximating the sphere by a sum of segments of equal length, J. Discrete Comput. Geom. 9 (1993), 131-144. [33] J. Bourgain and V.D. Milman, Distances between normed spaces, their subspaces and quotient spaces, Integral Eq. Operator Th. 9 (1986), 31-46. [34] J. Bourgain and V.D. Milman, New volume ratio properties for convex symmetric bodies in R n, Invent. Math. 88 (1987), 319-340. [35] J. Bourgain and S.J. Szarek, The Banach-Mazur distance to the cube and the Dvoretzky-Rogers factorization, Israel J. Math. 62 (1988), 169-180. [36] J. Bourgain, T. Figiel and V.D. Milman, On Hilbertian subsets offinite metric spaces, Israel J. Math. 55 (1986), 147-152. [37] J. Bourgain, J. Lindenstrauss and V.D. Milman, Minkowski sums and symmetrizations, Lecture Notes in Math. 1317 (1988), 44-66. [38] J. Bourgain, J. Lindenstrauss and V.D. Milman, Estimates related to Steiner symmetrizations, Lecture Notes in Math. 1376 (1989), 264-273. [39] J. Bourgain, J. Lindenstrauss and V.D. Milman, Approximation of zonoids by zonotopes, Acta Math. 162 (1989), 73-141. [40] J. Bourgain, V.D. Milman and H. Wolfson, On the type of metric spaces, Trans. Amer. Math. Soc. 294 (1986), 295-317. [41] H.J. Brascamp and E.H. Lieb, Best constants in Young's inequality, its converse and its generalization to more than three functions, Adv. in Math. 20 (1976), 151-173. [42] Y. Brenier, Polar factorization and monotone rearrangement of vector-valued functions, Comm. Pure Appl. Math. 44 (1991), 375-417.
774
A.A. Giannopoulos and V.D. Milman
[43] Y.D. Burago and V.A. Zalgaller, Geometric Inequalities, Springer Series in Soviet Mathematics, SpringerVerlag, Berlin (1988). [44] L.A. Caffarelli, A-priori Estimates and the Geometry of the Monge-Ampkre Equation, Park City/IAS Mathematics Series, Vol. 2 (1992). [45] C. Carath6odory and E. Study, Zwei Beweise des Satzes dass der Kreis unter allen Figuren gleichen Urnfangs den grOssten Inhalt, Math. Ann. 68 (1909), 133-144. [46] S. Dar, Remarks on Bourgain's problem on slicing of convex bodies, Geometric Aspects of Functional Analysis, Operator Theory: Advances and Applications, Vol. 77 (1995), 61-66. [47] S. Dar, On the isotropic constant ofnonsymmetric convex bodies, Israel J. Math. 97 (1997), 151-156. [48] S. Dar, Isotropic constants of Schatten class spaces, Convex Geometric Analysis, MSRI Publications, Vol. 34 (1998), 77-80. [49] W.J. Davis, V.D. Milman and N. Tomczak-Jaegermann, The distance between certain n-dimensional spaces, Israel J. Math. 39 (1981), 1-15. [50] S.J. Dilworth, The dimension of Euclidean subspaces of quasi-normed spaces, Math. Proc. Cambridge Phil. Soc. 97 (1985), 311-320. [51] L.E. Dor, Potentials and isometric embeddings in L 1, Israel J. Math. 24 (1976), 260-268. [52] A. Dvoretzky, A theorem on convex bodies and applications to Banach spaces, Proc. Nat. Acad. Sci. USA 45 (1959), 223-226. [53] A. Dvoretzky, Some results on convex bodies and Banach spaces, Proc. Sympos. Linear Spaces, Jerusalem (1961), 123-161. [54] A. Dvoretzky and C.A. Rogers, Absolute and unconditional convergence in normed linear spaces, Proc. Nat. Acad. Sci. USA 36 (1950), 192-197. [55] S. Dilworth and S. Szarek, The cotype constant and almost Euclidean decomposition of finite dimensional normed spaces, Israel J. Math. 52 (1985), 82-96. [56] P. Enflo, A counterexample to the approximation property, Acta Math. 130 (1973), 309-317. [57] P. Enflo, Uniform homeomorphisms between Banach spaces, S6minaire Maurey-Schwartz 75-76, Expos6 no. 18, Ecole Polytechnique, Paris. [58] W. Fenchel, Indgalitds quadratiques entre les volumes mixtes des corps convexes, C. R. Acad. Sci. Paris 203 (1936), 647-650. [59] T. Figiel, A short proof of Dvoretzky's theorem, Compositio Math. 33 (1976), 297-301. [60] T. Figiel and N. Tomczak-Jaegermann, Projections onto Hilbertian subspaces of Banach spaces, Israel J. Math. 33 (1979), 155-171. [61 ] T. Figiel, J. Lindenstrauss and V.D. Milman, The dimension of almost spherical sections of convex bodies, Acta Math. 139 (1977), 53-94. [62] A.A. Giannopoulos, A note on the Banach-Mazur distance to the cube, Geometric Aspects of Functional Analysis, Operator Theory: Advances and Applications, Vol. 77 (1995), 67-73. [63] A.A. Giannopoulos, A proportional Dvoretzky-Rogers factorization result, Proc. Amer. Math. Soc. 124 (1996), 233-241. [64] A.A. Giannopoulos and V.D. Milman, Low M*-estimates on coordinate subspaces, J. Funct. Anal. 147 (1997), 457-484. [65] A.A. Giannopoulos and V.D. Milman, On the diameter of proportional sections of a symmetric convex body, International Mathematics Research Notices 1 (1997), 5-19. [66] A.A. Giannopoulos and V.D. Milman, How small can the intersection of a few rotations of a symmetric convex body be ?, C. R. Acad. Sci. Paris 325 (1997), 389-394. [67] A.A. Giannopoulos and V.D. Milman, Mean width and diameter of proportional sections of a symmetric convex body, J. Reine Angew. Math. 497 (1998), 113-139. [68] A.A. Giannopoulos and V.D. Milman, Extremal problems and isotropic positions of convex bodies, Israel J. Math. 117 (2000), 29-60. [69] A.A. Giannopoulos and M. Papadimitrakis, Isotropic surface area measures, Mathematika 46 (1999), 113. [70] A.A. Giannopoulos, I. Perissinaki and A. Tsolomitis, John's theorem for an arbitrary pair of convex bodies, Geom. Dedicata (to appear). [71 ] E.D. Gluskin, The diameter of the Minkowski compactum is approximately equal to n, Funct. Anal. Appl. 15 (1981), 72-73.
Euclidean structure in finite dimensional normed spaces
775
[72] E.D. Gluskin, Finite dimensional analogues of spaces without basis, Dokl. Akad. Nauk USSR 216 (1981), 1046-1050. [73] E.D. Gluskin, On distances between some symmetric spaces, J. Soviet Math. 22 (1983), 1841-1846. [74] E.D. Gluskin, Probability in the geometry of Banach spaces, Proc. Int. Congr. Berkeley, Vol. 2 (1986), 924-938. [75] Y. Gordon, Gaussian processes and almost spherical sections of convex bodies, Ann. Probab. 16 (1988), 180-188. [76] Y. Gordon, On Milman's inequality and random subspaces which escape through a mesh in •n, Lecture Notes in Math. 1317 (1988), 84-106. [77] Y. Gordon and N.J. Kalton, Local structure theory for quasi-normed spaces, Bull. Sci. Math. 118 (1994), 441-453. [78] Y. Gordon, O. Gu6don and M. Meyer, An isomorphic Dvoretzky's theorem for convex bodies, Studia Math. 127 (1998), 191-200. [79] Y. Gordon, M. Meyer and S. Reisner, Zonoids with minimal volume product - a new proof, Proc. Amer. Math. Soc. 104 (1988), 273-276. [80] W.T. Gowers, Symmetric block bases infinite-dimensional normed spaces, Israel J. Math. 68 (1989), 193219. [81] W.T. Gowers, Symmetric block bases of sequences with large average growth, Israel J. Math. 69 (1990), 129-151. [82] M. Gromov, Convex sets and Kiihler manifolds, Advances in Differential Geometry and Topology, World Scientific, Teaneck, NJ (1990), 1-38. [83] M. Gromov, Metric Structures for Riemannian and Non-Riemannian Spaces, based on "Structures m6triques des vari6t6s Riemanniennes", L. LaFontaine and E Pansu, eds, Birkh~iuser, Boston (1999) (with Appendices by M. Katz, E Pansu and S. Semmes; English translation by Sean M. Bates). [84] M. Gromov, Spaces and question, Proceedings of Visions in Mathematics Conference, Israel 1999, GAFA, Special Volume issue 1, GAFA (2000) (to appear). [85] M. Gromov and V.D. Milman, A topological application of the isoperimetric inequality, Amer. J. Math. 105 (1983), 843-854. [86] M. Gromov and V.D. Milman, Brunn theorem and a concentration of volume phenomenon for symmetric convex bodies, GAFA Seminar Notes, Tel Aviv University (1984). [87] O. Gu6don, Gaussian version of a theorem of Milman and Schechtman, Positivity 1 (1997), 1-5. [88] O. Gu6don, Kahane-Khinchine type inequalities for negative exponent, Mathematika 46 (1999), 165-173. [89] H. Hadwiger, Vorlesungen iiber Inhalt, Oberfliiche under lsoperimetrie, Springer, Berlin (1957). [90] L.H. Harper, Optimal numberings and isoperimetric problems on graphs, J. Combin. Theory 1 (1966), 385-393. [91] D. Hensley, Slicing convex bodies: bounds of slice area in terms of the body's covariance, Proc. Amer. Math. Soc. 79 (1980), 619-625. [92] J. Hoffman-JCrgensen, Sums of independent Banach space valued random variables, Studia Math. 52 (1974), 159-186. [93] L. H6rmander, Notions of Convexity, Progress in Math. 127, Birkh~iuser, Boston (1994). [94] F. John, Extremum problems with inequalities as subsidiary conditions, Courant Anniversary Volume, Interscience, New York (1948), 187-204. [95] W.B. Johnson and J. Lindenstrauss, Extensions of Lipschitz mappings into a Hilbert space, Conference in Modern Analysis and Probability, New Haven, CT (1982), 189-206. [96] W.B. Johnson and G. Schechtman, Embedding g.p m into g.~, Acta Math. 149 (1982), 71-85. [97] W.B. Johnson and G. Schechtman, Finite dimensional subspaces of Lp, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 837-870. [98] M. Junge, Proportional subspaces of spaces with unconditional basis have good volume properties, Geometric Aspects of Functional Analysis, Operator Theory: Advances and Applications, Vol. 77 (1995), 121-129. [99] M. Junge, On the hyperplane conjecture for quotient spaces of L p, Forum Math. 6 (1994), 617-635. [100] B.S. Kashin, Sections of some finite-dimensional sets and classes of smooth functions, Izv. Akad. Nauk SSSR Ser. Mat. 41 (1977), 334-351. [ 101 ] D. Klain, A short proof of Hadwiger's characterization theorem, Mathematika 42 (1995), 329-339.
776
A.A. Giannopoulos and V.D. Milman
[102] A.G. Khovanskii and A.V. Pukhlikov, Finitely additive measures on virtual polyhedra, St. Petersburg Math. J. 4 (1993), 337-356. [103] H. Knrthe, Contributions to the theory of convex bodies, Michigan Math. J. 4 (1957), 39-52. [104] H. Krnig, Type constants and (q, 2)-summing norms defined by n vectors, Israel J. Math. 37 (1980), 130138. [ 105] H. K6nig, M. Meyer and A. Pajor, The isotropy constants of the Schatten classes are bounded, Math. Ann. 312 (1998), 773-783. [106] S. Kwapiefi, Isomorphic characterizations of inner product spaces by orthogonal series with vector valued coefficients, Studia Math. 44 (1972), 583-595. [ 107] D.G. Larman and P. Mani, Almost ellipsoidal sections and projections of convex bodies, Math. Proc. Cambridge Phil. Soc. 77 (1975), 529-546. [108] R. Latata, On the equivalence between geometric and arithmetic means for log-concave measures, Convex Geometric Analysis, MSRI Publications, Vol. 34 (1998), 123-128. [109] M. Ledoux and M. Talagrand, Probability in Banach Spaces, Ergeb. Math. Grenzgeb., 3. Folge, B. 23, Springer, Berlin (1991). [110] P. Lrvy, Problkmes Concrets d'Analyse Fonctionelle, Gauthier-Villars, Paris (1951). [111] D.R. Lewis, Ellipsoids defined by Banach ideal norms, Mathematika 26 (1979), 18-29. [112] J. Lindenstrauss, Almost spherical sections, their existence and their applications, Jahresber. Deutsch. Math.-Verein., Jubil~iumstagung 1990, Teubner, Stuttgart (1990), 39-61. [ 113] J. Linhart, Approximation of a ball by zonotopes using uniform distribution on the sphere, Arch. Math. 53 (1989), 82-86. [114] J. Lindenstrauss and V.D. Milman, The local theory of normed spaces and its applications to convexity, Handbook of Convex Geometry, P.M. Gruber and J.M. Wills, eds, Elsevier, Amsterdam (1993), 11491220. [115] J. Lindenstrauss and A. Szankowski, On the Banach-Mazur Distance Between Spaces Having an Unconditional Basis, Math. Studies 122, North-Holland, Amsterdam (1986). [ 116] A. Litvak and N. Tomczak-Jaegermann, Random aspects of the behavior of high-dimensional convex bodies, Lecture Notes in Math. 1745 (to appear). [ 117] A. Litvak, V.D. Milman and A. Pajor, Covering numbers and "low M*-estimate "for quasi-convex bodies, Proc. Amer. Math. Soc. 127 (1999), 1499-1507. [ 118] A. Litvak, V.D. Milman and G. Schechtman, Averages of norms and quasi-norms, Math. Ann. 312 (1998), 95-124. [119] P. Mani, Random Steiner symmetrizations, Studia Sci. Math. Hungar. 21 (1986), 373-378. [120] P. Mankiewicz, Finite dimensional spaces with symmetry constant of order V/ff, Studia Math. 79 (1984), 193-200. [ 121] P. Mankiewicz and N. Tomczak-Jaegermann, Quotients offinite-dimensional Banach spaces; random phenomena, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [122] B. Maurey, Constructions de suites sym~triques, C. R. Acad. Sci. Paris 288 (1979), 679-681. [123] B. Maurey, Type, cotype and K-convexity, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [124] B. Maurey and G. Pisier, Caract~risation d'une classe d'espaces de Banach par des propri~t~s de s~ries al~atoires vectorielles, C. R. Acad. Sci. Paris 277 (1973), 687-690. [125] B. Maurey and G. Pisier, Series de variables aleatoires vectorielles independentes et proprietes geometriques des espaces de Banach, Studia Math. 58 (1976), 45-90. [ 126] R.J. McCann, Existence and uniqueness of monotone measure preserving maps, Duke Math. J. 80 (1995), 309-323. [127] P. McMullen, Valuations and Euler-type relations on certain classes of convex polytopes, Proc. London Math. Soc. 35 (1977), 113-135. [128] P. McMullen, Continuous translation invariant valuations on the space of compact convex sets, Arch. Math. 34 (1980), 377-384. [129] M. Meyer, Une characterisation volumique de certains ~spaces norm~s, Israel J. Math. 55 (1986), 317326.
Euclidean structure in finite dimensional n o r m e d spaces
777
[130] M. Meyer and A. Pajor, On Santal6's inequality, Lecture Notes in Math. 1376, Springer, Berlin (1989), 261-263. [131] M. Meyer and A. Pajor, On the Blaschke-Santal6 inequality, Arch. Math. 55 (1990), 82-93. [ 132] V.D. Milman, New proof of the theorem of Dvoretzky on sections of convex bodies, Funct. Anal. Appl. 5 (1971), 28-37. [133] V.D. Milman, Geometrical inequalities and mixed volumes in the local theory of Banach spaces, Ast6risque 131 (1985), 373-400. [134] V.D. Milman, Random subspaces of proportional dimension of finite dimensional normed spaces: approach through the isoperimetric inequality, Lecture Notes in Math. 1166 (1985), 106-115. [ 135] V.D. Milman, Almost Euclidean quotient spaces of subspaces offinite dimensional normed spaces, Proc. Amer. Math. Soc. 94 (1985), 445-449. [136] V.D. Milman, Inegalit~ de Brunn-Minkowski inverse et applications ?~la th~orie locale des espaces norm~s, C. R. Acad. Sci. Paris 302 (1986), 25-28. [137] V.D. Milman, The concentration phenomenon and linear structure of finite-dimensional normed spaces, Proc. ICM, Berkeley (1986), 961-975. [138] V.D. Milman, A few observations on the connection between local theory and some other fields, Lecture Notes in Math. 1317 (1988), 283-289. [139] V.D. Milman, Isomorphic symmetrization and geometric inequalities, Lecture Notes in Math. 1317 (1988), 107-131. [140] V.D. Milman, A note on a low M*-estimate, Geometry of Banach Spaces, Proceedings of a Conference held in Strobl, Austria, 1989, EF. Muller and W. Schachermayer, eds, LMS Lecture Note Series 158, Cambridge University Press (1990), 219-229. [141] V.D. Milman, Spectrum of a position of a convex body and linear duality relations, Israel Math. Conf. Proceedings 3, Festschrift in Honor of Professor I. Piatetski-Shapiro, Weizmann Science Press of Israel (1990), 151-162. [142] V.D. Milman, Some applications of duality relations, Lecture Notes in Math. 1469 (1991), 13-40. [143] V.D. Milman, Dvoretzky's theorem- Thirty years later, Geom. Funct. Anal. 2 (1992), 455-479. [144] V.D. Milman, Isomorphic Euclidean regularization of quasi-norms in •n, C. R. Acad. Sci. Paris 321 (1995), 879-884. [145] V.D. Milman, Proportional quotients offinite dimensional normed spaces, Linear and Complex Analysis, Problem book 3, V.E Havin and N.K. Nikolski, eds, Lecture Notes in Math. 1573 (1994), 3-5. [146] V.D. Milman and A. Pajor, Cas limites dans les in~galit~s du type de Khinchine et applications g~om~triques, C. R. Acad. Sci. Paris 308 (1989), 91-96. [ 147] V.D. Milman and A. Pajor, Isotropic position and inertia ellipsoids and zonoids of the unit ball ofa normed n-dimensional space, Lecture Notes in Math. 1376, Springer, Berlin (1989), 64-104. [148] V.D. Milman and A. Pajor, Entropy and asymptotic geometry ofnonsymmetric convex bodies, Adv. Math. 152 (2000), 314-335. [149] V.D. Milman and G. Pisier, Banach spaces with a weak cotype 2 property, Israel J. Math. 54 (1986), 139-158. [150] V.D. Milman and G. Schechtman, Asymptotic Theory of Finite Dimensional Normed Spaces, Lecture Notes in Math. 1200, Springer, Berlin (1986). [151] V.D. Milman and G. Schechtman, An "isomorphic" version of Dvoretzky's theorem, C. R. Acad. Sci. Paris 321 (1995), 541-544. [152] V.D. Milman and G. Schechtman, Global versus Local asymptotic theories of finite-dimensional normed spaces, Duke Math. J. 90 (1997), 73-93. [ 153] V.D. Milman and H. Wolfson, Minkowski spaces with extremal distance from Euclidean spaces, Israel J. Math. 29 (1978), 113-130. [154] V.D. Milman and H. Wolfson, Topics in Finite Metric Spaces, GAFA Seminar Notes, Tel Aviv University (1984). [155] A. Pajor and N. Tomczak-Jaegermann, Remarques sur les nombres d'entropie d'un op~rateur et de son transpose, C. R. Acad. Sci. Paris 301 (1985), 743-746. [156] A. Pajor and N. Tomczak-Jaegermann, Subspaces of small codimension of finite dimensional Banach spaces, Proc. Amer. Math. Soc. 97 (1986), 637-642.
778
A.A. Giannopoulos and V.D. Milman
[157] O. Palmon, The only convex body with extremal distance from the ball is the simplex, Israel J. Math. 80 (1992), 337-349. [158] G. Paouris, On the isotropic constant of nonsymmetric convex bodies, Lecture Notes in Math. 1745 (to appear). [159] C.M. Petty, Surface area of a convex body under affine transformations, Proc. Amer. Math. Soc. 12 (1961), 824-828. [ 160] G. Pisier, Sur les espaces de Banach de dimension finie a distance extremale d'un espace euclidien, S6minaire d'Analyse Fonctionelle (1978-79). [161] G. Pisier, Holomorphic semi-groups and the geometry of Banach spaces, Ann. of Math. 115 (1982), 375392. [162] G. Pisier, On the dimension of the s of Banach spaces, for 1 ~< p < 2, Trans. Amer. Math. Soc. 276 (1983), 201-211. [163] G. Pisier, Probabilistic methods in the geometry of Banach spaces, Lecture Notes in Math. 1206 (1986), 167-241. [164] G. Pisier, The Volume of Convex Bodies and Banach Space Geometry, Cambridge Tracts in Math. 94 (1989). [ 165] S. Reisner, Zonoids with minimal volume product, Math. Z. 192 (1986), 339-346. [166] S. Rolewicz, Metric Linear Spaces, Monografie Matematyczne 56, PWN-Polish Scientific Publishers, Warsaw (1972). [167] S. Rosset, Normalized symmetric functions, Newton's inequalities, and a new set of Stringer inequalities, Amer. Math. Monthly 96 (1989), 815-819. [168] M. Rudelson, Estimates on the weak distance between finite-dimensional Banach spaces, Israel J. Math. 89 (1995), 189-204. [169] M. Rudelson, Contact points ofconvex bodies, Israel J. Math. 101 (1997), 93-124. [ 170] M. Rudelson, Distances between nonsymmetric convex bodies and the MM*-estimate, Positivity 4 (2000), 161-178. [171] J. Saint Raymond, Sur le volume des corps convexes sym~triques, Sem. d'Initiation h l'Analyse, no. 11 (1980-81). [ 172] G. Schechtman, L~vy type inequality for a class ofmetric spaces, Martingale Theory in Harmonic Analysis and Banach Spaces, Springer-Verlag, Berlin (1981), 211-215. [173] G. Schechtman, More on embedding subspaces of Lp in s Comp. Math. 61 (1987), 159-170. [174] G. Schechtman, A remark concerning the dependence on e in Dvoretzky 's theorem, Lecture Notes in Math. 1376 (1989), 274-277. [175] G. Schechtman, Concentration results and applications, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [176] E. Schmidt, Die Brunn-Minkowski Ungleichung, Math. Nachr. 1 (1948), 81-157. [177] R. Schneider, Convex Bodies: The Brunn-Minkowski Theory, Encyclopedia of Mathematics and its Applications, Vol. 44, Cambridge University Press, Cambridge (1993). [178] R. Schneider, Simple valuations on convex sets, Mathematika 43 (1996), 32-39. [ 179] V.N. Sudakov, Gaussian random processes and measures of solid angles in Hilbert spaces, Soviet Math. Dokl. 12 (1971), 412-415. [ 180] A. Szankowski, On Dvoretzky's theorem on almost spherical sections of convex bodies, Israel J. Math. 17 (1974), 325-338. [181] S.J. Szarek, On Kashin's almost Euclidean orthogonal decomposition of s Bull. Acad. Polon. Sci. 26 (1978), 691-694. [ 182] S.J. Szarek, The finite dimensional basis problem, with an appendix on nets of Grassmann manifold, Acta Math. 159 (1983), 153-179. [183] S.J. Szarek, Spaces with large distance to s and random matrices, Amer. J. Math. 112 (1990), 899-942. [184] S.J. Szarek, On the geometry of the Banach-Mazur compactum, Lecture Notes in Math. 1470 (1991), 48-59. [ 185] S.J. Szarek, Computing summing norms and type constants on few vectors, Studia Math. 98 (1991), 147156. [ 186] S.J. Szarek and M. Talagrand, An isomorphic version of the Sauer-Shelah lemma and the Banach-Mazur distance to the cube, Lecture Notes in Math. 1376 (1989), 105-112.
Euclidean structure in finite dimensional normed spaces
779
[187] S.J. Szarek and N. Tomczak-Jaegermann, On nearly Euclidean decompositions of some classes of Banach spaces, Compositio Math. 40 (1980), 367-385. [188] M. Talagrand, Embedding subspaces of L! into s Proc. Amer. Math. Soc. 108 (1990), 363-369. [189] M. Talagrand, Sections of smooth convex bodies via majorizing measures, Acta Math. 175 (1995), 273300. [190] N. Tomczak-Jaegermann, The Banach-Mazur distance between the trace classes C~Z~,Proc. Amer. Math. Soc. 72 (1978), 305-308. [191] N. Tomczak-Jaegermann, Computing 2-summing norm with few vectors, Ark. Mat. 17 (1979), 273-277. [192] N. Tomczak-Jaegermann, The Banach-Mazur distance between symmetric spaces, Israel J. Math. 46 (1983), 40-66. [193] N. Tomczak-Jaegermann, The weak distance between Banach spaces, Math. Nachr. 119 (1984), 291-307. [194] N. Tomczak-Jaegermann, Dualit~ des nombres d'entropie pour des op~rateurs gt valeurs dans un espace de Hilbert, C. R. Acad. Sci. Paris 305 (1987), 299-301. [195] N. Tomczak-Jaegermann, Banach-Mazur Distances and Finite Dimensional Operator Ideals, Pitman Monographs, Vol. 38, Pitman, London (1989). [196] G. Wagner, On a new method for constructing good point sets on spheres, Discrete Comput. Geom. 9 (1993), 111-129.
This Page Intentionally Left Blank
CHAPTER
18
Renormings of Banach Spaces Gilles Godefroy Equipe d'Analyse, Universite Paris VI, 4, place Jussieu, F-75252 Paris cedex 05, France E-mail: [email protected]
Contents 1. Definitions and basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Separability of the dual space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Renormings of super-reflexive spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Smoothness of higher order . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Characterizing spaces by renormings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Some applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
H A N D B O O K OF THE G E O M E T R Y OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 781
784 794 800 806 814 822 829 830
This Page Intentionally Left Blank
Renormings of Banach spaces
783
Renorming a Banach space consists of replacing the given norm, which is usually provided by the very definition of the space, by another norm which may have better (or sometimes worse) properties of convexity or smoothness, or both. This operation is by nature geometric. We replace the original unit ball by another one, which has a different shape. In fact, geometric intuition is frequently useful and a few pictures might help in understanding certain proofs. However, the analytical point of view is of course necessary and proofs usually rely on careful computations. Computing in Banach spaces is quite difficult, since certain spaces have no basis and therefore there is no hope to use coordinates in calculations. Most results in renorming theory rely on topological assumptions. Since bases or substitutes for bases are not always available, computations have to be global, that is, to rely on coordinate-free functions such as, e.g., distances to convex sets. It turns out that such a "global calculus" suffices most of the time for constructing special norms with desirable properties. The fact that many important isomorphism classes admit natural characterizations in terms of existence of certain equivalent norms illustrates this point. Norms which enjoy good properties of convexity and/or smoothness can be computed under natural (and optimal) topological assumptions. These norms are frequently obtained by duality arguments and they are tightly connected with the linear structure. However, renormings can be used for constructing all kinds of functions on Banach spaces, by processing equivalent norms through the usual operations of analysis: linear combinations, suprema and infima, composition with real functions. Therefore they create a bridge between linear and nonlinear theory and lead in particular to a better understanding of Banach spaces considered as infinite-dimensional smooth manifolds, or as metric spaces. Finally, renorming theory is by its very definition an intermediate topic between isomorphic and isometric theory of Banach spaces. Connections work both ways. Sometimes an isomorphic assumption leads to a canonical construction of special norms, which in turn provides information on the isomorphism class under consideration. Sometimes the original norm itself, although an isometric object, informs us on the isomorphic properties of the space. Let us outline the contents of this article. Unless otherwise specified, the spaces we consider are real Banach spaces. Most of the results and proofs we display concern separable spaces. Nonseparable theory leads to delicate problems involving in particular infinite combinatorics. We refer to V. Zizler's article [ 176] "Nonseparable Banach spaces" in this Handbook for these topics. Section 1 presents the basic convexity and smoothness properties norms can have as well as the main duality results between convexity and smoothness. Some simple but important renormings of separable spaces are also shown there. In Section 2 one relates, through renormings, the existence of"nontrivial" differentiable functions on a separable space with the separability of its dual space. Variational principles are applied to subdifferentiability, when smoothness is available, while harmonic behaviour of smooth functions is displayed in nonsmooth spaces. Cantor derivations in dual spaces are an operative tool in Section 2, and in Section 3 as well, where they provide an alternative approach to uniformly convex renormings of super-reflexive spaces. This allows for dispensing with the classical martingale approach but still provides quantitative results in a canonical way. Section 4 deals with smoothness of higher order. The gist of this section is that while C 1 smoothness is available under topological assumptions, C 2 smoothness leads to much more restrictive conditions of a quantitative nature, while C ~ smoothness
784
G. Godefroy
of norms forces the containment of classical sequence spaces. Section 5 displays several important isomorphism classes of separable Banach spaces which allow for simple characterizations in terms of renormings: Hilbert spaces, reflexive spaces, spaces containing I 1(N), subspaces of c0(N) and order-continuous lattices. Finally, Section 6 provides various applications of renormings: uniform approximation of continuous functions by differentiable ones, of real-valued Lipschitz functions by differences of convex functions on super-reflexive spaces, linearization of Lipschitz isomorphisms in subspaces of c0(N) and weak sequential completeness of extremely rough spaces. It should be stressed that this chapter is by no means an exhaustive work on renormings, even for separable spaces. It is rather an attempt to provide the reader with various relevant techniques and with motivations for the whole theory, in an accessible and when possible nontechnical way. Needless to say, choices had to be made and the selection of topics has been greatly influenced by the author's own taste. The sections themselves contain no references. The results are sometimes, but not always, the optimal available results. The proofs often consist of a sketch, where several computations are simply outlined. Experts might consider them as "complete" proofs, where only technical details should be added, while less experienced readers should find them helpful for understanding the proofs which are given in the referenced articles. Each section is complemented with an extensive "Notes and comments" paragraph, where references are given, related results and applications are provided, and open problems are mentioned. A bibliography section concludes the chapter. The notation and the terminology we use are classical and can be found in [130]. We should however point out the unfortunate fact that two terminologies are currently used for designating important convexity properties. We choose to use in this chapter the word "convex" instead of "rotund". So for instance, the meaning of the sentence "locally uniformly convex" is identical with the meaning of "locally uniformly rotund" in related works. Also, "smooth" and "differentiable" have the same meaning throughout this work. We refer to [ 104] in this Handbook for basic concepts and results used in this chapter.
1. Definitions and basic properties Let us recall basic definitions of special convexity properties norms can have. DEFINITION 1.1. Let X be a Banach space, and II 9 II be a norm on X. (i) The norm II 9 II is strictly convex if whenever (x, y) E X 2 are such that 2(llx II2 + IlYll2) = IIx -+- Yll2, one has x = y. (ii) The norm II 9 I1 is locally uniformly convex (1.u.c.) if whenever x E X and the sequence (Xn) C X are such that lim2(llx II2 -+-IlXn II2) --IIx + Xnll2 --0, then lim Ilx - xn II = O.
Renormings of Banach spaces
785
(iii) The norm II 9 II is uniformly convex (u.c.) if whenever (Xn) C X and (Yn) C X are bounded sequences in X such that lim2(llxn II2 -+- IlYn II2) - Ilxn -t- Yn II2 -- 0, then lim Ilxn - Yn II = 0. It is obvious from the definition that a uniformly convex norm is locally uniformly convex and that a locally uniformly convex norm is strictly convex. These notions can be alternatively defined in a n o n h o m o g e n e o u s manner: a norm is strictly convex if and only if the unit sphere Sx does not contain any nondegenerate segment. It is locally uniformly convex if the length of a segment joining any given point of Sx to another point of Sx is controlled by the norm of the midpoint. It is uniformly convex if the length of a segment joining two points of Sx is uniformly controlled by the norm of the midpoint. This last characterization naturally leads to the modulus of uniform convexity, defined for e E [0, 2] by
8x(e) -- inf{ 1 -
x+y) 2
; (x, y)
9
(Sx) 2 IIx - yll
~ e]
It is clear that II 9 II is uniformly convex if and only if 8x(e) > 0 for every e > 0. EXAMPLE 1.2. The natural norm of the Hilbert space 12 is uniformly convex, as shown by the parallelogram identity 2(llx II2 + Ilyll 2) - I I x -+- yll 2 -+-IIx - yll 2, In fact, this identity shows that
8/2 ( e ) - -
1 - (1
c2
-
-T) 1/2.
Uniformly convex spaces enjoy a simply shown but fundamental property. PROPOSITION 1.3. Any uniformly convex space X is super-reflexive. PROOF. We first show that X is reflexive. Pick t E X** with Iltll = 1. By Goldstine's theorem, the unit ball Bx of X is weak* dense in the bidual unit ball Bx**. Pick e > 0. Since II 9 II is uniformly convex, there exists 8 > 0 such that if (x, y) E (Sx) 2 are such that IIx + y II > 2(1 - 8), then IIx - y II < e. Since the bidual norm on X** is weak* lower semicontinuous, there is a convex neighbourhood V of t in (Bx**, w*) such that Ilull > 1 - 8 for all u E V. It follows that the II 9II-diameter of the nonempty set Bx M V is less than e. Since t belongs to the weak* closure of that set, we have by weak* lower semi-continuity of the norm that II 9 II - dist(t, X) ~< e. Since e is arbitrary, it follows that t E X, hence X** = X. To conclude the proof, we observe that in the above notation, the same 8 > 0 works for the same e and for every space Y which is finitely representable in X; this is indeed clear
G. Godefroy
786
since the inequalities involve only 2-dimensional subspaces. It follows that any such Y is reflexive, hence X is super-reflexive. Vq It turns out that the converse to Proposition 1.3. holds true, up to a renorming (see Theorem 3.2. below). This proposition shows that separable spaces do not in general admit equivalent uniformly convex norms. However, a positive result holds true if we are ready to drop the uniformity. The proof will allow us to formulate a dual version as well. PROPOSITION 1.4. Any separable space X has an equivalent locally uniformly convex norm. Any space with separable dual has an equivalent norm whose dual norm is locally uniformly convex. PROOF. We first show that X has an equivalent strictly convex norm. Indeed, let (fn)n/>l be a sequence in Bx, which separates X; such a sequence exists since, e.g., the compact space (Bx,, w*) is metrizable and therefore separable. If N is the original norm on X, we define a new norm I . I on X by oo
Ix[2 -- N ( x ) 2 -+- Z
2-n fn (X) 2
n=l
It follows easily from the separating property of the sequence (fn) and the strict convexity of the real valued function g(x) = x 2 that [. [ is an equivalent strictly convex norm. Let n o w (Xn)n>/Obe an increasing sequence of finite dimensional subspaces of X whose union is dense in X, with X0 = {0}. Let I 9[ be a strictly convex norm on X. For any n ~> 0, we define dn on X by
dn(x) -- I 9I - dist(x, Xn), and we define II 9 II on X by the formula O43
Ilxll 2 - ~
2-ndn(x) 2
n=0
It is easily checked that 1[ . 11 is an equivalent norm on X. We pick x E X and a sequence (xk) C X such that 2(llx II2 + Ilxk II2) - l i m IIx + x~ll 2 - O.
(1)
Since the seminorms dn are convex, the corresponding expressions obtained by substituting II l[ by dn in (1) are all positive. Therefore it follows from (1) that for every n ~> 0, lim 2(dn(x) 2 -+-dn(Xk) 2) - dn(x + Xk) 2 - - 0 , k
Renormings of Banach spaces
787
and thus lim dn (xk) = dn (x)
for every n ~> 1.
k
Pick now e > 0. There is n such that dn(x) < e, hence dn(xk) < e for k large enough. Since do(xk) is bounded, the sequence (xk) is bounded. Since Xn is finite dimensional, its bounded subsets are compact and it follows that (xk) has a subsequence which is (3e)Cauchy. Since ~ was arbitrary, a diagonalization provides a Cauchy subsequence. It follows easily that the original sequence (xk) itself is convergent to a limit y. But since (1) holds true with do, we have Ix + y l 2 - 2 ( I x l 2 + lYl 2) and since the norm I 9I is strictly convex, it follows that x = y, which shows that II 9 II is locally uniformly convex. To conclude the proof, we now investigate the dual case. A first observation is that when X = Y* is a separable dual, the separating sequence (fn) can be taken within By. Hence the functionals which define I 9I are all lower semi-continuous (1.s.c.) with respect to the pointwise convergence on Y, provided that the original norm N is a dual norm, and it follows that I 9[ itself is 1.s.c. with respect to the weak* topology. Hence I 9I is dual to a norm on Y, as can be checked by the bipolar theorem. Then we observe that since I 9I is weak* 1.s.c. and the spaces Xn are finite dimensional, the seminorms dn are weak* 1.s.c. and therefore so is II 9 II. Hence the norm we construct in the dual case is a dual norm. E] REMARKS 1.5. (1) Note that this proof shows in particular that if Y is separable, there is an equivalent norm on Y whose dual norm is strictly convex. Indeed, if Y is separable and the separating sequence (fn) is contained in Y, the norm I 9I is a strictly convex dual norm. (2) Theorem 1.4 relies on the consideration of a series of convex functionals; in this case, of distances from an increasing sequence of finite dimensional subspaces. Similar series, which increase convexity, turn out to be a very efficient tool in renorming theory (see, e.g., the proofs of Theorems 3.2 and 5.4 below). They also allow for transferring locally uniformly convex norms from a space to another, which is useful in the nonseparable theory. For instance, if Y is a dual space with a dual 1.u.c. norm [I 9II and T is a weak* to weak continuous linear map from Y to X with dense range, we define a sequence of equivalent norms [ 9In on X by the formula I x l ~ - inf{ Ilx - Zyll 2 + n -1 [yl2; y 6 Y} and an equivalent norm N by oo
N(x)-Z2-nIxl2. n--1
This norm N is an equivalent locally uniformly convex norm on X. Combining such arguments with "Cantor derivation" arguments (see the proofs of Proposition 2.6 and Theo-
G. Godefroy
788
rem 3.2) provides renorming theorems through a global approach, which does not refer to coordinates. Norms and convex functions on Banach spaces can enjoy two distinct fundamental properties of smoothness, which we now define. DEFINITION 1.6. A function F defined from a Banach space X into a Banach space Y is Fr6chet differentiable at x E X if there is a continuous linear map T from X into Y such that F ( x + h) = F ( x ) + T ( h ) + o(h), where limllhll__,0 IIo(h)ll/llhll - O. A norm II 9II on X is said to be GSteaux differentiable (or in short, G-smooth) at x ~ S x if there exists f E S x , such that for all y ~ S x ,
Ilxll]
lim t-1 [llx -q- t y l l t--+0
-
f(y).
(1)
If moreover the limit in (1) is uniform on y E S x , then the norm is said to be Fr6chet differentiable (or in short, F-smooth) at x. Observe that in the notation of (1) one has f (x) = 1. The linear form f is called the G~teaux or Fr6chet (in short, G- or F-) derivative of II 9 II at x. Obviously a norm is F - s m o o t h if and only if it is Fr6chet differentiable as a function from X\{0} to R. Hence, when a norm or a function is simply called differentiable, it will always be meant in the Fr6chet sense. Gfiteaux smoothness of convex functions refers to directional differentiability. By homogeneity, and since obviously no norm is smooth at 0, smoothness of norms is usually considered only on the unit sphere. We will usually say that a norm is G - s m o o t h or F - s m o o t h if it is so outside {0}. EXAMPLE 1.7. (1) The parallelogram identity and the chain rule immediately imply that the natural norm of the Hilbert space 12 is F-smooth. (2) Let l l be the space of all absolutely convergent series, equipped with its natural norm I[ 9Ill, that is, if x = (xi)i >>.1, oo
Ilxlll-~_~lxil. i=1
This norm is G-smooth at x
=
(xi) E Sll if and only if Xi ~ 0 for all i. Indeed, if Xi = 0
then clearly the norm is not smooth in the direction of ei, with e i ( j ) - 6/. On the other hand, if xi =/=0 for all i and h = (hi) E 11, we have that
l i m t -1 t--+O
[xi -t- thil - Ixi[ - sign(xi)thi i--1
Indeed, if i is such that Ithil ~ Ixi I, then
Ixi -t- thi[ -
Ixil = sign(xi)thi.
-0.
(1)
789
Renormings of Banach spaces
If J denotes the set of all its such that Ithil > Ix/I, then
~l
Ixi-+-thil - Ixil -sign(xi)thil ~ 4 ~ l t h i l .
i~J
ieJ
It follows that the norm is smooth in the direction h. We now claim that the natural norm of 11 is nowhere F-smooth. Indeed, if it were F - s m o o t h at x = (xi), its F-derivative would coincide with its G-derivative, i.e., with the linear form f ( h ) - Z i ~ = l sign(xi)hi. To check that the limit in (1) is not uniform in h 9 Sll, it suffices to consider the sequence en defined above. If we substitute h = en and tn = - 2 x n in (1), the left-hand side of the equation equals -2.sign(xn), and this contradicts uniformity since lim tn = 0. On the infinite-dimensional Banach space l l, we thus have a continuous (even Lipschitz) function, namely II 9 II1, which is nowhere F-differentiable. We shall see later that the existence of such a function defined on a Banach space X is closely related to the structure of X. DEFINITION 1.8. A norm II 9 II on a Banach space X is uniformly F - s m o o t h (in short, UF-smooth) if the limit limt~0 t-l[llx + thll- Ilxll] exists uniformly on (x, h) 9 ( S x ) 2. Let us mention that there is a weaker notion of uniform smoothness, which is called uniform G~teaux-smoothness. A norm II 9 II is uniformly G~teaux-smooth (in short, UG-smooth) if for every h 9 Sx, the limit in Definition 1.8. exists uniformly in x 9 Sx. This notion is somehow trivial in separable spaces since any separable space has an equivalent uniformly G~teaux-smooth norm. The nonseparable theory, for which we refer to [176], is in this respect much more interesting. The modulus of smoothness p x ( r ) is defined for any r > 0 by the formula
px(r)
--supJ/ IIx + rYll +
2
IIx - rYll
- 1- Ilxll = Ilyll-
1}.
It is easily checked that II 9II is UF-smooth if and only if lim
r--+0
px(r) "g
= 0.
EXAMPLE 1.9. The space 12 is UF-smooth. Indeed, since ]Ix + h[[2 - ((x -k- h,x + h)) 1/2, it follows that (1 +
_ 1.
There is a natural duality between convexity and smoothness. Let us first consider the superreflexive case, where this duality can be expressed in quantitative form. The modulus of convexity of X and the modulus of smoothness of X* equipped with the dual norm are related through the following formula, which follows from Fenchel duality.
G. Godefroy
790
THEOREM 1.10. One has p x , (r) = s u p { r e / 2 - 6x(e); 0 <<.e <~2}. F r o m this formula and the characterizations of UC- and U F - s m o o t h norms by their moduli, it follows that we have COROLLARY 1.1 1. A Banach space is uniformly smooth if and only if its dual X* is uni-
formly convex. It now follows from Proposition 1.3 that we have COROLLARY 1.12. Any uniformly smooth Banach space is super-reflexive. We now investigate dual characterizations of F - or G-smoothness when these properties hold without uniformity. In this case we have to expect only a partial duality: indeed such properties do not imply reflexivity, and therefore a distinction occurs between the linear forms which attain their norms and those which don't. We will use extremal structure of convex sets, through the following notions. DEFINITION 1.13. Let C be a closed b o u n d e d convex subset of a Banach space Y. A point x E C is exposed in C by f E Y* if f ( x ) = s u p { f ( t ) ; t E C} and if t E C and f ( t ) = f ( x ) implies that t = x. If, moreover, there exists for any ~ > 0 some e > 0 such that IIt - x II < as soon as t E C and f (t) > f (x) - e, then x is said to be strongly exposed by f in C. We now state and prove a fundamental duality lemma, which is a pointwise version of T h e o r e m 1.10. In what follows, we identify X with a subspace of X**, and X* = Y in the notation of Definition 1.13. LEMMA 1.14. Let X be a Banach space. Pick x E Sx and f ~ Sx, such that f (x) = 1.
The norm of X is G~teaux (resp. Frdchet) smooth at x if and only if f is exposed (resp. strongly exposed) by x in Bx,. PROOF. A s s u m e that II 9 II is G - s m o o t h at x, and pick f as above. Since Ilfll = 1, we have IlYll/> f ( Y ) for all y E X. It follows that f ( h ) = l i m t ~ 0 t - l [ l l x + t h l l - Ilxll] for all h 6 X, which uniquely determines f . Hence f is exposed in B x , by x. If m o r e o v e r II 9 II is F - s m o o t h at x, then for any 3 > 0 there is e > 0 such that
IIx -+- h ll-+-IIx - h ll ~ 2 + ~ IIh II
whenever Ilhll < e.
For any g E B x, and all h E X, one has
f (x + h) + g(x - h) <<,IIx + h II + IIx - h II. Thus for all h E X with IIh II < e one has
6 ( f - g)(h) <~ 1 - g(x) + ~llhll. z
Renormings of Banach spaces
791
Hence if g is such that g(x) > 1 - 8e/2, then IIf - gl[ ~ 8. For showing the converse, we observe that limt~0+ t -1 [llx + thll - Ilxll], which exists by convexity, is a subadditive function of h, while the corresponding limit when t --+ 0 is superadditive. It follows that the n o r m is G - s m o o t h if and only if these two limits are equal for all h E X. Hence if the n o r m is not G - s m o o t h at x, there is an h for which these two limits are different. It follows that there exist two distinct linear forms g and 1 on the 2-dimensional space generated by x and h with norm one such that g(x) = 1(x) = 1. These two linear forms can be lifted to two distinct elements of B x , by the H a h n - B a n a c h theorem. This proves the converse in the G - s m o o t h case. Finally, if F - s m o o t h n e s s fails, there is ~ > 0 and a sequence hn with lim IIh~ II = 0 and hn r 0 such that
IIx + hn II + IIx - hn II ~ 2 + 6 Ilhn IILet fn and gn in B x , be such that f n ( x + h n ) = IIx -+-h~ II and g n ( x - h n ) = IIx - h n l l . We have l i m f n ( x ) = l i m g n ( x ) = 1 and ( f , - g , ) ( h , ) >~ ~llh~ll, hence IIf~ - g~ll /> ~. This concludes the proof. D REMARK 1.15. A simple but important application of L e m m a 1.14 is that any F - s m o o t h n o r m is actually C 1. This extends, with similar proof, to any convex F-differentiable function, which is necessarily C 1. On the other hand, G - s m o o t h n e s s provides natural examples of differentiable functions (although in a weaker sense) whose derivative is not continuous. Indeed the mean value theorem implies that the continuity at x of the G-derivative implies F - s m o o t h n e s s at x.
COROLLARY 1.16. Let X be a Banach space. I f the dual norm is strictly convex, then the norm is Ggtteaux smooth on X. I f the dual norm is locally uniformly convex, then the norm is Frdchet smooth on X. PROOF. Pick x 6 Sx. If (f, g) E ( S x , ) 2 are such that f ( x ) = g(x) = 1, then IIf -+- gll = 2, and thus f = g if the dual n o r m is strictly convex. If this n o r m is m o r e o v e r locally uniformly convex and (f, g) ~ ( B x , ) 2 are such that f (x) = 1 and g(x) > 1 - e, then 2(llfll 2 -+-Ilgll 2) - I I f
+ gll 2 < 4e.
This implies IIf - g II < ~ if e > 0 is small enough. L e m m a 1.14 concludes the proof.
D
REMARK 1.17. The converse implications do not hold in Corollary 1.16. For instance, there exists on any nonreflexive space with separable dual an equivalent F - s m o o t h norm whose dual n o r m is not strictly convex. The fundamental result for smooth renorming of separable spaces now reads as follows.
PROPOSITION 1.18. Every separable Banach space X has an equivalent Ggtteaux smooth norm. I f moreover X* is separable, then X has an equivalent Fr~chet smooth norm.
792
G. Godefroy
PROOF. By Remark 1.5, if X is separable there is an equivalent norm on X whose dual norm is strictly convex. By Corollary 1.16, this norm is G-smooth. If X* is separable, there is by Proposition 1.4 a norm on X whose dual norm is locally uniformly convex on X*, and by Corollary 1.16, this norm is F-smooth. [2
Notes and comments. The basic definitions which are recalled here can be found in [36]. More recent references include [48,46] and [90]. The notion of locally uniformly convex norm is due to Lovaglia [ 131 ]. The parallelogram identity has been known for more than two thousand years. Its interpretation in terms of uniformly convex norms is however more recent. Reflexivity of uniformly convex spaces was shown independently in [137] and [ 145], the present proof is due to J. Ringrose [48, cf. p. 49]. Super-reflexivity has been defined by R.C. James, who proved a deep theorem which extends Proposition 1.3 [99]: if a Banach space X is such that 8x(2 - e) > 0 for some e > 0, then X is superreflexive. In other words, X is superreflexive as soon as the space R 2 equipped with the I 1 norm is not finitely representable in X. Locally uniformly convex equivalent norms on separable spaces have been constructed by Kadets [ 110]. We follow a modification of his approach here [35]. Kadets' purpose was to show that all separable infinite dimensional Banach spaces are homeomorphic ([111]; see [17] or [90]). This remarkable theorem does not close the subject, since uniform or Lipschitz isomorphisms lead to nontrivial isomorphism classes which are sometimes, but not always, distinct from the linear ones (see [106] and [78]). Dual locally uniformly convex renormings of separable duals are shown in [ 112] and independently in [8]. The same approach works for renorming a dual space in a dual way such that the norm is locally uniformly convex on a prescribed separable subspace. This permits to renorm any nonreflexive space in such a way that it is not isometric to a dual space [35]. A variation of this technique was used in [67] to show that the approximation property does not imply the bounded approximation property. A different approach, which relies on "strongly bicontractive" renormings of spaces with a shrinking basis, provides a renorming of every space with separable dual by a norm which has some kind of "uniform smoothness": more precisely, no proper subspace of the dual space is )~-norming, where )~ > 1/2 is given [81]. When such a space has the 00-1-B.A.E, its dual space has the metric approximation property. Remark 1.5 presents what is known as the transfer method, which originates in [74] and has been subsequently developed in a series of papers [83,84,57,139]. Roughly speaking, the idea consists of considering a hereditary property (P) of norms, and to check whether (P) can be "transferred" from X to Y when there is a linear operator T :Y --+ X such that T** is one-to-one. This technique is useful in the nonseparable theory, and we refer to V. Zizler's article [ 176] in this Handbook for some of its applications. Uniform Fr6chet smoothness leads to super-reflexivity (see, e.g., [12] or [90]) and is by now well-understood, but uniform G~teaux smoothness (UG) is much more delicate to handle. It is known that any space with a UG-smooth norm is a subspace of a weakly compactly generated space [60], see [ 176] for the definition of weakly compactly generated spaces. The duality formula (Theorem 1.10) which relates the moduli of convexity and smoothness is due to Lindenstrauss [128]. Several moduli, which refer to weaker "asymptotic" notions are defined in [138]. They are relevant to appropriate extensions of uniform con-
Renormings of Banach spaces
793
vexity or smoothness (in a weaker, asymptotic sense) to certain nonreflexive spaces, such as Theorem 5.7 below. These notions have recently been applied to nonlinear theory [78, 107], following the pioneering work of Heinrich and Mankiewicz [96] showing that superreflexivity as well as moduli of convexity and smoothness of equivalent renormings are stable under uniform homeomorphisms. A complex analogue of uniform convexity is studied in [33], under the terminology "PL-convex". An interesting point is that many natural nonreflexive complex spaces enjoy this property, such as duals of C*-algebras, which turn out to be 2-uniformly PL-convex. This result due to Haagerup (see [33]) improves on a previous work of [168]. We refer to [148, Chapter 9] for applications of this notion to the noncommutative version of Grothendieck's theorem. Extensions of given norms from a subspace to the whole space, while keeping special properties of the given norm, is usually possible when qualitative or quantitative rotundity properties are concerned [59,167]. Smoothness properties behave differently, and in fact G-smoothness cannot be preserved even in the separable case (see [46, Section 2.8]). It is not known whether a space, such that every G-smooth norm can be extended in a G-smooth way to an equivalent norm on any separable super-space, is isomorphic to c0(N). It is shown in [142] (cf., e.g., [48]) that the Hilbert space has a larger modulus of convexity and a smaller modulus of smoothness than any normed space. The exact value of the modulus of convexity (and by duality of the modulus of smoothness) is actually an isometric characterization of the Hilbert space 12. In fact, the values of the modulus on certain sequences suffice for characterizing the Hilbert space: it is shown in [4] that if 32 denotes the modulus of convexity of 12 and if ~x(e) = ~2(e) for some e > 0 such that e/2 is not the sine of an even divisor of zr, then X is isometric to a Hilbert space. We refer to [5] for many isometric characterizations of the Hilbert space, and to [ 13] for a related very versatile "modulus for all seasons". Extremal structure of convex sets is studied for quite a long time and the literature on this topic is huge. We refer to the chapter "infinite dimensional convexity" by V. Fonf, J. Lindenstrauss and R.R. Phelps in this Handbook [69] and to [32] for integral representation theory, which refines the classical Krein-Milman theorem by representing any point in a metrizable compact convex set as a barycenter of a probability measure carried by the extreme points. Weak-star exposed points turn out to be more difficult to handle than extreme points, since even in the metrizable case their topological structure can be quite complicated, e.g., they can form a non-Borel set ([115]; see [38]). Investigating exposition and its strong version in convex sets leads to "dentability" of convex sets and to the Radon-Nikodym property of Banach spaces, for which we refer to [50,24]. The fundamental duality Lemma 1.14 is due to Smulyan [161 ]. The examples mentioned in Remark 1.17 can be found, e.g., in [175] or in [22] where "generic" constructions lead to a precise computation of the topological complexity of the set of counterexamples. Fr6chet smooth norms on spaces with a separable dual have been constructed independently in [109] and [ 116]. Note that it is shown in Section 2 below that the second part of Proposition 1.18 has a converse (Theorem 2.2). Let us quote for the record [46, Theorem 2.7.1 ] which gathers what can be done under separability assumptions: any separable Banach space has an equivalent norm which is simultaneously locally uniformly convex, uniformly convex in every direction and uniformly
794
G. Godefroy
G~teaux smooth. Any Banach space with separable dual has an equivalent norm which is simultaneously locally uniformly convex, weakly uniformly convex, Fr6chet smooth, uniformly G~teaux smooth and whose dual norm is locally uniformly convex.
2. Separability of the dual space We first recall a classical notation. DEFINITION 2.1. Let X be a Banach space. A real-valued function b defined on X is called a bump function if the set supp(b) = {x 6 X; b(x) 7/= 0} is a bounded nonempty subset of X. A real-valued function is called C 1 o n X if it is differentiable (in the sense of Definition 1.6) and if its differential f ' is continuous from X to X* equipped with their norm topologies. Note that by Remark 1.15 a convex F-smooth function is C 1. We state now a fundamental characterization of spaces with separable dual. THEOREM 2.2. Let X be a separable Banach space. The following assertions are equivalent. (i) The dual X* is separable. (ii) There exists a C 1-smooth bump function on X. (iii) There exists an equivalent Frdchet differentiable norm on X. PROOF. (i) :=~ (iii): this was shown in Proposition 1.18. (iii) = , (ii): let b0 be a C ~ smooth bump function from the real line to itself such that supp(b0) C [1, 2]. If II 9II is an equivalent F-smooth norm on X, then b(x) = b0(llxll) is a C 1 smooth bump function on X. (ii) = , (i): let b be a C 1-smooth bump function. Pick any f E X*. We let 7r = b -2 - f . The function ~ takes values in R U {+oo}, it is bounded from below and lower semicontinuous. By Ekeland's variational principle, there is a slight perturbation of ~p which attains its minimum. In other words, for any e > 0, there is xe in X such that 7r (xe + h) 7z(x~)/> ellhll for all h E X. This implies that for all h 6 X and t > 0, one has t -1 [ b - 2 ( x e -+- th) - b - 2 ( x e ) ] ~ f ( h ) - ellhll. Letting t tend to 0, we obtain by the chain rule
l[-2b-3(x~)b'(x~) - f l ~ E and it follows that the linear span of the set D = {b'(x); x E X} is norm dense in X*. Since X is separable and b ' is continuous, D is separable and it follows that X* is separable as well. This concludes the proof of Theorem 2.2. U]
795
Renormings of Banach spaces
A first consequence of this result is a subdifferentiability property of lower semicontinuous functions on spaces with separable dual. We say that a function g is subdifferentiable at x E X if there exists f E X* such that when lim IIh II -- 0, one has
g(x + h) >~g(x) + f (h) + o(llhll). COROLLARY 2.3. Let X be a Banach space such that X* is separable. Every lower semicontinuous function defined on X is subdifferentiable at every point of a dense subset of X. PROOF. Let g : X --+ R be a 1.s.c. function, and U be a n o n e m p t y open subset of X. Since g is 1.s.c., there is a n o n e m p t y open ball B -- B(x0; 5) in U and an a E R such that g (x) > a for all x E B. Let ot = inf{g(x); Ilx - x011 = 5}. T h e o r e m 2.2 shows that there is a C 1smooth bump function b with supp b C B such that/3 = ot - inf8 (g + b) > 0. We define ~p as follows:
~ ( x ) = g + b(x)
if ilx - x01l ~< 6,
~p(x) = + o o
otherwise.
Clearly ~p is 1.s.c. and b o u n d e d below. M o r e o v e r infsc ~p ~> inf8 ~p + / 3 . variational principle (cf., e.g., [44]) provides us with a C 1-smooth bump IIb011~ ~3/2 such that g + b + b0 attains its m i n i m u m on X at some x0, lib011~ ~3/2. Since b + b0 is differentiable at x0, it follows easily that tiable at x0.
Now the smooth function b0 with and x0 E B since g is subdifferenD
Note that since there exists (even if X = R) continuous and nowhere differentiable functions, the dense set constructed in Corollary 2.3 is in general a m e a g e r set. However, for convex functions we have:
COROLLARY 2.4. Let X be a Banach space such that X* is separable. Every continuous convex function is Fr~chet differentiable at every point of a dense G~ subset of X. PROOF. Let g : X --+ R be a continuous convex function. A standard convexity argument shows that g is F - s m o o t h at x E X if and only if
t--+O\ hESx
.al j
It follows that g is F - s m o o t h as soon as - g is subdifferentiable. Now Corollary 2.3. shows that g is F - s m o o t h at every point of a dense set. M o r e o v e r since a continuous convex function is locally Lipschitz, the set Ue consisting of all x E X where the limit in (1) is strictly less than a given e > 0 is open. It follows that the set of points where a convex function is F - s m o o t h is a G~ set. This concludes the proof. D We recall that a Banach space X is called an Asplund space if every continuous convex function on X is F - s m o o t h at every point of a dense G~ subset of X. By Corollary 2.4 and its converse shown below, a separable space is Asplund if and only if its dual is separable. For showing the converse of Corollary 2.4 we need a notion of uniform nonsmoothness.
G. Godefroy
796
DEFINITION 2.5. Let X be a Banach space. A norm II 9 II is rough if there is some e > 0 such that for all x E Sx, lim sup t - l [ l l x -t- thll-+-IIx - t h l l - 2]/> e. t--+O hESx
The computations in E x a m p l e 1.7(2) show that the natural norm of ll is 2-rough. This also follows from Fact 2.7 below. It is clear that a rough norm is nowhere F - s m o o t h . Consequently, such a norm cannot exist on spaces with separable dual. But conversely, we have: PROPOSITION 2.6. Let X be a separable Banach space. The following assertions are equivalent: (i) The dual X* is not separable. (ii) There exists an equivalent rough norm on X. PROOF. (ii) implies (i) is clear by Corollary 2.4. Conversely assume (i). We will consider a "Cantor derivation" on the set of weak* closed subsets of X*. For any e > 0 and any weak* closed subset F of X*, we define the following notation: F e' = {x 6 F; II 9II-diam(V n F ) > e for all weak* open neighbourhoods V of x }. If ot is any countable ordinal, we d e f i n e F f f +1 - - ( F ~ot) e ! and if 13 is a limit ordinal, we set F [ -- n ~ <~ F~. Set K -- B x , . Since K is weak* metrizable compact, there exists for any e > 0 a countable ordinal or(e) such that K~(e)+l = K~(e). Since X* is nonseparable, there exists some e > 0 such that K~ (~) -- D ~ 0. Hence we have found a weak* compact subset D of X* such that every weak* open n o n e m p t y subset of D has II 9 II-diameter at least e. If C denotes the weak* closed convex hull of D, it follows that every weak* open slice of C has II 9II-diameter at least e. Note that D and C are both symmetric with respect to 0. If we define now B0 = K + C, one checks easily that every weak* open slice of B0 has II 9II-diameter at least e. Moreover, B0 is the unit ball of some equivalent dual norm on X*. Our conclusion now follows from the following fact. FACT 2.7. Assume that f o r some 8 > O, every weak* open slice of B x , has norm diameter at least 8. Then the norm on X is rough. Indeed, pick any x E Sx. For any positive t and v, we pick f and g in B x , such that m i n ( f ( x ) , g ( x ) ) > 1 - tv8 and [If - gl[ > 8(1 - v). Let h ~ Sx be such that ( f - g)(h) > 8 (1 - v). We have IIx 9 thll 9 IIx - thll ~ f (x -4- th) 4- g(x - th)
Renormings of Banach spaces
797
and an easy computation shows IIx -+- thll + IIx - thll ~ 2 + t6(1 - 3v). Since t and v were arbitrary, this proves that the norm is 6-rough. REMARK 2.8. (1) Proposition 2.6 shows in particular that Corollary 2.4. has a strong converse: if X is separable and every continuous convex function on X has at least one point of F-smoothness, then X* is separable. (2) A modification of the proof of (i) implies (ii) in Proposition 2.6 allows to show that a dual space X* has the weak* dentability property if (and only if) every separable subspace Y of X has a separable dual. Indeed, if X* fails to be weak* dentable, there is a weak* compact convex subset C of X* with no small weak* open slices. Proceeding as above, we construct an equivalent rough norm 11 . l] on X. Now, a separable exhaustion argument provides a separable subspace Y of X such that the restriction of ]J . 11to Y is rough, and then Proposition 2.6 shows that Y* is not separable. Theorem 2.2 and Proposition 2.6 are tightly connected with "harmonic" behaviour of smooth functions on nonsmooth spaces. More precisely, the following holds. PROPOSITION 2.9. Let X be a separable Banach space such that X* is not separable, and let Y be any Banach space. Let U be a bounded open subset of X, and let OU be its boundary. Let f be a continuous function from U to Y whose restriction to U is Fdifferentiable. Then" (i) the set f (U) is contained in the weak closure of f (OU) in Y. In particular one has suPu IIf II -- suP0u IIf Jl. Hence f satisfies the maximum principle. (ii) If the derivative f1(x) is a compact operator for every x ~ U, then f (U) is contained in the norm closure of f (OU) in Y. (iii) If Y* is separable, then f (U) is contained in the norm closure of f (OU) in Y. PROOF. If the weak closure of f ( O U ) does not contain f ( U ) , there exists a linear map T from Y to a finite dimensional space R n such that the closure of (T o f ) ( O U ) in R n does not contain (T o f ) ( U ) . Composing with an appropriate smooth bump function b from R n to R, we construct a smooth bump B = b o T o f on X. This contradiction with Theorem 2.2 shows (i), and (iii) can be shown in a completely similar manner, using this time a smooth bump function b from Y to R provided by Theorem 2.2. We outline the proof of (ii), which requires the use of rough norms provided by Proposition 2.6. A first step consists of showing that the restriction of a rough norm II 9II to any finite codimensional subspace is still rough. Then compactness of the derivative provides at every x E U a finite codimensional subspace Nx such that the norm of the restriction of f ' ( x ) to Nx is very small. Since II 9II is rough on Nx, we can find a direction h ~ Nx along which f does not vary much while II 9 II is significantly increasing. In this way, starting from an arbitrary x0 6 U, we construct a path along which f does not change much and which will eventually cross OU since U is bounded, and (ii) follows. D
798
G. Godefroy
REMARK 2.10. Let X be a separable Banach space. If X* is separable, then Theorem 2.2 and the use of partitions of unity (or of Theorem 6.2) show that if F is any closed subset of X and f is any real valued continuous function on F, then there is a continuous extension of f to X such that f is C 1 on X \ F. On the other hand, if X* is not separable, pick g X*\{0}. Set f = Igl, and F = Sx. If I were a continuous function on Bx extending f and having a C 1-smooth restriction to the open unit ball, then applying Proposition 2.9(i) to the level sets of g would show that 1 = f on Bx. But f is not smooth at 0. Hence continuous functions on a closed subset of a separable space can be extended in a differentiable manner exactly when the dual is separable.
Notes and comments. Theorem 2.2 combines the work of [ 109] and [ 116] (for (i) implies (iii)) and of [51] (for (ii) implies (i)). It should be noted that the implication still holds if the bump function is simply assumed to be differentiable [46, p. 59]. If the bump function is Lipschitz, this can be seen by applying the smooth variational principle (see [44] in this Handbook) to a rough norm constructed in Proposition 2.6. Ekeland's variational principle (see [44]) has been applied to many different problems, for which we refer to [ 10]. Subdifferentiability of 1.s.c. functions on Asplund spaces is demonstrated in [56], using [ 19]. We refer to [98] for a recent survey on this topic. Generic F-smoothness of convex functions on spaces with separable dual is due to Lindenstrauss and Asplund [ 127,8]. Corollary 2.4 can be made more precise, using the notion of porous sets [ 151 ]. Note that the proof of this corollary shows that the set of points where a convex function is F-smooth is topologically simple, namely G~ regardless of the space. On the other hand, the set of G-smooth points can be very complicated in nonseparable spaces [97]. It is shown in [140] through topological games that if a Banach space has an equivalent G-smooth norm, then every convex function is G-smooth on the complement of a meager set. It is not known whether this latter property is equivalent in full generality to saying that every continuous convex function is G-smooth on a dense set. We refer to [58] for a comprehensive survey of G~teaux smoothness of convex functions on Banach spaces. Understanding the set of points where a given norm is F-smooth is an active field of research. The topological situation is quite clear, but the measure theoretic analogue, which would say that "norms on Asplund spaces are almost everywhere differentiable" is much harder to handle, since on one hand it is nontrivial to define properly what "almost everywhere" means, and on the other hand natural conjectures turn out to be false. We refer in particular to [133] and [134] where equivalent norms on the Hilbert space, and on infinite dimensional super-reflexive spaces, whose set of Fr6chet smoothness is Aronszajn null are constructed. The behaviour of the space c0(N) in this respect is unclear. Pioneering works on rough norms include [37] and [119]. Proposition 2.6 is a result of [125], which characterizes non-Asplund spaces by the existence of an equivalent rough norm, regardless of the density character, with essentially the same proof. We refer to [ 158] for quantitative estimates on roughness and related results. Proposition 2.6 leads the way to Cantor-Bendixon derivations and their applications to Banach space theory, which are displayed in Section 3 below. Remark 2.8.2) is in [125]. We refer to [162,163] and [164] for a comprehensive investigation of the duality between Asplund spaces and dual spaces with the Radon-Nikodym property. Note that the class of separable spaces which have an equivalent G~teaux smooth but rough norm lies strictly between the class of spaces
Renormings of Banach spaces
799
containing 11 and the class of spaces with nonseparable dual [45]. We refer to [73] and references therein for related weak forms of the Radon-Nikodym property. It should be noted that rough norms exist on every non-Asplund space, while smooth norms fail to exist in general on Asplund spaces; examples are provided by R. Haydon's fundamental work [95]. Let us note that [95] contains also positive renorming results, and one of the original features of Haydon's renormings is that norms can be defined by recursion. This should be compared with the crucial use of recursion techniques in the definition of important examples of Banach spaces [ 171 ]. It is shown in [ 101 ], under the continuum hypothesis, that there are Asplund spaces with no equivalent norm having Mazur's intersection property. No such example has been constructed so far which would dispense with using special axioms of set theory. It is not known whether (not necessarily equivalent) C 1 smooth norms exist on every Asplund space. The behaviour of smooth functions on Banach spaces and Banach manifolds is investigated in the seminal paper [18], where the maximum principle is shown for smooth functions on Banach spaces containing I i. We refer to [46] and references therein for a recent account of the theory. The general principle which lies behind Proposition 2.9 is that functions satisfy the maximum principle when their order of smoothness exceeds the order of smoothness of the space on which they are defined. Approximation of continuous functions by smooth ones is investigated in Section 6 below, where some applications of renormings are displayed. The classical approximation technique, for which we refer to [46, Chapter VIII], consists of showing the existence of smooth partitions of unity. This is difficult in the nonseparable case, and for instance it is still open in full generality whether a Banach space X with an F-smooth norm has C 1-smooth partitions of unity. This is true when the space X is weakly compactly generated (cf., e.g., [46, Chapter VIII]), hence in particular every reflexive space has C 1-smooth partitions of unity [ 169]. Let us note that when a smooth approximation of a Banach space valued function is obtained from locally finite coverings and smooth real-valued partitions of unity, the approximating function has derivatives which are of finite rank at every point. The "harmonic" behaviour of smooth functions on nonsmooth spaces leads naturally to the problem of finding a "Poisson formula" for representing the values inside U by the values on 0 U. Very little seems to be known about this question which goes back to [ 18]. It is shown in [11], among other things, that if X is an infinite dimensional Banach space which has a C 1 smooth norm (nonnecessarily an equivalent one!), then X is C 1 diffeomorphic to X\{0}. One of the motivations for such results is that it easily implies the failure of Rolle's theorem in infinite dimensional spaces: there exists, e.g., on 12 a C~ bump function b such that b' (x) 7~ 0 for every x such that b(x) ~ O. The proof in [ 11 ] relies in part on a renorming technique, namely on Bessaga's "incomplete norm" approach [ 16]. It is natural to "integrate linear theory", that is, to investigate which theorems remain true when linear operators between Banach spaces are replaced by smooth functions. For instance, it is conjectured that any C 1 function with uniformly continuous derivative from c0(N) to a Banach space which does not contain c0(N) maps weakly Cauchy sequences into norm convergent ones (see [91 ]). This would mean that co enjoys a nonlinear form of Pelczyfiski's property (V). This field of research is nearly unexplored.
800
G. Godefroy
3. Renormings of super-reflexive spaces We will make a crucial use in this section of a modification of the "Cantor derivation" which has been used in the proof of Proposition 2.6. Let C be a weak* compact convex subset of a dual space X*, and pick any e > 0. We denote
De(C) = {x ~ C; II 9II-diam(S) > e for every weak* open slice of C with x ~ C }. Note that De(C) is convex, and that we can define by induction D~' for any positive integer n. With this notation, we have LEMMA 3.1. Let X be a Banach space.The space X is super-reflexive if and only if for any e > O, there is an integer n(e) such that D~ (e) (Bx, ) -- 0. PROOF. Let (YTk) be a finite increasing sequence of Boolean algebras generated by partitions of [0, 1] into subintervals, with S0 = {0, [0, 1]}. An X-valued martingale of height n is a sequence (Mk)0~
Renormings of Banach spaces
801
Let II 9II be the original n o r m on X*; distances d ( . , C) to convex sets are taken with respect to this original norm. We define a convex function f on X* by the following formula: oc
f(x)-
Nk
2_ k
[Ix[] + Zk=, n~l= - ~ k d ( x , K ; ) .
We define the new n o r m I 9 [ as the Minkowski functional of the set C -- {x 6 X*; f (x) ~< 1 }. It is clear that I 9 I is an equivalent n o r m on X*. We show that this n o r m is uniformly convex. This follows from the claim below via an easy computation which is left to the reader. CLAIM 3.3. Let x and y in X* be such that f (x) = f (y) = 1 and [Ix - YI[ >/e. Then
f
( ) x+y 2
<.l32n2(e/8)
Indeed, pick k such that e / 8 ~< 2 -~ < e / 4 . We denote by n=Max{j~>O;
xEK~andyEK~}.
Note that n < Ark since [Ix - yl[ ~> e. We let ), = e/4Nk. SUB-CLAIM. For some integer I with 1 <~1 <. Nk - n, one has
_ (d(x
n
2
'
+ d(y,
n+l )) - d I/ x + Y ~,
K~+I I" > y
(*)
~ ]' 2
Indeed if not, we show by induction on I that for all 1 ~< l ~< Ark -- n, one has
l (d(x K'~ +') + cl(y, K'~+')) < Iv. 2 '
(**)
This is true for I -- 1 since x and y belong to K~ and [[x - y [[ >/e, hence x+y - T - 6 Kn+l k 9 If (**) is true for l, there are x' and y' in K ; +1 such that (l[x - x'[[ + [[y - y'[[)/2 < Iv, hence [Ix' - y' 1[ >~ e / 2 . It follows that x I + y, E K~ +/+1 .
x+y
Since we have ][ 2
d
x~+y' 2
][ < l v , it follows that
x + y Kn+l+l) 2 ' k
G. Godefroy
802
However, if ( , ) fails for l, this implies that (**) holds for (1 + 1), which concludes the induction. We can now apply (**) with I - Ark - n to get 1
Nk
e
-~(d(x, K ~ k) + d(y, K~ )) < (Nk - n)?, <~N~:?' = -~.
Hence there exist x I and yl in K ffk such that IIx'-
YZl[ > e/2. This implies that
(x' + y~)/2 E K ffk+l, which is certainly impossible since this latter set is empty. Hence the sub-claim is shown. In order to finish the proof of the claim, it suffices to notice that the functions II 9 II and
d ( . , K kn ) are convex, hence the sub-claim shows that f
x + y ) <~1 2
2 -~
e
Nk "4Nk
The claim follows since Nk <~ n(e/8) and 2 -~ ~> e/8. This concludes the proof of the theorem, ff] Let X be a super-reflexive space, and let 6(.) be the modulus of convexity of an equivalent UC-norm II 9II on X*. Pick e > 0, and let y ~ Bx, be such that IlYll > 1 - ~(e). Let x E Sx be such that y(x) > 1 - 8(e). Set S = { f 6 Bx,; f ( x ) > 1 - 6(e)}. If (f, g) 6 S 2 then I[(f + g)/2l[ > 1 - 6(e) and thus Ilf - gll < e. It follows that y q~ De(Bx,), and thus De(Bx,) C (1 - 6(e))Bx,. By homogeneity, it follows that for all n one has Dn(Bx*) C (1 - 6(e))nBx ,. If no is such that (1 - 6(e)) no < e and n(e) = no + 1, then clearly D n(e~(Bx,) -- 0. This shows the that the modulus of convexity provides a quantitative control on the sequence n(e). Conversely, the proof of T h e o r e m 3.3. shows that the sequence n(e) provides quantitative estimates on the modulus of convexity. More precisely, Claim 3.3. shows that if there is p i> 1 such that n(e) = O(e-P), then 61. I is of power type, that is, there is (C, q) ~ R 2 such that 61. i(e) ~> Ceq. Such a control on n(e) is indeed available. For proving this, we come back to the proof of Proposition 2.6, and to the notation defined there. For any e > 0, we define the Szlenk index of X as
Sz(X, e) = min{ot; (Bx,)~ _ 0}. It follows from the above that if X is super-reflexive, then Sz(X, e) 6 N for any e > 0. We have now LEMMA 3.4. Let X be a Banach space such that Sz(X, e) ~ N for any e > O. Then there exist constants C and p such that Sz(X, e) <~Ce p for all e > O. PROOF. It follows from the definition that for any 8 > 0 and r / > 0, one has
Sz(X, ~o) <~Sz(X, ~)Sz(X, o).
Renormings of Banach spaces
803
In particular, we have for all k E N that Sz(X, 2 -k) ~< Sz(X, 1/2) ~. Pick now any e 6 (0, 1), and choose k such that 2 -(k+l~ < e <~ 2 -k. By the above one has
Sz(X, e) <~Sz(X, 1/2) k+l and thus Sz(X, e) <. Sz(X, 1/2) l+l~
and the lemma follows.
We denote by L 2 (X) the space of Bochner-measurable square integrable functions from [0, 1], equipped with the Lebesgue measure, to X. The next lemma relies on the fact that the Szlenk derivation applied on the space L 2 (X*) controls the slice derivation on the space
X*. LEMMA 3.5. Let X be a super-reflexive Banach space. Then there are constants C and q
such that n(e) <. Ca -q for all e > O. PROOF. First we note that the space L2(X *) is super-reflexive. Indeed, by Theorem 3.2. there is an equivalent uniformly convex norm on X* and one checks that the norm it induces on L2(X *) is uniformly convex as well. It follows in particular that Sz(L2(X), e) E N for all e > 0. We claim now that
n(e) <. S z ( L 2 ( X ) , 2 ) .
(1)
Indeed fix e > 0 and pick n < n(e). By the proof of L e m m a 3.1. there exists an (e/2)separated martingale (Mk) of height n with values in Bx,. Let us denote L0 -- BL2(X, ) and by induction Lk -- (Lk-1) t~/2 with the notation of Proposition 2.6. The sets Lk are therefore relevant to the "Cantor derivation" in the space LZ(x*). One has Mj ~ L n - l - j for all 0 ~< j <~ n - 1. This is obvious if j ----n - 1 and it can be checked for all j through backwards induction and the following observation: let F be a finite subset of the unit ball of a Banach space E and x E conv(F) be such that Ily - xll i> c~ for every y 6 F. The constant function fo = l[o, llX E LZ(E) is a weak limit in L2(E) of a sequence of functions ( f , ) which take their values into F. Indeed, since y E conv(F), there is a partition of [0, 1] into finitely many intervals I1,12 . . . . . Ik and an F-valued function f which is constant on each interval Ij and such that
fo f (t) dt - x. m
m
If we denote f the 1-periodic function on R which extends f and if we let fn (t) - f (nt), it is easily checked that (fn) works. Moreover Illl0,1]x - fn IIL2(E) ~ ot since fn is F-valued. Therefore any weakly open subset of BL2 (E) that contains f0 has diameter at least or.
G. Godefroy
804
In particular, one has Mo e L n - 1 and thus Ln-1 ~ 0, and (1) follows. Finally, (1) and Lemma 3.4. imply immediately the conclusion. [3 From Lemma 3.5. and the computations we made for proving Theorem 3.2 we deduce THEOREM 3.6. Let X be a super-reflexive Banach space. There exist constants C a n d q a n d a uniformly convex norm I 9I on X such that 6(x,I. I)(e) >~ CE q. There exist constants C I a n d p a n d a uniformly smooth norm 11 . 11on X such that P(x, II. II)(7) ~< C~rlp. It turns out that one necessarily has p ~< 2 ~< q. When X is infinite dimensional, this follows from Dvoretzky's theorem. REMARKS 3.7. (1) Equivalent uniformly convex norms on super-reflexive spaces with a modulus of power type appear therefore as canonical objects. If one performs the natural "slice derivation" on bounded sets of the space, starting from an arbitrary unit ball, one obtains for each e > 0 a finite sequence of convex sets. The Minkowski functional of a weighted series of distances to these convex sets is a uniformly convex norm with a modulus of power type. Roughly speaking, the removal of small slices "rounds up" unit balls. (2) The two norms obtained in Theorem 3.6 can be combined to construct an equivalent norm on any super-reflexive space X which is simultaneously UF and UC with moduli of power type. Let us give a formula which works: if II 9 Ill is UC and II 9 112 is UF, we define for all f e X* and all n ~> 3 an equivalent norm II 9 IIn on x * by the formula * 2) 1/2 [[flln9 -- (([[flll*)2 + n - 1 (11f[[2) Then the norm defined on X by oo
Ix 12 -- ~
n -3 IIx II2n
n=3
is UC and UF. If moreover 611. Ill and Pll 9112are of power type, then the moduli of convexity and smoothness of I 9I are both of power type. Lemma 3.1 is mainly a reformulation of the fact that superreflexivity is equivalent to knowing that the height of e-subtrees in the unit ball is bounded for any e > 0 ([ 100]; see [12]). Theorem 3.2 is Enflo's renorming characterization of superreflexivity [53]. The present proof is due to G. Lancien [123]. Lemma 3.4 is a usual submultiplicativity argument, which is in this case a simple geometric observation, and which leads to power-type behaviour. We refer, e.g., to [66] for its use in probability theory. Lemma 3.5 provides a way to control the dentability index of X by the Szlenk index of L Z(x). It is interesting to notice that even this more geometrical approach requires the use of the space LZ(x), which gives a quantitative control on trees that geometry fails to provide. Theorem 3.6 is Pisier's quantitative improvement of Enflo's result [ 147]. The exponent in the modulus of convexity is greater than the cotype and they are in general different [ 146]; dually, the exponent in the modulus of smoothness is (in general strictly) less than the type. Moreover, the best exponents are not attained in general, and terms of
Notes a n d comments.
Renormings of Banach spaces
805
smaller order show up. We refer to the negative solution to the three-space problem for 12 [54] and to the examples of spaces which are "close to be Hilbertian" without being Hilbertian (see [103,149]) for such phenomenons. Remark 3.7(2) is a modification of Asplund's averaging technique [9], which is in [102]. It is not known whether both exponents can be kept in general. The Szlenk index has been defined in [165], in a different but equivalent way, and used there for showing that there is no universal space for the class of separable reflexive spaces. When applied to super-reflexive spaces, this index is a natural number, but for general Asplund spaces it is an ordinal (which is countable when the space is separable). Szlenk's work led to further research in two related directions which we now describe. The existence of a universal space for a given isomorphism class (C) is tightly related with the topological nature of (C). This point of view originates in [23], and it has been investigated and put into its proper frame by Bossard [20]. It is interesting to notice that most of the natural isomorphism classes are, in a sense which is made precise in [20], non-Borel sets. Moreover, there is no constructive way of picking a separable Banach space in each isomorphism class; in other words, there is no analytic section of the equivalence relation "being linearly isomorphic" [21 ]. This topological approach has been used by G. Lancien [123], who applied the Kunen-Martin theorem for showing the existence of a universal function ~ from the set of countable ordinals to itself such that if the Szlenk index of a Banach space X is ot < col, then its dentability index (which is defined through the derivations De) is less than ~p(ot). Since these indices are separably determined, it follows that if the Asplund space X is such that the Szlenk indices of the separable subspaces of X are uniformly bounded by some ot < col, then X has an equivalent norm whose dual norm is 1.u.c., hence in particular an equivalent F-smooth norm. In other words, spaces whose separable subspaces are "uniformly Asplund" have equivalent F-smooth norms [ 123]. The construction of the dual 1.u.c. norm relies on a formula similar to the series from the proof of Theorem 3.2. A modification of the derivation approach has been subsequently used by Raja ([153,152])who provided a transparent proof, and significative improvements, of Troyanski's theorem [170] asserting that any Banach space which has a strictly convex norm with the Kadets-Klee property has an equivalent 1.u.c. norm (cf., e.g., [ 176]). It is easy to see that invertible isometries of the original norm are still isometries for the new norm when the Enflo-Pisier renorming is performed. In other words, one can improve the convexity and smoothness of the norm, while increasing, or at least keeping, the "unitary group" of invertible isometries. It is an open question whether a separable Banach space such that the unitary group acts transitively on the sphere is isometric to the Hilbert space. Note that when the norm is quasi-transitive (that is, when the orbits of points in the sphere are dense in the sphere), this norm necessarily has an optimal modulus of convexity and smoothness [68], as can be seen, e.g., with the smooth variational principle (see [46, Section 4.5]). It is not known whether every superreflexive space has an equivalent quasi-transitive norm (see [25]). Super-reflexivity, and the moduli of convexity and smoothness, are tightly related with the "girth" of a Banach space, that is, with the infimum of the length of curves contained in the sphere and joining two antipodal points of the sphere. This girth is greater than 2 if and only if the space is super-reflexive (see [160] and references therein). The moduli of a
G. Godefroy
806
given norm on X also have a strong impact on the behaviour of sequences and series in X. We refer to [49] and references therein for a comprehensive account of this topic.
4. Smoothness of higher order As shown in Section 2, separable spaces on which there a r e C 1 smooth bump functions are characterized by the separability of their dual. We show in this section that the existence of C 2 or C ~ smooth bumps provides much stronger information on the space. We first define polynomials in an infinite dimensional setting. DEFINITION 4.1. Let X and Y be Banach spaces. A Y-valued monomial of degree n is a function f " X --+ Y such that there exists a Y-valued continuous n-linear form F such that f (x) = F (x, x . . . . . x) for all x 6 X. A polynomial on X is a finite linear combination of monomials. Recall that a function g from a Banach space X to a Banach space Y is differentiable at x if there exists T : X --+ Y continuous and linear such that g (x + h) -- g (x) + T (h) + o(h), with limh~0 o(h)/llhll = 0. In this case, T = gt(x) is called the differential of g at x. We say that g is C 1 if it is everywhere differentiable and if g ' : X --+ L(X, Y) is continuous. Since L(X, Y) is a Banach space, we can of course iterate and define C k functions for all k E N. Note that Schwarz's lemma permits to identify d (n) f (x) with a continuous n-linear symmetric form from X n to Y. We say that a function is C ~ if it is C ~ for all k 6 N. Finally, if f is C ~ , we can define for all x 6 X and all n 6 N the Taylor polynomial Tn f (x) by the formula /7
Tnf (x)(h) = Z ( 1 / n ! ) d ( n ) f (x)(h, h, ..., h). k--O We say that f is analytic if for every x 6 X, there is 6(x) > 0 such that f = l i m n ~ Tnf(x) uniformly on B(x, 3(x)). We say that a norm is C k, C ~ , analytic . . . if it is so outside {0}. EXAMPLES 4.2. (1) If p = 2n is an even integer, the natural norm on L p is analytic, since we can write Ilfllp = ( p ( f ) ) l / p , where P is a monomial of degree p. Note that absolute values are not needed by positivity of even powers. (2) Let us define on c0(N) the function ~((xi)) - - ~_~i~=1X 2 i . Then the set C = {x co; ~b(x) ~< 1} is the unit ball of an equivalent norm on co. It follows from the analytic case of the implicit function theorem that this norm is analytic on co. The results from this section will show that the above classical spaces are the canonical examples of spaces which have C ~ smooth equivalent norms. In order to sketch the proofs, we need a compact variational principle, valid for spaces which do not contain an isomorphic copy of c0(N). We also state and prove a quantitative version of this principle for spaces with a nontrivial cotype.
Renormings of Banach spaces
807
LEMMA 4.3. Let X be a Banach space, let U be a bounded symmetric open subset of X, and let f : U --+ R be a continuous function, such that f (O) <~0 and inf{f(x); x ~ OU} = m > O. For any compact subset K of U, we let f K ( x ) = sup{f(x + k); k ~ K} for all x E X such thatx + K C U. (i) If X does not contain an isomorphic copy of c0(N), then there exist a symmetric compact set K and a symmetric neighbourhood V of 0 with K + V C U, such that for every 6 > O, there exists a finite subset K~ of K such that i n f { f K ~ ( x ) - fK~(0); x 6 V,
Ilxll ~ ~} >
0.
(ii) If X has a finite cotype q, there exist a symmetric compact set K, a symmetric neighbourhood V of 0 with K + V C U and A > 0 such that fK (x) ~> fK (0) + A IIx IIq
for all x ~ V. PROOF. We first prove (i). We let xo = O. If xo, x 1. . . . .
Xn
have been constructed, we define
Kn--{~eixi;i=o 8 i E { - 1 , 1 } } and
En--
xEX;
for a l l k 6 K ~ ,
k+xEUandf(k+x)~<~-(1-
m
2 - n ) }.
We let an = sup{ IIx II; x E En } and we choose Xn+l E En such that IlXn+lll ~ C~n/2. Then we define K = Un~>0 K~. Since K is contained in U, it is bounded. Since X does not contain c0(N), we have by Bessaga-Petczyfiski's theorem that K is compact and limx~ = 0, therefore lim oe~ = 0. Since f is continuous, we have f ( x ) <~ m / 2 for all x E K, and thus K A OU = 0. Hence there exists a symmetric neighbourhood V of 0 with K+VCU. We pick 6 > 0, and we choose n such that an < 6. Pick x 6 V such that IIx II/> ~. Since otn < 6, we have fKn (x) > m(1 - 2 - n ) / 2 . On the other hand, fK~ (0) -- max { ft:._, (Xn), fK~_, ( - x ~ ) } -- fK._, (Xn) and since Xn E En-1 this implies fKn (0) ~< m(1 - 2n-1)/2. Therefore fKn (x) > fK~ (0) + 2 - ~ - 1 m and this shows (i). The proof of (ii) is similar. We assume for convenience that m -- 1 and U C Bx. Using the same notation, we set
G. Godefroy
808 and
En - {x 9 X; for all k 9 K . , k + x 9 U and fKn (k .Af_X) ~ fKn (0) .Af_(2C) -1 [ixllq }, where C is the cotype q constant of X. We let as before an = sup{ Ilx II; x e En } and we choose Xn+l 9 En such that IIx~+~ II/> c~/2. Since X has a nontrivial cotype, it does not contain c0(N) and thus we have as before that lim an = 0. We have now
supf (ei)
6iXi
<~ (2c)-lllxnll q + s u p f
i=1
(ei)
6iXi
~< (2C) -1
i=1
[]Xil[ q
9
i=1
Since X has cotype q and U C B x this implies
sup
f
@i)
6iXi
1
~ -~.
i=1
It follows that f (x) ~< 1/2 for all x 9 K and thus U n ) 0 Kn - K C U. Let V be a symmetric neighbourhood of 0 such that K + V C U. Pick x 9 V with x # 0. We have IIx II > ~n for n large enough. Since Kn is an increasing sequence of sets, we have
f K ( X ) - - l i m fKn (x) t> lim fKn (0) + ( 2 c ) - l l l x l l q - fK(O) + ( 2 c ) - l l l x l l q 11 fl This shows (ii).
[]
Our first application of L e m m a 4.3 is the following uniformization result. THEOREM 4.4. Let X be a Banach space which does not contain an isomorphic copy of ,(k) co(N), and let k 9 N. If there is a Ck-smooth bump function bo defined on X such that o 0 is locally Lipschitz, then there is a Ck-smooth bump function b defined on X such that b (k) is globally Lipschitz on X. PROOF. We give the proof for k = 1; trivial modifications provide the general case. We may and do assume that b0(0) = 1, that bo(x) = b o ( - x ) for all x and that bo(x) = 0 if IIx II/> 1. We let f = 1 - b0, and U = {x 9 X; f ( x ) # 1 }. We apply L e m m a 4.3 to f , U and m = 1 to get V and K as in (i). Since f~ is locally Lipschitz, there is fl > 0 such that f l B x C V and f t is Lipschitz on K + f l B x . Pick 0 < 3 < fl/2. L e m m a 4.3 provides K~ and e > 0 such that if ~ < IIx II ~ then
f K~ (x) >>f x~ (0) + e. For x 9 fi B x , we define now
~ ( x ) = Z ( f (x + y ) yeKa
f(y))2
Renormings of Banach spaces
809
Then ~ ( 0 ) - 0, the derivative (Tr)' is Lipschitz on f i B x and ~p(x) ~> 6 2 if x E f l B x and [Ix II ~> 8. Composing with an appropriate smooth function from R to R provides the conclusion. D The uniform behaviour of derivatives provides important information on the space, through the construction of uniformly smooth equivalent norms, as sketched in the following result. P R O P O S I T I O N 4.5. Let X be a Banach space. I f there exists a C 1-smooth bump function b defined on X whose derivative b ~ is globally Lipschitz on X, then there is an equivalent norm N ( . ) on X whose derivative is globally Lipschitz. In particular, N is uniformly
smooth and tON(rl) ~ Crl 2. PROOF. We outline the construction of the norm. We may and do assume that b(0) > 0, that b is an even function and that 0 ~< b(x) ~< 1 for all x E X. Let us define c/)(x) --
F
OG
b ( s x ) ds
OG
and gr(x) = 4)(x) -1 for x -r O, with gr(0) --O. Substitution shows that gr(tx) --Itlgr(x) for all x and all t 6 R. It follows that there exist a > 0 and b > 0 such that allxll ~< gr(x) ~< b IIx II for all x 6 X. There is k E R such that for all x with 89~< IIx II ~< 3 one has
4)(x) --
b ( s x ) ds.
k It follows that 4 / i s Lipschitz on the unit sphere. We define now ~(x) = ~ ( x ) 2. Clearly (al[xi]) 2 ~< ~(x) ~< (bllxll) 2. One checks by direct computation that the derivative ~' exists and is globally Lipschitz on X. We now use ~ and a convexification procedure to construct the equivalent norm. Let F be the largest convex function which is less than ~ on X. The epigraph of F is the closed convex hull of the epigraph of ~, and F can be estimated through the formula
where of course the 0ti'S are positive a n d z i n = l o/i - 1. Let C be the Lipschitz constant of ~'. For all x and h, one has F ( x § h) + F ( x - h) - 2 F ( x ) ~< 2CIIhll 2.
(1)
For proving (1), we pick any 6 > 0 and find (Ofi) and (Xi) as above with x -- Z i L 1 olixi and 17
Z i=1
oti~(xi) < F ( x ) + ~.
G. Godefroy
810 Then we have
F ( x + h) + F ( x - h) - 2F(x)
Z
Oti~(Xi @ h) +
i=1
Oti~(Xi -- h) - 2 Z i--1
Oli~(Xi) -+-2e.
i=1
Thus F ( x + h) + F ( x - h) - 2F(x)
n Z Oli(~(Xi + h) + ~(Xi -- h) - 2 ~ ( x i ) ) -k- 2e ~< 2CIIh[I 2 + 2e i=1
by the Mean Value Theorem. Since e is arbitrary, this proves (1). Since F is convex, (1) shows that F is C 1 and that F t is Lipschitz. The norm N is simply the Minkowski functional of K = {x ~ X; F(x) ~< 1}. ff] By the Mean Value Theorem, every C 2 smooth function has a locally Lipschitz derivative. Hence Theorem 4.4 and Proposition 4.5 imply the following corollary. COROLLARY 4.6. Let X be a Banach space. If there exists a C 2 smooth bump function on X, then X contains an isomorphic copy of co(N) or X is superreflexive of type 2. This corollary, when compared with Theorem 2.2, shows that C 2 smoothness is a much more restrictive condition than C 1 smoothness. The next theorem shows that C ~ smoothness provides considerable information on the space. THEOREM 4.7. Let X be a Banach space which does not contain an isomorphic copy of c0(N), and such that there exists a C ~ smooth bump function defined on X. Then: (i) There exist an integer k ~ 1, a 2k-form a(.) on X, c > 0 and d > 0 such that f o r all xEX, cllxll 2k ~ a(x) <~dllxll 2k.
(ii) There exists an integer p ~ 1 such that inf{q; X of cotype q } - 2p, and X is of cotype 2p. (iii) X contains an isomorphic copy of l2p, where p is the integer defined in (ii). PROOF. The proof of this theorem is quite involved and we will only sketch the proof of (i), namely the existence of a so-called separating polynomial. We recall that Tk f (x) denotes the Taylor polynomial of order k of f at x. FACT. Let f be a C k smooth function, and let K C X be a compact set. For every e > O, there exists 6 > 0 such that f o r every x ~ K and every h ~ X with ]lh II <<,6, one has I f (x + h ) -
T k f (x)(h) I ~
ellhll ~.
Renormings of Banach spaces
811
The fact is an easy consequence of Taylor's formula with integral remainder. Together with the compact variational principle, it allows us to show the existence of separating polynomials on spaces whose cotype is controlled from above by the "order of smoothness". LEMMA 4.8. Let q ~ 1 be an integer. Let X be a Banach space of cotype q, on which there exists a C q smooth bump function dp. then there exist polynomials P1, P2 . . . . . Pn of degree at most q on X such that Pi (0) -- 0 for all 1 <~ i <~n, and for all x E X with [Ix ]l -- 1,
m/ xlP, x l To prove the lemma, we may and do assume that 4~(0) = - 1 , that 4~ is an even function and finally that 4~ is uniformly continuous. This last condition can be obtained through Theorem 4.4 since X has a finite cotype and in particular does not contain co (N). We denote f = ~b + 1 and U = {x E X; f ( x ) ~- 1 }. L e m m a 4.3(ii) provides us with a symmetric compact subset K of U, a symmetric neighbourhood V of 0 such that K + V C U, and A > 0 such that for all x E V,
fK (X) >~ fK (0) + A IIx IIq 9 By the above Fact, there is fl > 0 such that fl Bx C V and for every x E K and every h E X with Ilhll ~3, one has
Since f is uniformly continuous on K + V, there is a finite subset K0 -- {x l, of K such that for all h E V, one has
fKo(h)>/fK(h)--(A)fl
X2 .....
Xn}
q.
We denote Pq(xi)(h) - Tq(xi)(h) - f ( x i ) . Pick any h E X with ] l h l l - / 3 . By definition, one has
fKo(h) -- sup{f(xi + h); 1 ~ i ~< n} For any 1 <~ i <~n, f (xi + h) - f (xi) ~ Pq(xi)(h) + (A)flq, and therefore
f K o ( h ) - fKo(O) <~max[Pq(xi)(h)]-+- ( 4 ) f l q . It follows that
G. Godefroy
812
It suffices now to set Pi (x) -- (A@r This finite set of polynomials satisfies the conclusion of the lemma. To prove Theorem 4.7(i), we notice that by Corollary 4.6, the space X is superreflexive and therefore it has a finite cotype. Thus we can find an integer q such has X has cotype q, and certainly it has a cq smooth bump. Lemma 4.8 provides us with separating polynomials P1, P2 . . . . . Pn. We let P - - ~ i n = l p2. Let us write P = a l + a2 -+-...-k- ak, where each aj is a j-form on X. Since IPI ~> 1 on Sx, we have 1 sup [aj(x) I ) l<~j~k k
for all x with
IIx II = 1. Hence if we let k
Q-~a~
(k')/j
j=l
then Q is a 2(k!)-form and Ia(x)l ~ k -2(k!) if Ilxll = 1. Therefore Q works.
Notes and comments.
D
Polynomials on Banach spaces have been intensively studied (see [ 141,3,85]). They are applied among other things to a better understanding of holomorphy in infinite dimensional complex spaces. Lemma 4.3 is from [41] (see [46, Section 5.2]). It can roughly be interpreted by saying that Banach spaces which do not contain co are such that there is "some compactness" in them which allows some uniformization. This uniformization takes a quantitative form when finite cotype is assumed. Quite naturally, in view of the Krein-Milman theorem, this weak form of compactness is related with extremal structure; for instance Lemma 4.3 easily implies that unit balls of spaces not containing co have extremal compact subsets ([173]; see [46, Section 5.2]). Theorem 4.4 is an example of this uniformization; it was originally shown in [63]. Proposition 4.5 relies on a construction of [126] and arguments from [63]. The Lipschitz character of the derivative b t is crucially important: there are C~-smooth functions ~ on 12 such that the convexification F (in the notation of the proof of Proposition 4.5) is not even G~teaux smooth [46, Example 2.5.6]. Theorem 4.7 is Deville's fundamental theorem on C~-smooth Banach spaces [39,40], which shows that Maurey-Pisier's theorem [ 136] has a global analogue for spaces which have C~C-smooth bumps. This theorem states in particular that spaces which have C ~ smooth norms are containing the "obvious" examples 12p or co. Hence the problem of containment of a classical sequence space, which by Tsirelson's example ([ 171]; see [27]) has a negative answer for general Banach spaces, has a strong positive answer for C *c smooth Banach spaces. Let us mention that Deville's theorem has no nonseparable analogue (cf., e.g., [ 176]). Note that Corollary 4.6 already shows that C2-smooth spaces contain co or are reflexive, while by Gowers' example of a space with separable dual which does not contain co or a reflexive infinite dimensional subspace [87], even weaker forms of this conjecture are false for C 1-smooth spaces. Contrasting with the Tsirelson phenomenon, the notion of stable Banach space [118] provides an isometric condition of "interchange of limits"
Renormings of Banach spaces
813
which ensures the existence of subspaces isomorphic to lp, with arbitrarily small BanachMazur distance. So for instance, stable norms on lP are not distorted. Stability of the norm allows to show that any subspace X on L 1 contains a subspace isomorphic to I p where p is the best type of X [89], which improves on Rosenthal's theorems ([ 156,157]). The deep relations between commutation of quantifiers, topological games, nondistortability and the containment of "nice subspaces", has recently been a very active direction of research, for which we refer in particular to [143] and [88]. Theorem 4.7 as stated does not exhaust the contents of Deville's theorem. For instance, if X has a C ~ - s m o o t h bump function, has cotype 2n and no infinite dimensional subspace has cotype < 2n, then X contains a copy of 12n , and this leads through a simple minimality argument to 12n subspaces of X where 2n can be less than the best cotype of X. For instance, if X = 12 G 14, this result provides 12, while Theorem 4.7(iii) gives 14. Also, if p >~ 2 is an integer and X has a CP-smooth bump function and is saturated with subspaces of cotype p, then X has cotype p. It follows for instance that the space X -- (~--~n~__l/~)2, which has a C 1 norm with Lipschitzian derivative, and thus a norm with modulus of smoothness of power type 2, has no C2-smooth norm (see [46, Sections 5.4 and 5.5]). It does not even have any twice Fr6chet differentiable norm [47]. It is not known whether every reflexive space which has a C2-smooth norm is a UMD space. A related problem is to know whether a reflexive space which has a C ~ - s m o o t h bump function, and is therefore of exact cotype 2p, has a uniformly convex norm of power type 2p. This question might be related with finding the lowest possible degree for the separating form in Theorem 4.7. Even in the separable case, it is not clear whether for k > 1 bumps and norms define different smoothness notions: that is, it is not known whether a separable space which has a C2-smooth bump has a C2-smooth equivalent norm. Haydon [95] showed among many other results that it is not so for general Banach spaces, and that there are Banach spaces with a C ~ - s m o o t h bump and without a G-smooth equivalent norm. Smoothness of higher order is related with norms which "locally depend upon finitely many coordinates". This property means that any point x ~ 0 has a neighbourhood V such that the norm factors on V through a finite rank linear map. It follows from Lemma 4.3 that the existence of such a norm implies that the space is saturated with subspaces isomorphic to co. These norms naturally lead to the consideration of spaces with countable boundary. The existence of C ~ - s m o o t h norms on C(K)-spaces, where K is a countable compact, is shown in [94]. This result is improved in [42,43] where "countable generalized boundaries" are used to show uniform approximation of continuous convex functions by C ~ smooth convex functions on several spaces, including the Hilbert space. However, there are equivalent norms on the Hilbert space that cannot be approximated by norms with uniformly continuous second derivatives [ 141 ]. It should be stressed that the usual "inf-convolution" formulas provide C 1-smooth approximants with Lipschitz derivatives, but fail to provide 2 C -smoothness, even in finite-dimensional spaces. Analytic approximation has first been considered in [ 120], where it is shown to hold for continuous functions if there exists a separating polynomial (hence by Theorem 4.7 if X is reflexive and has a C ~ - s m o o t h bump). Using Gaussian substitutes for bumps, it has been shown to hold in co for uniformly continuous functions [31,71 ]. It is not known whether continuous functions on co are uniform limits of analytic functions. Let us mention that even the smoothness behaviour of convex functions on co is not fully elucidated. For instance, it is not known whether every contin-
814
G. Godefroy
uous convex function f on co has a point x of "Lipschitz-smoothness", that is, such that f ( x + h) = f ( x ) + f ' ( x ) ( h ) + O(llhll 2) (see [55] where this notion is introduced). The best order of smoothness for renormings of classical spaces has been thoroughly investigated. The natural norm of L p spaces is optimally smooth (see [46, Section 5.1]), which follows in particular from its quasi-transitivity. More generally, best smoothness of Orlicz spaces has been carefully studied by Maleev and Troyanski, e.g., in [ 132].
5. Characterizing spaces by renormings Results from Section 2 show that if X is separable, the dual X* is separable if and only if X has an equivalent F-smooth norm, while X* is not separable if and only if X has an equivalent rough norm. Hence separability of the dual space is naturally characterized in terms of equivalent norms. The same is true for super-reflexivity as shown in Section 3. In this chapter we will investigate more classes of Banach spaces which can be characterized in terms of existence of certain equivalent norms on the space or on its dual. The Banach space which admits the largest number of characterizations is certainly the Hilbert space. Not surprisingly, it can also be characterized in terms of existence of certain norms on X and X*. Our first statement is a very simple fact. PROPOSITION 5.1. Let X be a Banach space equipped with the norm II 9 II. If the convex function defined by f (x) = Ilxll 2 is twice differentiable at O, then X is isomorphic to a Hilbert space. PROOF. We clearly have f~(0) = 0. Hence the second order Taylor polynomial T2 of f at 0 is a symmetric bilinear form. Since f ( x ) - T2(x) = o(llxl12), there is e > 0 such that T2(x) /> Iix112/2 whenever Ilxll -- e. It follows that I x l - ~/T2(x) is an equivalent Hilbertian norm on X. [2 Note that the above function f is clearly differentiable at 0 for any Banach space X. Hence this simple example demonstrates that smoothness of order 2 is considerably stronger than smoothness of order 1. This can also be said of the related following statement, since any reflexive space has an equivalent C 1 norm (see Theorem 2.2.). THEOREM 5.2. Let X be a Banach space such that X and X* admit equivalent C2-smooth norms. Then X is isomorphic to a Hilbert space. PROOF. If X* contains an isomorphic copy of co, then X* contains a copy of l ~ and thus X* is not an Asplund space, which contradicts the existence of an equivalent Frechetsmooth norm. By Theorem 4.3 above, it follows that X* is a superreflexive space of type 2. Hence X is superreflexive as well, and by Theorem 4.3. again it is superreflexive of type 2. But since X* has type 2, X has cotype 2, and then Kwapiefi's theorem provides the conclusion. U] REMARK 5.3. Let us outline a simpler proof of this result in the case where the norm II 9 II of s and its dual norm II 9 I1" are both C 2. We define the convex function g by
Renormings of Banach spaces
815
g(x) = IIx 112/2. Recall that the (Fenchel) dual function g* to g is defined on X* by Fenchel duality through the formula g*(x*) = sup{x*(x) - g(x); x E X}. In this case one checks easily that g*(x*) = Iix*112/2. We pick any xo r 0 in X. The convexity of the functions g and g* shows that if f (xo) - x~ then (g*)'(x~)) - xo. Since g* is C 2, there is 6 > 0 such that for some c 6 R and all h* 6 X* with IIh* II < 6, one has
g*(x~) + h*) - g*(x~)) - h*(xo) <~cllh* II2. An easy computation using Fenchel duality now shows that
g(xo + h) - g(xo) - x~)(h) >~ Iih112/4c for all h c X such that IIh II < 2c3. But then Taylor's expansion of order 2 shows that
g" (xo)(h, h) >~ Iih112/2c for all h 6 X. Hence [g"(xo)(h, h)] 1/2 = N ( h ) is an equivalent Hilbertian n o r m on X. THEOREM 5.4. Let X be a separable Banach space. The following assertions are equiv-
alent: (i) X is reflexive. (ii) There is an equivalent norm II 9 IIM on X such that every bounded sequence (Xn) such that limm limn IlXm + xn IIM = 2 limn IlXn IIM is norm convergent. PROOF. If II 9II denotes some equivalent n o r m on a Banach space X and x ~ X, we define the symmetrized type n o r m II 9 I1~ by
Ily IIx = IIx Ily II + Y II + IIx Ily II - y II. For every x e X, II 9 IIx is an equivalent n o r m on X such that 211yll ~ Ilyll~ ~ (2 + 211xll)llyll for all y. To check the triangle inequality, one uses the fact that for fixed u and v in X, the function s(r) = Ilru + vii + Ilru - vii is convex and even on R and thus is increasing on R +. We fix now a countable dense Q-vector subspace of X, which we denote C, and we choose a sequence (pc)ccC of positive real numbers such that Y~c~c pc(1 + Ilcll) < ~ . We define a map A from the set of equivalent norms on X into itself as follows:
A(II. II)(x) -- ~
P~llxll~.
ccC
We first prove that (i) implies (ii). Let II 9 II be a strictly convex equivalent norm. It turns out that A(A(II . II)) = II 9 IIM satisfies the conclusion of the theorem. This relies on the following crucial fact.
816
G. Godefroy
FACT. Let us denote II 9Ill = A(II. II). Let (Xn) C X be a sequence such that Ilxn II - 1 f o r all n and lim lim Ilxm -+-Xn Ill -- 2 lim Ilxn Ill, m
n
n
then there is a subsequence (x 88 o f (xn) such that f o r all y E X and ~, fl ~ 0 one has
lim lim Ily -+- oeXZm-4-/3XZnII - lim Ily + (~ -4- ~)X'm lie n m To show this fact, we extract a subsequence (x'/7) such that for every c 6 C, y E C and or, fl E Q + , the limits lim lim IIY -4- c~x~ -4-/3x'~ IIc m
n
exist, which is easy through a diagonal argument. The assumptions of the fact and a classical convexity argument imply that for all c E C,
lim IIx'm + X'n IIc -- 2 lim IIx'nlice lim n n If we let c -- 0, we obtain since
(Xn) is normalized that
/
lim lim IIx'm -4- x n II - 2 m
n
and thus for c~, fl ~> 0 one has limlim IIotXtm + m
/7
~x'nll-~ + ~.
(1)
Similarly we have for all c E C that
lim lim IIc~x~m+ ~X'n IIc -- (c~ + ~) lim IIx~mIIc, m /7 m
(2)
Let y E C and or, t3 E Q+. We apply (2) to c = (or + f l ) - l y . Using (1), we obtain that limlim(lly + m
/7
otx'm +
fix'n II -+- Ily - ~X'm - - ~X'n II)
--limm(lly + (~ + 13)X'm II + IlY -- (~ +/~)X'm II). By the triangle inequality, we have for r/6 { - 1 , 1 } that limlim Ily -4- r/(oex'm + ~X'n) II m
n
immlyl ~
<~ 1
+ Ootx m + n m y
~< lim Ily + ~(o~ + r m
m II
o~+/~
+ ~'.
(3)
Renormings o f Banach spaces
817
and now it follows from (3) that we have l i m l i m IlY + ~X'm + ~X'n II - lim Ily + (oe + ~)X'm II. m
n
m
We proved the fact for y E C and or,/3 E Q + . An obvious density argument concludes the proof. In the notation of the fact and under the same assumptions, one obtains through an inductive procedure the following statement: given s > 0, there is a subsequence (xff) of (Xn) such that for all k E N and all positive real numbers oil, 0/2 . . . . . Otk, one has k
OliX i
(4)
/> (1 - - e ) ~ O t i . i=1
i=1
In particular, it follows from (4) and Mazur's theorem that (x~I) has no weakly null subsequence. We now conclude the proof of (i) implies (ii). Starting from a strictly convex norm ]l 9II, we let II 9Ill - A(II. II) and we denote II 9IIM -- A(II 9Ill). We consider a sequence (Xn) such that lim lim Ilxm -+- x/7 IIM - 2 lim m
/7
/7
Ilxn IIM.
Since the n o r m II 9 IIM is strictly convex, it suffices to show that any such sequence has a norm convergent subsequence. Since X is reflexive, passing to a subsequence we m a y assume that Xn = x -Jr-Yn where yn is weakly null and lim Ily~ II = A exists. If A = 0 we are done. A s s u m e it is not so. Then we m a y also assume that IlYn Ill = 1 for all n. We now apply the Fact to II 9 II1 and II 9 IIM = A(II 9 Ill) to find a further subsequence which we still denote (x/7) such that for all y E X, one has Y lim lim II Y + Xm + X/7 I1~ - 2 lim ~ + X m m
/7
m
Letting y -- - 2 x , it follows with the above notation that (5)
lim lim IIYm + Y/7 I1~ - 2 lim IIYm I1~ - 2. m
n
m
Choose y* E X such that IlY* II - 1 and lim
IIAy*
- yn II - o. It follows from (5) that
lim lim II Y* + Y* Ill - 2 lim II Y* Ill. m
n
m
We m a y now apply again the fact, this time with II 9 II and II 9 Ill = A(II. II), and its consequence to conclude that (y*) has no weakly null subsequence. But this is a contradiction since (Yn) is weakly null. We now prove that (ii) implies (i). Pick f E X* with IlfllM -- 1. Let (x/7) be a sequence such that IIx~ll ~ 1 for all n and lim f ( x n ) = 1. It is clear that the sequence (Xn) satisfies
818
G. Godefroy
the assumption of (ii) and thus it is norm convergent. Its limit x satisfies IIx IIM = f ( x ) = 1. Hence every f E X* attains its norm and thus by James' theorem X is reflexive. Let us mention that a proper use of Goldstine's theorem provides an alternative proof which does not rely on James' theorem. [3 PROPOSITION 5.5. Let X be a Banach space. The following assertions are equivalent: (i) X contains a subspace isomorphic to l 1(N). (ii) There is an equivalentnorm II 9l] on X and t ~ X**\{0} such that Ilx + tll - Ilxll + Iltll f o r every x E X.
Let us mention that when X is separable, these conditions are also equivalent to the following: for every finite dimensional subspace F of X and every e > 0, there is h E X\{0} such that for all x E F and ot E R, one has IIx + ~hll/> (1 - e)(llxll + IIc~hll). It is therefore clear that the above conditions define an "extreme" form of roughness. The canonical norm of 11 (N) provides the simplest example. PROOF. To prove that (ii) implies (i), we observe that the local reflexivity principle permits to construct, using t and a simple inductive argument, a subspace of X isomorphic to 11 (N). We simply outline the proof of (i) implies (ii). Starting from a subspace Y of X which is isomorphic to l l(N), we identify the bidual of Y with a subspace of X**, and we pick any completely singular t E Y**\{0}; in other words, we pick a nontrivial ultrafilter of N, and t is the weak* limit in X** along this ultrafilter of a sequence in X equivalent to the canonical basis of l 1(N). We observe now that a completely singular element t of (l l) **, when considered as a function on the Hilbert cube C (that is, on the unit ball of l ~ equipped with the weak* topology), has at every point of C a weak* oscillation equal to 211tll. Then we lift C to a minimal weak* compact subset K of B x , . Simple geometrical considerations permit to construct a dual unit ball B in X* such that if Osc(t) denotes the oscillation of t : ( B , weak*) --+ R and N = sup8 Itl, then O s c ( t ) ( f ) = 2N for all t E B. The proof is concluded by the following lemma: LEMMA 5.6. Let X be a Banach space and t E X**. The following are equivalent: (i) IIx + t ll--IIx II + Iltll f o r every x E X. (ii) / f Osc(t) denotes the oscillation o f t: ( B x , , weak*) --+ R, then O s c ( t ) ( f ) = 211tll f o r all f E B x , .
PROOF. Let 7" denote the infimum of all weak* continuous functions on B x , which are greater or equal of t. Since t is affine, 7"is concave upper semi-continuous, hence by H a h n Banach theorem, it is the infimum of the weak* continuous affine functions on B x , which are greater or equal to t. This translates into the following formula: ~'(f) -- inf{f(x) + [it - xl[; x E X} for every f E B x , . A straightforward computation now shows that if (i) is true, then 7"(f) Iltl] for every f E B x , , and (ii) follows. Conversely, if (ii) holds, given any e > 0 and any x E X, there is f0 E B x , such that f o ( x ) -- Ilxll, and by (ii) we can pick f E B x , "close to
Renormings of Banach spaces fo" suchthat t ( f ) > and (i) follows.
IltlI-E and f ( x ) >
Ilxll-e. Wehavethen f ( t + x ) >
819
Iltll + Ilxll-2E, []
Lemma 5.6 will be used in Section 6 when roughness of norms will be related with weak sequential completeness of certain spaces. Proposition 5.5 states a characterization of spaces containing l l (N) by an extremal form of roughness. Dually, it turns out that subspaces of c0(N) can be characterized by the existence of equivalent norms enjoying optimal properties of smoothness, as shown by the following THEOREM 5.7. Let X be a separable Banach space. The following are equivalent: (i) X is isomorphic to a subspace of co(N). (ii) There is an equivalent norm [I 911on X and k > 0 such that if (fn) is a sequence in Bx* which isweak* convergentto f andlimllfn - fl] = e > O, then Ilfll ~< 1 - k s . PROOF. (i) implies (ii) simply follows from the fact that the natural norm of c0(N) satisfies (ii) with k -- 1, and the property goes to subspaces with a straightforward proof. Let us mention that the canonical norm of c0(N) (and the equivalence) show that if (ii) can be done with some k > 0, then it can be done with k = 1. The proof of (ii) implies (i) is more involved. We first observe that (ii) clearly implies that the weak* and norm topologies agree on the dual unit sphere, from which it follows that X* is separable. We now outline the proof in the case when X* has a Schauder basis, that is, when X has a shrinking basis. Let (e~) be such a basis. We denote by En,k the linear span of {e j; n <. j <. k}. A duality argument shows that if (ii) is satisfied, there is a constant t > 0 such that for all weakly null sequence (x~) and all x, one has limsup IIx + xnll ~ max(llxll, (t) - 1 limsup IIx~ II).
(1)
Fix a summable sequence (Sn) $ O. The convergence in the left-hand side of (1) is uniform in x in the unit sphere of any finite dimensional space and (xn) gliding to infinity along the basis (e,,), so we can choose 1 = no < n l < n2 < -.. so that for each k ~> 1, if x is in the unit sphere of El,nk and y is in Enk+l,oo with Ilyll ~< t, then Ilx + yll ~< 1 + sk. Define the desired blocking {Fn } of {En} by setting Fk = Enk_l,nk for k = 1,2 . . . . . It is enough to check that {F2~ }nC~=l and {FZn-1 }~~1621 are both co-decompositions, since then X is the direct sum of the spaces (y~,, {Fzn }~--1)c0 and (y~.~ {F2~-1 }~~1621)c0. To check that, for example, {F2n }n~__j is a co-decomposition, it is sufficient to observe that if xk is in F2k and supk Ilxk [ ~< t, then for each m - 1, 2 . . . . for which II ~km_-I xk II ~> 1, the inequality
Xk ~< (1 -+-62m+1) k--1
Xk k=l
is true. This completes the proof when X has a shrinking basis. A straightforward modification of this proof provides the result under the assumption that X has a shrinking finite dimensional decomposition.
820
G. Godefroy
The general case can be deduced from the above argument by using the JohnsonRosenthal theorem which states that any Banach space X with separable dual has a subspace Y such that both Y and X~ Y have shrinking finite dimensional decompositions. The conclusion follows from the fact that (ii) goes to subspaces and quotient spaces, and that "being a subspace of c0(N)" is a three-space property. D Theorem 5.7 will be applied in Section 6 to Lipschitz isomorphisms. Renormings can sometimes be combined with special structures on Banach spaces, such as order structures. The following result characterizes "nice" lattices through the existence of special lattice norms. A Banach lattice is a Banach space X equipped with a lattice structure which is compatible with the vector lattice operations and such that Ilxll :
II Ix l II
(2)
for all x 6 X. A norm which satisfies (2) is called a lattice norm. Recall that a Banach lattice L is said to be order continuous if every downward directed net with infimum 0 converges in norm to 0. In other words, L satisfies a "Lebesgue dominated convergence theorem". Order continuity turns out to be equivalent with weak compactness of the order intervals. They are characterized among Banach lattices by the following result. THEOREM 5.8. Let X be a Banach lattice. Then X is order-continuous if and only if there is an equivalent lattice norm II 9 II on S such that the weak and norm topologies agree on the unit sphere o f II 9 II. PROOF. We only outline the arguments. If X is not order continuous there is a sequence (Yn) of disjoint positive vectors which is equivalent to the unit vector basis of c0(N) and y 6 X with Yn <<,Y for all n. Then for any lattice norm II 9 II on s we have lim IlYn - - Y II = Ily II, and (yn - y) tends weakly but not strongly to y. We outline the proof of the converse implication. If X = L 1(#), where # is a probability measure, the canonical injection from L2(#) into X has dense range, and the natural norm of L2(#) is a uniformly convex lattice norm. Now if we apply the formula of Remark 1.5.1) with Y = L2 (/z) and the canonical injection as T, the norm N is a locally uniformly convex lattice norm on X since clearly the computation of Remark 1.5.2) does not damage the lattice structure. The case of separable spaces V with an unconditional basis can be obtained in the same way through an injective map from 12(N) into V. The general case can be done by similar arguments, using Boolean algebras of projections. Fq Notes and comments. Proposition 5.1 is a simple and well-known exercise. Its relevance lies in the fact that outside the Hilbert space, functions that will be defined by series of squares of norms or of distances might be C 1 but will not be twice differentiable in general. Such functions are important, e.g., in the proof of Borwein-Preiss variational principle [ 19]. We refer to [5] for numerous characterizations of the Hilbert space. Theorem 5.2 follows, as many such characterizations, from Kwapiefi's theorem [121 ]. For the elementary proof of Theorem 5.2 we refer to [65]. Theorem 5.4 is shown in [144]. The main result of [ 144] is actually more general than Theorem 5.4 since it asserts the existence, on any
Renormings of Banach spaces
821
separable Banach space, of an equivalent norm such that any weakly relatively compact sequence which satisfies the condition (ii) of Theorem 5.4 is norm convergent. The proof of Theorem 5.4 relies in particular on the characterization of bidual types due to Maurey [135]. Maurey's work makes the link between Theorem 5.4 and Proposition 5.5, since [135] contains also a characterization of separable spaces containing l 1 by a property of any given norm: a separable space (X, [[ . II) contains an isomorphic copy of 11 if and only if there exists t 6 X** such that ]Ix + t l[ = [Ix - t l[ for every x 6 X, that is, if and only if there exists a symmetric bidual type. This result is in turn related with the ball topology which is defined and studied in [77], where it is shown for instance that if a Banach space X equipped with a given norm does not contain an isomorphic copy of I 1, then the weakest topology r for which the closed balls are r-closed coincides with a Hausdorff locally convex topology on the bounded subsets of X. This result asserts that spaces not containing l 1 have "quite smooth" norms, and it follows among other things that they have at most one isometric predual [77]. Proposition 5.5 is in [76], and the remark following it is in [77]. Theorem 5.7 is an isomorphic version from [78] of the nearly isometric characterization [114] of subspaces of co. Condition (ii), which asserts an optimal form of asymptotic smoothness of the norm, can be expressed with a smoothness modulus which goes back to [138]. This proof of Theorem 5.7 follows [107]. The original proof, which is closer to Kalton-Werner's approach, is more involved but provides a better control of isomorphism constants. Johnson-Rosenthal's "twisted sum" theorem is in [105]. Theorem 5.8 is shown and applied in [34]. We refer to [159] for the general theory of Banach lattices. The spaces co(F) have equivalent lattice 1.u.c. norms for arbitrary sets F , and this turns out to be important in nonseparable renorming theory (see [46, Section 7.1]) since linear injective maps into co(F) are frequently used in this context. The space L1 of any probability space has an equivalent lattice norm which is 1.u.c. and uniformly convex in every direction [122] (for definitions see, e.g., [176]). This can be obtained by the transfer method (see Remark 1.5.2). Note however that if F is uncountable, the space co(F) has no equivalent norm which is uniformly convex in every direction, so the result from [122] does not extend to arbitrary order continuous Banach lattices. Symmetric norms have also been investigated in great detail. One of their interesting features is that when they are defined on a separable sequence space with a symmetric basis, they lead to norms with similar properties on the relevant nonseparable spaces; Day's norm [37] is an early example of this phenomenon. Therefore the impossibility of certain nonseparable renormings implies that no symmetric norm exists with the given property on the relevant separable space (see [921). The few results we show in this chapter are far from exhausting the list of isomorphism classes which can be characterized in terms of equivalent norms. Let us recall for instance that by [60], separable Banach spaces are exactly the subspaces of l ~ (N) which can be given an equivalent UG-smooth norm. By [117], a separable space X has an equivalent weak-star uniformly Kadets-Klee norm (of power type) if and only if the Szlenk index of X is coo, and optimal quantitative results along these lines are established in [79]. An important problem, relevant to approximation property when bases are not available, is whether every separable space with the bounded approximation property can be renormed in order to have the metric approximation property. Since by [26] every separable space which has the metric approximation property satisfies it with a commuting sequence of
G. Godefroy
822
operators, this renorming problem is equivalent to asking if the bounded approximation property and its commutative version are equivalent. Nonseparable theory provides natural characterizations of "topologically nice" spaces in terms of norms. We mention for instance the recent converse to the classical AmirLindenstrauss theorem [6] shown in [113]: a Banach space of density character o91 has a projectional resolution of identity for every equivalent norm if and only if its dual unit ball is a Corson compact in the weak-star topology. Equivalences between topological properties of weak-star dual balls and special renormings of nonseparable Banach spaces is a very rich theory. For instance, the dual unit ball is a uniform Eberlein compact (for definitions see [ 15,7]) for the weak-star topology if and only if X has an equivalent UGnorm [60].
6. Some applications We present in this chapter various applications of renorming techniques: approximation of continuous functions by smooth ones or by differences of convex functions, linearization of Lipschitz isomorphisms between certain Banach spaces, weak sequential completeness of "extremely rough" spaces. The following result is usually called "Asplund averaging technique". THEOREM 6.1. If X has a separable dual there exists on X a locally uniformly convex equivalent norm whose dual norm is locally uniformly convex. PROOF. Theorem 6.1 can be obtained as an application of Baire category theorem on the set E of all equivalent norms on X. First we observe that this set E, when equipped with the topology of uniform convergence on bounded subsets of X, is homeomorphic to a complete metric space, and thus it is a Baire space. It turns out that the sets L and L, of equivalent 1.u.c. norms (and norms whose dual norm is 1.u.c.) contain dense G~ subsets of E, and thus by Baire category theorem their intersection is nonempty; in fact, "almost every" norm in the Baire sense belongs to L A L,. To check that L contains a dense G3 subset, we fix r0 6 L and for p E E we let
G(p, j ) -
{q ~ E; sup{ [ p 2 ( x ) + j - l r g ( x ) -
q2(x)l; r o ( x ) - 1} < j - 2 }
and for k E N, we set
Gk -- U { G ( p , j); p ~ E, j / > k} and finally
G-["qGk. k>~l It is obvious that Gk is open in E for all k. It is not difficult to check that Gk is dense for all k and that every norm in G is locally uniformly convex. The proof for L, is identical,
Renormings of Banach spaces
823
using the fact that the space E is h o m e o m o r p h i c to the space of equivalent norms on X* which are dual to a norm on X equipped with its natural topology, ff] The motivation for T h e o r e m 6.1 is that it leads to a simple proof of approximation of continuous functions by smooth ones. Note that Proposition 1.4 asserts that the following result applies in particular to spaces which have a separable dual. THEOREM 6.2. Let X be a Banach space which has a locally uniformly convex norm with locally uniformly convex dual norm, and let Y be an arbitrary Banach space. Then every continuous map from X to Y is a uniform limit on X o f C 1 smooth maps. PROOF. We prove it when Y = R. Let f be a continuous function from X to R. We denote fn = s u p { - n , inf(f, n)} the "truncation" of f at level n, and gk = fk+l -- fk. It is clear that f - ~ k = 0 gk pointwise. More precisely, given any x 6 X, there is s > 0 and n 6 N such that gk (Y) = 0 for all k ) n and all y such that [[y - x [1 < s. Moreover, if a function h is a uniform limit of a sequence (r,) of smooth functions, we m a y ensure that r, (t) -- 0 for all n if h(t) = 0 by considering (s, o rn), where the sn's are smooth real functions which vanish on a well-chosen neighbourhood of 0. It easily follows that it suffices to prove T h e o r e m 6.2 for b o u n d e d functions. Let [[ . [[ be an 1.u.c. n o r m on X with 1.u.c. dual norm, which exists by T h e o r e m 6.1. If f is a b o u n d e d function from X to R, we define h ( x ) = f ( x ) + [[x[[ 2. Since [1 . [[2 is C 1 smooth, it suffices to show that h is uniform limit of C 1 smooth functions. If we apply the above truncation procedure to h, the corresponding sequence (gk) consists of functions with bounded support. By the same token as before, such functions are uniform limits of smooth functions on X if they are uniform limits of such functions on b o u n d e d subsets of X. Summarizing, we have now that it suffices to show that b o u n d e d continuous functions on X are uniform limits on b o u n d e d sets of C 1 smooth functions. We now proceed to prove this latter fact. Let C be a closed convex subset of X. We define p (x) -- dist(x, C), and we claim that p is F - s m o o t h at every x E X \ C . Indeed, for any e ~> 0, denote
D ? p ( x ) -- { f ~ X*" p ( x + h) >~ p ( x ) + f (h) - s for all h E X ] . The n o n h o m o g e n e o u s version of L e m m a 1.14, which follows from a similar proof, says that p is F - s m o o t h at x if D o P ( X ) is a singleton {f} and if fn E D s with Sn tending to 0 implies that [[f , - f [1 tends to 0. But since p is a Lipschitz function with constant 1 and x r C, we have lim [[f , [[ = 1 and 1[f [[ = 1. M o r e o v e r (fn + f )
cD~p(x)
and thus lim ][fn + f [I = 2. Since the dual norm is 1.u.c. it follows that lim ]lfn - f l[ = 0 and thus p is F - s m o o t h on X \ C . If we consider now the function p2, it is easily seen to be F - s m o o t h on X, and even C 1 since it is convex. We now show that every continuous b o u n d e d function f on Sx is uniform limit on Sx of restrictions of C 1 smooth functions on X. Indeed assume that - 1 < f (x) < 1 for all
G. Godefroy
824
x ~ Sx, pick e > 0 and N E N such that 7 < 2Ne. For 0 ~< i ~< 2 N + 1, we consider the interval
Ii=
i-l-N N
'
and we let Oi - f - l ( I i ) . If Ci - 0 then g(x) - ~
i-N] N
If 1 ~< i ~< 2 N , we let Ci -c-o--fi-9(Bx\(Oi-1 U Oi g Q i + I ) ) . is s-close to f on Sx. Hence we may assume that Ci :/: 13 for
1 ~< i ~< 2N. For these i's, we define ri (x) - dist2(x, Ci). The functions ri create some kind of smooth partition o f u n i t y on Sx. Indeed pick xo ~ Sx, and i0 such that xo ~ Qio. Since f is continuous, there is ~ > 0 such that if IIx - yll < ~ and y E Sx then y ~ (Oi0-1 u Oio u Qio+l)- Since the n o r m of X is 1.u.c. it follows that xo ~ Cio, and thus rio (xo) > 0. Since xo was arbitrary, we have ~ 2 N 1 ri (x) > 0 for all x E Sx. if we let now ro(x) - (1 - IIx 112)2, we have that ro -- 0 on Sx and ro > 0 elsewhere. For 1 ~< i ~< 2 N and x E X, we define
h i ( x ) --
ro(x) + ri (x) 2N
Y~d=o rj (x ) The hi 's are C 1 functions on X and the function f is "nearly constant" on the support of each hi. It follows by an easy computation that if we denote 0/i the midpoint of each interval li, the function g(x) -- E 2 N 1 otihi(x) is C 1 smooth on X and we have I f ( x ) - g(x)l < s for every x ~ Sx. It suffices for concluding the proof to use a stereographic projection. Let us equip the space Y = X G R with the n o r m II (x, r) Ilr - (llxll 2 -+- r2) 1/2. This n o r m II 9 Ilr and its dual n o r m are clearly 1.u.c. We denote H -- {(x, 1); x 6 X} and we define p from Y\{0} to Sy by p ( y ) -- (llyll r) -1 y; let pl be the restriction of p to H . If f is a b o u n d e d continuous function from H to R and B is a closed b o u n d e d subset of H , we extend ( f o p 11), defined on the closed subset pl (B) of Sx, to a b o u n d e d continuous function f l on Sx. For any s > 0, there exists by the above a C 1 smooth function g on Y such that If(Y) - g(Y)l < s for every y E Sy. Hence I f - g o Pl] < s on B, and since g o pl is C 1 smooth on H this concludes the proof of T h e o r e m 6.2. D We now consider another kind of approximation, namely approximation of Lipschitz functions by differences of convex functions which are b o u n d e d on b o u n d e d sets. W h e n such an approximation is possible, one could say that the metric structure of the space is s o m e h o w "close" to the affine structure. This turns out to be a characterization of superreflexive spaces. THEOREM 6.3. Let X be a Banach space. The following assertions are equivalent: (i) X is superreflexive. (ii) Every Lipschitz function is uniform limit on X of a sequence of functions (gn) such
that f o r every n, gn = Cn - d n , where the Cn's and dn's are convex functions which are bounded on bounded sets.
Renormings of Banach spaces
825
PROOF. Assume (i). Let II 9 II be an equivalent uniformly convex norm on X provided by Theorem 3.2. We will show (ii) with uniform convergence on bounded subsets of X and indicate later how to obtain uniform convergence on X. Let f be a Lipschitz function on X. We may and do assume that the Lipschitz constant of f is 1. We define
fn(x)-
inf {f(y)-+- n(211xll 2 -+- 211Yll2 -IIx + yl12)}.
y6X
(1)
If we let cn(x) -- 2nllxll 2 and dn(x) = Cn(X) - f~(x) it is clear that the functions Cn and d~ are convex and that c~ is bounded on bounded sets. Moreover the sequence (fn) is uniformly convergent to f on bounded subsets of X. Indeed if we let x -- y in (1) we see that f (x) ~> f~ (x) for every x 6 X and every n 6 N. On the other hand, for every (x, y) 6 X 2 and every n 6 N one has 211xll 2 -4- 211yll2 - I I x -+- yll 2 ~ ( l l x l l - Ilyll) 2 ~ 0,
(2)
hence the sequence of functions (fn) is increasing. Let now n ~> 3 and y 6 X be such that
f ( Y ) + n(211x II2 -+-2llyll 2 - IIx -+-yll 2) ~ f ( x ) .
(3)
It follows from (2) and (3) that n ( l l x l l - Ilyll) 2 ~ f ( x ) - f ( y ) <~ IIx - yll and through an easy computation this implies that Ilyll ~ 2(1 + Ilxll),
(4)
therefore we may assume that y satisfies (4) in the computation of fn (x) and it follows that fn and therefore dn are bounded on bounded sets. Moreover, it follows from (4) that under the condition (3) one has 211xll 2 + 211ylle - I I x -4- yll 2 ~ (3/n)(1 + Ilxll).
(5)
Since the norm II 9II is uniformly convex, this implies (see Definition 1.1.) that for n large enough, fn is uniformly close to f on bounded sets. This sequence (fn) converges to f uniformly on X when the modulus of convexity of the norm is of power type 2 (e.g., when the norm is Hilbertian). In general however, a modification of the formula is needed. Assume that the norm II 9 II is uniformly convex with a modulus of power type p for some p i> 2. Such a norm exists by Theorem 3.6. Then the sequence of functions defined by
gn(x) -- inf { f ( y ) + n(2 p-111xl[ p + 2 p-I Ilyl] p - Ilx + yIIP)} y~X
converges to f uniformly on X. The computations are similar to the above but slightly more involved.
G. Godefroy
826
We only sketch the proof of (ii) implies (i). The following is true: if X is not superreflexive there exists a 1-Lipschitz function defined on X such that for every pair {c, d} of continuous convex functions bounded on Bx one has
suplf(x)
- (c - d ) ( x ) I >1 1/4.
Bx
When X = c0(N) the function f can be constructed as follows: let T be the set consisting of all finite sequences with values in {-1, 1}, which we embed into c0(N) in the obvious way. Let E be the subset of T consisting of sequences with even length. Then the function f defined on c0(N) by f ( x ) = dist(x, E) works. The idea of the proof is that T is an infinite dyadic tree, and that f oscillates between 0 and 1 on each branch of T. Assuming that f = c - d with c and d convex and following an appropriate branch contradicts the boundedness of c and d. The general case of nonsuperreflexive spaces is obtained by a similar trick, using the fact that arbitrarily large finite trees grow in the unit ball of such spaces. [] REMARK 6.4. Denote by Convb(X) the cone of convex functions which are bounded on bounded sets, by UCb(X) the space of functions on X which are uniformly continuous on bounded sets, and by rb the topology of uniform convergence on bounded sets. Theorem 6.3 implies that X is superreflexive if and only if the rb-closed linear span of Convb (X) is equal to UCb (X). We continue with more applications of renormings to nonlinear theory. The next result relies on the characterization of subspaces of c0(N) given in Theorem 5.7. We recall that two Banach spaces X and Y are Lipschitz isomorphic if there is a bijective map F from X onto Y such that both F and F - 1 are Lipschitz maps. THEOREM 6.5. Let Y be a Banach space which is Lipschitz isomorphic to a linear subspace of c0(N). Then Y is linearly isomorphic to a subspace of c0(N). If Y is Lipschitz isomorphic to c0(N) then Y is linearly isomorphic to c0(N). PROOF. We only outline the arguments. The main topological tool we need for proving this result is Gorelik's principle, which provides a substitute for the lack of weak continuity of Lipschitz isomorphisms between Banach spaces X and Y. This principle can be stated as follows: Let U be a Lipschitz isomorphism from X onto Y, and let k be the Lipschitz constant of U -1 . Let X0 be a subspace of finite codimension of X. Let d > 0 and b > 0 be l such that d > kb. Then there exists a compact subset K of Y such that bByf C U(2dBxo) + K, where B I denotes the open unit ball. This topological result is an application of the Bartle-Graves selection theorem and of Brouwer's fixed point theorem. Gorelik's theorem permits to show the following "quasi-linearity" result: if (fn) is a weak* null sequence in Y*, then U(Bxo) asymptotically norms, up to a constant independent of X0 and (fn), the sequence (fn). In other words, when n is large enough, suPv(Bx0) Ifnl controls Ilfnll. This simply follows from Gorelik's principle and the fact that (fn) goes uniformly to 0 on K.
Renormings of Banach spaces
827
Recall that Theorem 5.7. provides a characterization of subspaces of c0(N) through the behaviour of weak* null sequences in the dual space. Assuming that X is a linear subspace of c0(N) and U is a Lipschitz isomorphism from X onto Y, it suffices therefore to construct an equivalent norm N on Y which satisfies the assumptions of this Theorem. This equivalent norm is provided by the following "maximal rate of change" formula: if f E Y*, define N(f)
- supJ I f ( U x - U x ' ) l ; (X,X') E X 2 x ~;kxt} I IIx - x' II
This norm N is an equivalent dual norm since it is clearly weak* 1.s.c. Moreover, N satisfies the assumptions of Theorem 5.7. For checking this point, we pick (in the notation of Theorem 5.7.) x and x t in X which nearly realize the supremum in the definition of N ( f ) , and we may assume without loss of generality that x = - x I and U (x) = - U (xl). Then we use the fact that since X is a subspace of c0(N), the set M of metric midpoints between x and x ~ essentially contains a ball of a finite codimensional vector subspace. Moreover the choice of x implies that f (t) is small for every t E M. On the other hand, Gorelik's principle implies that the set U (M) asymptotically norms the weak* null sequence ( f - fn) with vectors U(xn). Computing ( f - ( f - f~), U ( x ) - U(Xn)) shows that the norm N is asymptotically nearly additive on f and ( f - f , ) since the cross products are small, and this shows our claim and the first part of the theorem. The second part of the result follows from the first, since by a theorem of Heinrich and Mankiewicz, being a/Z or is stable under Lipschitz isomorphisms, and by a theorem of Johnson and Zippin every s162 subspace of c0(N) is linearly isomorphic to c0(N). D We conclude our section with a linear result. Rough norms were applied in Section 2 to prove the "harmonic" behaviour of certain smooth functions (see Proposition 2.9). The next statement shows that some extreme form of roughness bears applications in a different direction. PROPOSITION 6.6. Let X be a Banach space such that there exists a projection P f r o m the bidual X** onto X such that Ilull = Ilu - P(u)ll + IlP(u)ll f o r every u E X**. Then X is weakly sequentially complete.
PROOF. Pick u E X * * \ X , and denote t = u - P ( u ) . Our assumption on use L e m m a 5.6., which shows in particular that the restriction of t to B x , weak* continuity. Now Baire's theorem implies that t is not the pointwise a sequence of weak* continuous functions. In particular, it is not limit of sequence from X, and thus weak sequential completeness of X is shown.
P allows us to has no point of limit on B x , of a weak Cauchy D
Of course, weakly sequentially complete spaces cannot in general be renormed so as to satisfy the assumption of Proposition 6.6 since, for instance, they are in general not even complemented in their bidual. However, many natural examples of weakly sequentially complete spaces equipped with their natural norms turn out to satisfy this assumption. For instance, many natural quotient spaces of L 1 satisfy it.
828
G. Godefroy
Notes and comments. Theorem 6.1 is an example of Asplund's averaging technique which originated in [9]. The Baire category approach we follow here comes from [64]. It should be noted that the set of equivalent 1.u.c. norms is topologically quite complicated; in particular, it is not a Borel set for the natural topology on the set E of equivalent norms [22]. Dually, the set of C 1 smooth norms [22] on a separable Asplund space, or of C 1 smooth convex functions on a separable space [174], is not a Borel set. Many natural properties of norms or convex sets lead to residual sets, in other words, these properties are satisfied for almost every norm in the Baire category sense (see, e.g., [72]). Theorem 6.2 is shown in [172], and it has been extended further in [70] and [152]. It is shown in these two works that if X has a equivalent 1.u.c. norm, or alternatively if X has the Radon-Nikodym property, then X has Ck-smooth partitions of unity (where k 6 N or k - e~) provided that every equivalent norm is uniform limit on bounded sets of C ~smooth functions. The proofs rely on particular on the result that if X has the RadonNikodym property or if X has an equivalent 1.u.c. norm, then the norm topology of X has a a-discrete basis made up of convex sets. It is apparently not known whether every Banach space satisfies this conclusion (or the weaker one of existence of such a a-locally finite basis). Such results are important in the nonseparable theory, where a major problem is to know whether every Banach space which has a C 1-smooth equivalent norm has C 1-smooth partitions of unity. Several partial positive answers are available (see [176]) but the general conjecture still holds. It is not even known whether every space with a C 1 smooth norm is such that the set of C 1 smooth norms is dense in the set of all equivalent norms. Finally, no example is known of an Asplund space without C 1-smooth partitions of unity. Theorem 6.3 is a recent result of [28]. Its motivation is to relate as closely as possible convex and Lipschitz functions, and it asserts that this is possible in super-reflexive spaces and only there. So for instance there is no way to reduce analytic approximation of Lipschitz functions to approximation of norms in the nonreflexive case, and other methods have to be used for showing this approximation result in c0(N); this is done in [31 ] and independently in [71 ]. It is an intriguing fact that no simple formula exists so far which would provide uniform approximation by C 2 or analytic functions and which would replace the use of partitions of unity. The existence of distorted norms in minimal super-reflexive Banach spaces (such as I p with 1 < p < oo) follows easily by Theorem 6.3 from the existence of a distorted Lipschitz function. Much more precise results than Theorem 6.3 are actually shown in [28]. For instance, the uniform convergence on bounded sets of the sequence ( f n ) given by Eq. (1) in the proof is actually equivalent to the uniform convexity of the norm, and locally uniformly convex renormings provide with the same formula uniform approximation on compact sets. On the other hand, there is a Lipschitz function on loc (N) which is not even a pointwise limit of a sequence of differences of convex functions. This follows from the fact that continuous convex functions are weakly 1.s.c. and in particular weakly Borel, while the norm and weak Borel structure on loc (N) are distinct [ 166]. More characterizations of super-reflexive spaces are shown by Cepedello [29,30]: for instance, X is super-reflexive if and only if for every closed subset F of X, there is a difference of two convex functions f = c - d such that f attains its strong minimum exactly on F. Remark 6.4 is from [28].
Renormings of Banach spaces
829
Theorem 6.5 is the main result of [78] and the proof is taken from this paper. Gorelik's principle, which is one of the main tools in the proof, originates in [86] and has been expanded in [ 106]. Among the milestones of nonlinear geometry of Banach spaces, one should mention [52,154,96] and [106]. We refer to [14] for an up-to-date account of the whole theory. The maximal rate of change argument which lies behind the definition of the equivalent norm is related with D. Preiss' theorem on the differentiability of real-valued Lipschitz maps on Asplund spaces [150] and its proof. Johnson-Zippin's theorem on s subspaces of c0(N) is in [108]. Theorem 6.5 somehow means that c0(N) is the "smoothest" of all / ~ spaces, and under this form it can be extended in various ways; for instance, c0(N) is the only separable s space which can be equivalently renormed into an M-ideal in its bidual ([82]; see [93]). However natural conjectures in this direction remain unsolved; for instance, it is not known whether c0(N) is the only s separable space which has Petczyfiski's property (u). Nonlinear theory of Banach spaces contains many open questions. To mention a few of them, it is not known whether a Banach space which is Lipschitz isomorphic to 11 (N) is linearly isomorphic to it, nor whether a space which is uniformly homeomorphic to c0(N) is linearly isomorphic to it. Moreover, no example is known of a couple of separable Banach spaces which are Lipschitz isomorphic but not linearly isomorphic. Separable examples for uniform homeomorphisms [ 155,1], and nonseparable examples for Lipschitz isomorphisms [2], are available. Proposition 6.6 is a simple application of Baire's lemma and goes back to [75]. It has been used in a variety of situations; we refer to [93, Chapter IV] where many of these applications are displayed, and for a more recent work to [80]. It is shown in [75] that the quotient of L 1 by a subspace X whose unit ball is closed in L 1 equipped with the topology of convergence in measure is weakly sequentially complete. This applies in particular to the Hardy space X = HI(D), considered as a subspace of L 1(T), and provides a proof of the Mooney-Havin theorem. Many natural generalizations of the Mooney-Havin theorem can be obtained by the same token (see, e.g., [93]). Proposition 6.6 itself can be widely generalized: for instance, any space X such that there exists a projection P from X** onto X such that I I I - 2PII = 1 is weakly sequentially complete; this follows from [135] or from [77]. We refer to [81] for the related notion of u-ideal and its applications. As mentioned before, there are weakly sequentially complete Banach spaces which are not complemented in their bidual (e.g., the kernel of a quotient map from 11 onto co [129]), and thus Proposition 6.6 as stated has no converse. However, it can be shown that under the assumption of Proposition 6.6, if (xn) is a sequence in X and a 6 R are such that for every f ~ X*, one has limn sup{If(x1 - xk)l; I, k ~> n} ~< al[fll, then dist(t, X) ~< a for every weak-star cluster point t to (xn) in (X**, w*). It is not clear which weakly sequentially complete spaces can be renormed to have such a quantitative form of weak sequential completeness.
Acknowledgement I am glad to thank V~iclav Zizler, who introduced me to renorming theory, and Mari~in Fabian for their help in preparing this chapter.
830
G. Godefroy
References [1] I. Aharoni and J. Lindenstrauss, An extension of a result of Ribe, Israel J. Math. 52 (1985), 50-64. [21 I. Aharoni and J. Lindenstrauss, Uniform equivalence between Banach spaces, Bull. Amer. Math. Soc. 84 (1978), 281-283. [3] R. Alencar, R.M. Aron and S. Dineen, A reflexive space of holomorphic functions in infinitely many variables, Proc. Amer. Math. Soc. 90 (1984), 407--411. [41 J. Alonso and C. Benitez, Some characteristic and non-characteristic properties of inner product spaces, J. Approx. Theory 55 (1988), 318-323. [5] D. Amir, Characterizations of Inner Product Spaces, Operator Theory: Advances and Applications, Vol. 20, Birkh~iuser (1986). [6] D. Amir and J. Lindenstrauss, The structure of weakly compact sets in Banach spaces, Ann. of Math. 88 (1968), 35-44. [7] S. Argyros and V. Farmaki, On the structure of weakly compact subsets of Hilbert spaces and applications to the geometry of Banach spaces, Trans. Amer. Math. Soc. 289 (1985), 409-427. [8] E. Asplund, Fr~chet differentiability of convex functions, Acta Math. 121 (1968), 31-47. [91 E. Asplund, Averaged norms, Israel J. Math. 5 (1967), 227-233. [lO1 J. P. Aubin and I. Ekeland, Applied Linear Analysis, Wiley, New York (1984). [11] D. Azagra and T. Dobrowolski, Smooth negligibility of compact sets in infinite-dimensional Banach spaces, Math. Ann. 312 (1998), 445-463. [12] B. Beauzamy, Introduction to Banach Spaces and their Geometry, 2nd edn, North-Holland Mathematics Studies, Vol. 68 (1985). [13] C. Benitez, K. Przelawski and D. Yost, A universal modulus for normed spaces, Studia Math. 127 (1998), 21-46. [14] Y. Benyamini and J. Lindenstrauss, Geometric Nonlinear Functional Analysis, Vol. 1, Colloquium Publications, Vol. 48, Amer. Math. Soc. (2000). [15] Y. Benyamini and T. Starbird, Embedding weakly compact sets into Hilbert spaces, Israel J. Math. 23 (1976), 137-141. [16] C. Bessaga, Every infinite-dimensional Hilbert space is diffeomorphic with its unit sphere, Bull. Acad. Pol. Sci. 14 (1966), 27-31. [17] C. Bessaga and A. Petczyliski, Selected Topics in Infinite Dimensional Topology, Polish Scientific Publishers, Warszawa (1975). [18] R. Bonic and J. Frampton, Smooth functions on Banach manifolds, J. Math. Mech. 15 (1966), 877-898. [19] J.M. Borwein and D. Preiss, A smooth variational principle with applications to subdifferentiability and to differentiability of convex functions, Trans. Amer. Math. Soc. 303 (1987), 517-527. [20] B. Bossard, Codage des espaces de Banach s~parables. Familles analytiques ou conanalytiques d'espaces de Banach, Note aux C. R. Acad. Sci. Paris 316 (1993), 1005-1010. [21] B. Bossard, Thkse, Universit6 de Paris VI (1994). [22] B. Bossard, G. Godefroy and R. Kaufman, Hurewicz's theorems and renormings of Banach spaces, J. Funct. Anal. 140 (1996), 142-150. [23] J. Bourgain, The Szlenk index and operators on C(K) spaces, Bull. Soc. Math. Belgique 31 (1979), 87117. [24] R.D. Bourgin, Geometric Aspects of Convex Sets with the Radon-Nikodym Property, Lecture Notes in Math. 993, Springer-Verlag (1983). [25] E Cabello Sanchez, Regards sur le problbme des rotations de Mazur, Extracta Mathematicae 12 (1997), 97-116. [26] P. Casazza and N.J. Kalton, Notes on approximation properties in separable Banach spaces, REX. Mtiller and W. Schachermayer, eds, London Math. Soc. Lecture Notes 158, Cambridge Univ. Press (1990), 49-63. [27] P. Casazza and T.J. Shura, Tsirelson's Space, Lecture Notes in Math. 1363, Springer-Verlag (1989). [28] M. Cepedello, Approximation of Lipschitz functions by A-convex functions in Banach spaces, Israel J. Math. 106 (1998), 269-284. [29] M. CepedeUo, On regularization in superreflexive Banach spaces by infimal convolution formulas, Studia Math. 129 (1998), 265-284.
Renormings o f Banach spaces
831
[30] M. Cepedello, Two characterizations of super-reflexive Banach spaces by the behaviour of differences of convex functions, Preprint (1999). [31] M. Cepedello and E H~ijek, Analytic approximations of uniformly continuous operators, Journal of Mathematical Analysis and Applications (2001), to appear. [32] G. Choquet, Lectures on Analysis, W.A. Benjamin, New York (1969). [33] W. Davis, D.J.H. Garling and N. Tomczak-Jaegermann, The complex convexity of quasi-normed spaces, J. Funct. Anal. 55 (1984), 110-150. [34] W. Davis, N. Ghoussoub and J. Lindenstrauss, A lattice renorming theorem and applications to vectorvalued processes, Trans. Amer. Math. Soc. 263 (1983), 531-540. [35] W. Davis and W.B. Johnson, A renorming of non-reflexive Banach spaces, Proc. Amer. Math. Soc. 37 (1973), 386-489. [36] M.M. Day, Normed Linear Spaces, 3rd edn, Springer-Verlag (1973). [37] M.M. Day, Strict convexity and smoothness ofnormed spaces, Trans. Amer. Math. Soc. 78 (1955), 516528. [38] G. Debs, G. Godefroy and J. Saint Raymond, Topological properties of the set of norm-attaining linear functionals, Canadian J. Math. 47 (1995), 318-329. [39] R. Deville, A characterization of C~-smooth Banach spaces, Proc. London Math. Soc. 22 (1990), 13-17. [40] R. Deville, Geometrical implications of the existence of very smooth bump functions in Banach spaces, Israel J. Math. 6 (1989), 1-22. [41 ] R. Deville and M. Fabian, Principes variationnels et diff~rentiabilit~ d'applications d~finies sur un espace de Banach, Publ. Math. Besan~on 10 (1989), 79-102. [42] R. Deville, V. Fonf and E H~ijek, Analytic and C k approximations of norms in separable Banach spaces, Studia Math. 120 (1996), 61-74. [43] R. Deville, V. Fonf and E H~ijek, Analytic and polyhedral approximation of convex bodies in separable polyhedral spaces, Israel J. Math. 105 (1998), 139-154. [44] R. Deville and N. Ghoussoub, Perturbed minimization principles and applications, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 393-435. [45] R. Deville, G. Godefroy, D. Hare and V. Zizler, Differentiability of convex functions and the convex point of continuity property in Banach spaces, Israel J. Math. 59 (1987), 245-255. [46] R. Deville, G. Godefroy and V. Zizler, Smoothness and renormings in Banach spaces, Pitman Monographs and Surveys, Vol. 64, Longman (1993). [47] R. Deville, R. Gonzalo and J.A. Jaramillo, Renormings of Lp(Lq), Math. Proc. Cambridge Phil. Soc. 126 (1999), 155-169. [48] J. Diestel, Geometry of Banach Spaces - Selected Topics, Lecture Notes in Math. 485, Springer-Verlag (1975). [49] J. Diestel, Sequences and Series in Banach Spaces, Graduate Texts in Math., Springer-Verlag (1984). [50] J. Diestel and J.J. Uhl, Vector Measures, Math. Surveys, Vol. 15, Amer. Math. Soc. (1977). [51] I. Ekeland and G. Lebourg, Generic Fr~chet differentiability and perturbed optimization problems in Banach spaces, Trans. Amer. Math. Soc. 224 (1976), 193-216. [52] E Enflo, On the nonexistence of uniform homeomorphisms between LP-spaces, Ark. Mat. 8 (1969), 103105. [53] E Enflo, Banach spaces which can be given an equivalent uniformly convex norm, Israel J. Math. 13 (1972), 281-288. [54] E Enflo, J. Lindenstrauss and G. Pisier, On the "Three space problem", Math. Scand. 36 (1975), 189-210. [55] M. Fabian, Lipschitz-smooth points of convex functions and isomorphic characterizations of Hilbert spaces, Proc. London Math. Soc. 51 (1985), 113-126. [56] M. Fabian, Subdifferentiability and trustworthiness in the light of a new variational principle of Borwein and Preiss, Acta Univ. Carolinae 30 (1989), 51-56. [57] M. Fabian, On a dual locally uniformly rotund norm on a dual Va~dk space, Studia Math. 101 (1991), 69-81. [58] M. Fabian, Differentiability of Convex Functions and Topology- Weak Asplund Spaces, Wiley (1997). [59] M. Fabian, On an extension of norms from a subspace to the whole Banach space keeping their rotundity, Studia Math. 112 (1995), 203-211.
832
G. Godefroy
[60] M. Fabian, G. Godefroy and V. Zizler, The structure of uniformly G~teaux smooth Banach spaces, Israel J. Math., to appear. [61] M. Fabian, P. H~jek and V. Zizler, Uniform Eberlein compacta and uniformly G~teaux smooth norms, Serdica Math. J. 23 (1997), 351-362. [62] M. Fabian, E Habala, E H~jek, V. Montesinos, J. Pelant and V. Zizler, Functional Analysis and Infinite Dimensional Geometry, Canad. Math. Society Book, Springer-Verlag (2001), to appear. [63] M. Fabian, J.H.M. Whitfield and V. Zizler, Norms with locally Lipschitzian derivatives, Israel J. Math. 44 (1983), 262-276. [64] M. Fabian, L. Zajf6ek and V. Zizler, On residuality of the set of rotund norms on a Banach space, Math. Ann. 258 (1981/82), 349-351. [65] M. Fabian and V. Zizler, An elementary approach to some questions in higher order smoothness in Banach spaces, Extracta Mathematicae, to appear. [66] W. Feller, An Introduction to Probability Theory and its Applications, Vol. II, 2nd edn, Wiley (1971). [67] T. Figiel and W.B. Johnson, The approximation property does not imply the bounded approximation property, Proc. Amer. Mat. Soc. 41 (1973), 197-200. [68] C. Finet, Uniform convexity properties of norms on a superreflexive Banach space, Israel J. Math. 53 (1986), 81-92. [69] V. Fonf, J. Lindenstrauss and R.R. Phelps, Infinite dimensional convexity, Handbook of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 599-670. [70] J. Frontisi, Smooth partitions of unity in Banach spaces, Rocky Mountain J. Math. 25 (1995), 1295-1304. [71] R. Fry, Analytic approximation on c 0, J. Funct. Anal. 158 (1998), 509-520. [72] E Georgiev, Mazur's intersection property and a Krein-Milman type theorem for almost all closed, convex and bounded subsets ofa Banach space, Proc. Amer. Math. Soc. 104 (1988), 157-164. [73] N. Ghoussoub, G. Godefroy, B. Maurey and W. Schachermayer, Some topological and geometrical structures in Banach spaces, Mem. Amer. Math. Soc. 378 (1987). [74] G. Godefroy, Existence de normes trks lisses sur certains espaces de Banach, Bull. Sci. Math. 2 106 (1982), 63-68. [75] G. Godefroy, Sous-espaces bien disposes de L 1. Applications, Trans. Amer. Mat. Soc. 286 (1984), 227249. [76] G. Godefroy, Metric characterizations of first Baire class functions and octahedral norms, Studia Math. 95 (1989), 1-15. [77] G. Godefroy and N.J. Kalton, The ball topology and its applications, Contemporary Math. 85 (1989), 195-238. [78] G. Godefroy, N.J. Kalton and G. Lancien, Lipschitz isomorphisms and subspaces of c0, Geom. Funct. Anal. 10 (2000), 798-820. [79] G. Godefroy, N.J. Kalton and G. Lancien, Szlenk indices and uniform homeomorphisms, Trans. Amer. Math. Soc., to appear. [80] G. Godefroy, N.J. Kalton and D. Li, On subspaces of L 1 which embed into l 1, J. Reine Angew. Math. 471 (1996), 43-75. [81] G. Godefroy, N.J. Kalton and ED. Saphar, Unconditional ideals in Banach spaces, Studia Math. 104 (1993), 13-59. [82] G. Godefroy and D. Li, Some natural families of M-ideals, Math. Scand. 66 (1990), 249-263. [83] G. Godefroy, S. Troyanski, J. Whitfield and V. Zizler, Smoothness in weakly compactly generated Banach spaces, J. Funct. Anal. 52 (1983), 344-352. [84] G. Godefroy, S. Troyanski, J. Whitfield and V. Zizler, Locally uniformly rotund renormings and injections into co(F), Canadian Math. Bull. 27 (1984), 494-500. [85] R. Gonzalo, M. Gonzalez and J.A. Jaramillo, Symmetric polynomials on function spaces, J. London Math. Soc. 59 (1999), 681-697. [86] E. Gorelik, The uniform nonequivalence of L p and 1p, Israel J. Math. 87 (1994), 1-8. [87] W.T. Gowers, A Banach space not containing co, ll or a reflexive subspace, Trans. Amer. Math. Soc. 344 ( 1994 ), 407-420. [88] W.T. Gowers, Ramsey theory methods in Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published).
R e n o r m i n g s o f B a n a c h spaces
833
[89] S. Guerre-Delabribre and M. Levy, Espaces 1p dans les sous-espaces de L 1, Trans. Amer. Math. Soc. 279 (1983), 611-616. [90] E Habala, P. H~jek and V. Zizler, Introduction to Banach Spaces, I, II, Matfyspress, Prague (1996). [91] P. H~ijek, Smooth functions on co, Israel J. Math. 104 (1998), 17-27. [92] P. H~ijek and V. Zizler, Remarks on symmetric smooth norms, Bull. Austral. Math. Soc. 52 (1995), 225229. [93] E Harmand, D. Werner and W. Werner, M-Ideals in Banach Spaces and Banach Algebras, Lecture Notes in Math. 1547, Springer-Verlag (1993). [94] R. Haydon, Normes inddfiniment diffdrentiables sur certains espaces de Banach, Note aux C. R. Acad. Sci. Paris 315 (1992), 1175-1178. [95] R. Haydon, Trees in renorming theory, Proc. London Math. Soc. 78 (1999), 541-584. [96] S. Heinrich and E Mankiewicz, Applications of ultrapowers to the uniform and Lipschitz classification of Banach spaces, Studia Math. 73 (1982), 225-25I. [97] E Holick3), M. Smfdek and L. Zajf6ek, Convex functions with a non-Borel set of G~teaux differentiability points, Comment. Math. Univ. Carol. 39 (1998), 469-482. [98] A. Ioffe, Variational methods in local and global non-smooth analysis, Nonlinear Analysis, Differential Equations and Control, EH. Clarke and R.J. Stern, eds, Kluwer (1999), 447-502. [99] R.C. James, Uniformly nonsquare Banach spaces, Ann. of Math. 80 (1964), 542-550. [100] R.C. James, Super-reflexive Banach spaces, Canadian J. Math. 24 (1972), 896-904. [101] M. Jimenez and J.E Moreno, Renorming Banach spaces with the Mazur intersection property, J. Funct. Anal. 144 (1997), 486-504. [102] K. John and V. Zizler, A short proof of a version of Asplund averaging theorem, Proc. Amer. Math. Soc. 73 (1979), 277-278. [103] W.B. Johnson, A reflexive space which is not sufficiently Euclidean, Studia Math. 60 (1976), 201-205. [104] W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84. [105] W.B. Johnson and H.P. Rosenthal, On weak*-basic sequences and their applications to the study of Banach spaces, Israel J. Math. 9 (1972), 77-92. [106] W.B. Johnson, J. Lindenstrauss and G. Schechtman, Banach spaces determined by their uniform structure, Geom. Funct. Anal. 6,3 (1996), 430-470. [107] W.B. Johnson, J. Lindenstrauss, D. Preiss and G. Schechtman, Affine approximation of Lipschitz maps between infinite dimensional Banach spaces, to appear. [108] W.B. Johnson and M. Zippin, On subspaces of quotients of (Y-~ Gn)lP and (Y-~ Gn)co, Israel J. Math. 13 (1972), 311-316. [109] M.I. Kadets, On weak and norm convergence, Dokl. Akad. Nauk SSSR 122 (1958), 13-16. [110] M.I. Kadets, On spaces isomorphic to locally uniformly rotund spaces, Izv. Vyss. Uc. Zav. Matem. 1 (1959), 51-57 and 1 (1961), 186-187. [ 111] M.I. Kadets, Proof of topological equivalence of separable infinite dimensional Banach spaces, Funct. Anal. Appl. 1 (1967), 53-62. [112] M.I. Kadets, Conditions of differentiability of the norm of a Banach space, Uspekhi Mat. Nauk 20 (1965), 183-187. [113] O. Kalenda, Valdivia compacta and equivalent norms, Studia Math. 138 (2) (2000), 179-191. [114] N.J. Kalton and D. Werner, Property (M), M-ideals and almost isometric structure of Banach spaces, J. Reine Angew. Math. 461 (1995), 137-178. [ 115] R. Kaufman, Topics on analytic sets, Fund. Math. 139 (1991), 215-220. [116] V.L. Klee, Mappings into normed linear spaces, Funct. Math. 49 (1960), 25-34. [117] H. Knaust, E. Odell and Th. Schlumprecht, On asymptotic structure, the Szlenk index and UKK properties in Banach spaces, Positivity 3 (1999), 173-199. [118] J.L. Krivine and B. Maurey, Espaces de Banach stables, Israel J. Math. 39 (1981), 273-295. [119] J. Kurzweil, On approximation in real Banach spaces, Studia Math. 14 (1954), 213-231. [120] J. Kurzweil, On approximation in real Banach spaces by analytic operations, Studia Math. 16 (1957), 124-129.
834
G. Godefroy
[ 121] S. Kwapiefi, Isomorphic characterization of inner product spaces by orthogonal series with vector valued coefficients, Studia Math. 44 (1972), 583-595. [122] D. Kutzarova and S. Troyanski, On equivalent lattice norms which are uniformly convex or uniformly differentiable in every direction in Banach lattices with a weak unit, Serdica Bulgaricae Math. Publ. 9 (1983), 249-262. [123] G. Lancien, Dentability indices and locally uniformly convex renormings, Rocky Mountain J. Math. 23 (1993), 635-647. [124] G. Lancien, On the Szlenk index and the weak*-dentability index, Quart. J. Math. Oxford (2) 47 (1996), 59-71. [ 125] E.B. Leach and J.H.M. Whitfield, Differentiable functions and rough norms on Banach spaces, Proc. Amer. Math. Soc. 33 (1972), 120-126. [ 126] M. Leduc, Densit~ de certaines familles d'hyperplans tangents, C. R. Acad. Sci. Paris, S6rie A 270 (1970), 326-328. [127] J. Lindenstrauss, On operators which attain their norm, Israel J. Math. 1 (1963), 139-148. [128] J. Lindenstrauss, On the modulus of smoothness and divergent series in Banach spaces, Michigan Math. J. 10 (1963), 241-252. [129] J. Lindenstrauss, On a certain subspace ofl 1, Bull. Acad. Pol. Sci. 12 (1964), 539-542. [130] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces, I and II, Springer-Verlag (1977 and 1979). [131] A.R. Lovaglia, Locally uniformly convex Banach spaces, Trans. Amer. Math. Soc. 78 (1955), 225-238. [132] R. Maleev and S. Troyanski, Smooth norms in Orlicz spaces, Canad. Math. Bull. 34 (1991), 74-82. [133] J. Matou~ek and E. Matou~kov~, A highly nonsmooth norm on Hilbert space, Israel J. Math. 112 (1999), 1-27. [134] E. Matou~kov~, An almost nowhere Fr~chet smooth norm on superreflexive spaces, Studia Math. 133 (1999), 93-99. [135] B. Maurey, Types and ll-subspaces, Longhorn Notes, Texas Functional Analysis Seminar (1982-83), 123-137. [136] B. Maurey and G. Pisier, S~ries de variables al~atoires vectorielles ind~pendantes et propri~t~s g~om~triques des espaces de Banach, Studia Math. 58 (1976), 45-90. [137] D.E Milman, On some criteria for the regularity of spaces of the type (B), Dokl. Akad. Nauk SSSR 20 (1938), 243-246. [138] V.D. Milman, Geometric theory of Banach spaces. 2. Geometry of the unit ball, Uspehi Mat. Nauk 26 (1971), 6(162), 73-149 (Russian). English translation: Russian Math. Surveys 26 (1971), 6, 79-163. [ 139] A. Molto, J. Orihuela and S. Troyanski, Locally uniformly rotund renorming and fragmentability, Proc. London Math. Soc. 75 (1997), 619-640. [140] I. Namioka, R.R. Phelps and D. Preiss, Smooth Banach spaces, weak Asplund spaces and monotone or usco mappings, Israel J. Math. 72 (1990), 257-279. [ 141 ] A.M. Nemirovski and E.M. Semenov, On polynomial approximation in function spaces, Mat. Sbornik 21 (1973), 255-277. [142] G. N6rdlander, The modulus of convexity in normed linear spaces, Ark. Mat. 4 (1960), 15-17. [143] E. Odell and Th. Schlumprecht, Distorsion, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [144] E. Odell and Th. Schlumprecht, On asymptotic properties of Banach spaces under renormings, J. Amer. Math. Soc. 11 (1998), 175-188. [145] B.J. Pettis, A proof that every uniformly convex space is reflexive, Duke Math. J. 5 (1939), 249-253. [ 146] G. Pisier, Un exemple concernant la superr~flexivit~, S6minaire Maurey-Schwartz, Vol. 2 (1974/75). [147] G. Pisier, Martingales with values in uniformly convex spaces, Israel J. Math. 20 (1975), 326-350. [148] G. Pisier, Factorization of Linear Operators and Geometry of Banach Spaces, CBMS Regional Conference Series in Mathematics, Vol. 60 (1986). [149] G. Pisier, Weak Hilbert spaces, Proc. London Math. Soc. 56 (1988), 547-579. [150] D. Preiss, Differentiability of Lipschitzfunctions on Banach spaces, J. Funct. Anal. 91 (1990), 312-345. [ 151 ] D. Preiss and L. Zajf6ek, Fr~chet differentiation of convex functions in Banach spaces with separable dual, Proc. Amer. Math. Soc. 91 (1984), 202-204. [ 152] M. Raja, Mesurabilit~ de Borel et renormages dans les espaces de Banach, Thbse, Universit6 de Bordeaux I (1998).
R e n o r m i n g s o f Banach spaces
835
[153] M. Raja, On locally uniformly rotund norms, Mathematika, to appear. [154] M. Ribe, On uniformly homeomorphic normed spaces, Ark. Mat. 14 (1976), 237-244. [155] M. Ribe, Existence of separable uniformly homeomorphic nonisomorphic Banach spaces, Israel J. Math. 48 (1984), 139-147. [156] H.E Rosenthal, On subspaces of L p, Ann. of Math. 97 (1973), 344-373. [157] H.E Rosenthal, Embeddings of L 1 into L |, Contemp. Math. 26 (1984), 335-349. [158] W. Schachermayer, A. Sersouri and E. Werner, Moduli ofnon-dentability and the Radon-Nikodym property in Banach spaces, Israel J. Math. 65 (1989), 225-257. [ 159] H.H. Schaefer, Banach Lattices and Positive Operators, Grundlehren der Math. Wiss. 215, Springer-Verlag (1974). [160] J.J. Sch~iffer, Geometry of Spheres in Normed Spaces, Lecture Notes Pure Appl. Math. 20, M. Dekker, New York (1976). [ 161 ] V.L. Smulyan, Sur la ddrivabilitd de la norme dans l'espace de Banach, C. R. Acad. Sci. URSS (Doklady), N. S. 27 (1940), 643-648. [162] C. Stegall, The Radon-Nikodym property in conjugate Banach spaces, Trans. Amer. Math. Soc. 206 (1975), 213-223. [163] C. Stegall, The duality between Asplund spaces and spaces with the Radon-Nikodym property, Israel J. Math. 29 (1978), 408-412. [164] C. Stegall, The Radon-Nikodym property in conjugate Banach spaces, II, Trans. Amer. Math. Soc. 264 (1981), 507-519. [165] W. Szlenk, The nonexistence of a separable reflexive Banach space universal for all separable reflexive Banach spaces, Studia Math. 30 (1968), 53-61. [166] M. Talagrand, Comparaison des bordliens d'un espace de Banach pour les topologies faibles et fortes, Indiana Math. J. 27 (1978), 1001-1004. [167] W.K. Tang, On extension of rotund norms, Note aux C. R. Acad. Sci. Paris 323 (1996), 487-490. [168] N. Tomczak-Jaegermann, The moduli of smoothness and convexity and the Rademacher averages of the trace classes Sp (1 <~ p < oc), Studia Math. 50 (1974), 163-182. [169] H. Torunczyk, Smooth partitions of unity on some nonseparable Banach spaces, Studia Math. 46 (1973), 43-51. [170] S. Troyanski, On a property of the norm which is close to local uniform convexity, Math. Ann. 271 (1985), 305-313. [171] B.S. Tsirelson, Not every Banach space contains I p or c0, Funct. Anal. Appl. 8 (1974), 139-141. [172] J. Vanderwerff, Smooth approximations in Banach spaces, Proc. Amer. Math. Soc. 115 (1992), 113-120. [173] J.H.M. Whitfield and V. Zizler, Extremal structure of convex sets in spaces not containing c O, Math. Z. 197 (1988), 219-221. [174] M. Yahdi, The topological complexity of sets of convex differentiable functions, Revista Math. 11 (1998), 79-91. [175] D. Yost, M-ideals, the strong 2-ball property and some renorming theorems, Proc. Amer. Mat. Soc. 81 (1981), 299-303. [176] V. Zizler, Nonseparable Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published).
This Page Intentionally Left Blank
CHAPTER
19
Finite Dimensional Subspaces of
Lp
William B. Johnson* Department of Mathematics, Texas A&M University, College Station, TX 77843, USA E-mail: johnson @math. tamu. edu
Gideon Schechtman t Department of Mathematics, The Weizmann Institute of Science, Rehovot, Israel E-mail: gideon @wisdom, weizmann, ac. il
Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
839
1.1. The good, the bad, the natural, and the complemented . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. The role of change of density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Subspaces o f e pn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Fine embeddings of subspaces of L p into e np . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
839 839 843 843
2.2. Natural embeddings of grk into epn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
852
n 2.3. erk subspaces of m-dimensional subspaces and quotients of gp
853
3. Finite dimensional subspaces of L p with special structure
....................
.........................
3.1. Subspaces with symmetric basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Subspaces with bad gl constant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Restricted invertibility and finite dimensional subspaces of L p of maximal distance to Euclidean spaces 4.1. Restricted invertibility of operators on epn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Subspaces of Lp with maximal distance to Hilbert spaces . . . . . . . . . . . . . . . . . . . . . . . 5. Complemented subspaces
........................................... 5.1. Fine embeddings of complemented subspaces of Lp in s n
......................
5.2. Finite decomposition and uniqueness of complements . . . . . . . . . . . . . . . . . . . . . . . . . References
.....................................................
854 854 858 859 859 862 864 864 865 868
*Supported in part by NSF DMS-9623260 and DMS-9900185, Texas Advanced Research Program 010366-163, and the US-Israel Binational Science Foundation. t Supported in part by the US-Israel Binational Science Foundation. H A N D B O O K OF THE GEOMETRY OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 837
838
W.B. Johnson and G. Schechtman
Abstract We discuss the finite dimensional structure theory of L p" in particular, the theory of restricted invertibility and classification of subspaces of s n
Finite dimensional subspaces of L p
839
1. Introduction 1.1. The good, the bad, the natural, a n d the c o m p l e m e n t e d There are many interesting problems about infinite dimensional subspaces of L p ( : = L p [0, 1]) which have finite dimensional analogues. For example, it has long been a central problem in Banach space theory to classify the complemented subspaces of L p up to isomorphism; the finite dimensional analogue is to find for any given C a description of the finite dimensional spaces which are C-isomorphic to C-complemented subspaces of L p. A lot is known about both the infinite dimensional (see [1]) and finite dimensional (see Section 5) versions of this complemented subspaces of L p problem, but in neither case does a classification seem to be close at hand. It sometimes happens that the finite dimensional version of an infinite dimensional problem leads to a theory which is much more interesting than the infinite dimensional theory. Take, for example, the problem of describing the subspaces of L p which embed isomorphically into a "smaller" L p space; namely, g p; for which there is a fairly simple answer (see [ 1]). Now it is clear that a finite dimensional subspace X of Lp embeds, with isomorphism constant 1 + e , into ~N p if N -- N ( e , X ) is sufficiently large. The attempt to estimate well N in terms of ~ and X (or the dimension of X) has led to a deep theory (see Sections 2.1 and 5.1). A sideline of this investigation also led to deeper understanding of how certain natural subspaces of L p (such as the span of a sequence of independent Gaussian random variables) are situated in L p (see Section 2.2). Besides being an interesting subject in its own right, the study of finite dimensional subspaces of L p is often needed in order to understand properties of infinite dimensional subspaces. For example, the easiest and best way to obtain subspaces of L p, 2 < p < cx~, which fail GL-l.u.st. (cf. [28, Section 9]) is to show that random large dimensional subspaces of g/7p are bad in a certain sense (see Section 3.2). The topic discussed herein which has the most applications is that of restricted invertibility (see Section 4). Basically the theorem says that an n by n matrix which has ones on /7 the diagonal and is of norm M, say, as operator on gp, must be invertible on a coordinate subspace of dimension at least 3 ( M ) n . One of the many consequences of this result is that certain finite dimensional subspaces X of L p contain well-complemented gp subspaces with n proportional to the dimension of X.
1.2. The role o f change o f density Generally the structure of the L p spaces is described for a fixed value of p. However, proofs of many of the results about L p for a fixed p use the entire scale of L~ spaces, 1 ~< p ~< ~ . Consider, for example, the proof [28, Section 4] that a subspace X of L p, 2 < p < cx~, is either isomorphic to a Hilbert space or contains a subspace which is complemented in L p and is isomorphic to gp. There one needs only to compare the L2 norm and the L p norm on X. In proofs of other theorems about L p it is necessary to change the measure before making a comparison between the L p norm and another norm. Since the technique of changing
840
W.B. Johnson and G. Schechtman
the measure ("making a change of density") is used in the proofs of most of the results we discuss in this article, we chose to devote this section to describing the change of densities that arise later. It turns out that the framework in which this technique is most naturally used is that of an L p (/z) space w h e n / z is a probability. For us there is no loss of generality in restricting to that case since the space s N is isometric to L p (/z) w h e n / z is any probability on {1 . . . . . N} for which/z({n}) > 0 for each 1 ~< n ~< N. For such a measure/z we denote L p (/z) by L N p (/z), or just L N p if/z assigns mass 1 / N to each integer n, 1 ~< n ~< N. A density on a probability space ($2,/z) is a strictly positive/z-measurable function g on $2 for which f g d/z = 1. Such a density g induces for fixed 0 < p < cx~ an isometry M - Mg,p from Lp(/z) onto Lp(g d/z) defined by M f -- g - 1 / p f . Sometimes a gain is achieved by making such a change of density; that is, by replacing Lp(/z) by its isometric copy L p (g d/z). The gain usually occurs because for some subspace E of L p (/z) and some value of r different from p, the space Mg,pE is better situated with respect to Lr(g d/z) than E is to Lr(/z). For example, it follows from the Pietsch factorization theorem [28, Section 10] that if an operator T from some space X into L1 (/z) has p-summing adjoint, then by replacing the original L1 space by another isometrically equivalent L1 space, T X is actually contained in L p. Formally, PROPOSITION 1. If T ' X --+ L1 (/Z) (/Z a probability measure) has p-summing adjoint, then there is a change of density g and an operator T" X --+ Lp(g d/z) so that Mg,pT is the compositi2n of T followed by the canonical injection from Lp(g d/z) into L1 (g d/z). Moreover, lITll - 7rp(T*) as long as T X has full support; i.e., there is no subset 12o C 12 with/z(I2o) < 1 so that Tx -- ls2oTx /z-a.e. for every x in X. To prove this factorization theorem, assume for simplicity that/z is a regular Borel measure on a compact space f2 and that T X has full support. Get a Pietsch measure v for the restriction of T* to C ( K ) , which means that v is a regular probability measure on K such that IIT*fll p <~ 7rp(T*)p f Ifl p dv for all f in C ( K ) . This same inequality is true if v is replace by its Radon-Nikod3)m derivative with respect to #, so one can assume that v is of the form g d/z with g >~ 0 and 1 = v S-2 -- f g d/z. It is easily checked that this g is the desired density provided that g is strictly positive/z-a.e., which it must be since T X has full support. (When T X does not have full support, reason the same way but at the end add a small constant function to g and renormalize.) The next change of density result, due to Lewis [43], gives useful information about finite dimensional subspaces of L p. THEOREM 2. Let/Z be a probability measure and let E be a k dimensional subspace of Lp(/Z), 0 < p < ~ , with full support. Then there is a density g so that Mg,pE has a basis
{fl . . . . . fk } which is orthonormal in L2(g d/z) and such that ~_kn=1 If/12 -- k. For a proof when 1 <~ p < c~ see [43]. The first step is to apply Lewis' lemma [28, Section 8] to get an operator T from s onto E for which top(T) -- 1 and I p , ( T -1) = N. The rest of the proof involves checking that the choices g "-- ( ~ ITeil2) 1/2 and f/ "= x//--kg-1/PTei satisfy the conditions of Theorem 2. Another proof, for the entire range 0 < p < cx~, is contained in [56].
Finite dimensional subspaces of L p
841
Recall that if T ' X --+ Y is an operator, y2(T) is the infimum of II T II II U II over all factorizations T -- SU with U" X --+ ~2 and S ' g 2 --+ Y. THEOREM 3. If E is a k-dimensional subspace of L p ( # ) then there is a projection P from Lp(lZ) onto E with Y2(P) ~< k ll/p-1/21. Since the B a n a c h - M a z u r distance from s k to s 2 is k I1/p-1/2l [28, Section 8], Theorem 3 implies that the distance of a k dimensional subspace of Lp(lZ) to s is maximized when the subspace is s k To prove Theorem 3, observe first that the case p = cx~ follows trivially from the fact proved in [28, Section 10] that 7rZ(IE) --- x/~ for every k dimensional space E. When p < cx~, in view of the comments made in [28, Section 10], there is no loss in generality in assuming the # is a probability measure and that E has full support. Theorem 2 says that we can further assume that E has a basis {fl . . . . . fk } which is orthonormal in L2(#) and such that If/" ]2 ~ k. Let P be the orthogonal projection onto E. If p < 2, a simple computation shows that ] ] P ' L p ( # ) ~ Lz(/z)]] ~< k 1/p-1/2 and so also y 2 ( P ' L p ( l Z ) --+ L p ( # ) ) <~k 1~p-I~2. The case p > 2 is even easier. For a generalization of Theorem 3 to spaces of type p > 1 see [60, 27.4, 28.6]. The change of density in Theorem 2 is used in Section 2.1 to show that a k dimensional subspace of L p well embeds into ~pn with n not too large. There what is needed is the relation between the L ~ and the Lp(g d/z) norms on Mg,pE. The relevant estimate follows from a trivial observation which we record for later reference.
Zkn=l
LEMMA 4. Let lZ be a probability measure and let E be a k dimensional subspace of Lp(lZ) which has a basis {fl . . . . . fk} which is orthonormal in Lz(/z) and such that 1]fi] 2 = k . Then f o r e a c h f in E, ]]f]]~ ~
][fl]~ ~<
There is a pretty characterization, due to Maurey [44,63, III.H.10] of subsets of L p ( # ) which are bounded in Lq after a change of density. THEOREM 5. Let # be a probability measure, 0 < p < q < cx~, and S a subset of L p ( # ) of full support. Then there is a density g so that Mg,FS C BLq(gd#) if and only if for all finite subsets So of S and {ax" x E So} C [0, 1], 1/q
(1)
laxX[ q x ESo
Lp(#)
x ESo
Assuming the Pietsch factorization theorem, for the case p - 1 and when S - T Bx for some operator T" X --+ L1 (#), Corollary 6 is little more than a restatement of Theorem 5. Recall [28, Section 5] that when T is an operator into a Banach lattice, M (q) (T) denotes the q-convexity norm of T. COROLLARY 6. Suppose T is an operator from a Banach space X into an L 1 space. Then M(P)(T) -- rCp,(T*).
W.B. Johnson and G. Schechtman
842
A consequence of Theorem 5 that we shall need in Section 4 is the result of [27] that an operator on Lp is bounded on L2 after an appropriate change of density: THEOREM 7. IfO < p < cx~, lZ is a probability measure, and T is an operator on Lp(lZ), 1. L2(gd/z) --+ L2(gd#)ll ~< 2KGllTIIfor some density g ~ 1/2, where then IIMg,pZM~,p KG is the constant in Grothendieck's inequality [28, Section 10]. If one does not wish to make a change to the measure g d#, then by throwing away the part of the measure space where g ~> 2 one gets: COROLLARY 8. If 0 < p < CXZ, # is a probability measure, and T is an operator on Lp(#), then IIRA T R A ' L z ( I z ) --+ Lz(/z)ll ~< 2KGIITII for some set A with lzA ~ 1/2, where RA is the restriction operator defined by RA f "-- 1A f . Theorem 7 is a fixed point version of the following factorization consequence (due to Maurey [44]) of Theorem 5. THEOREM 9. Let lZ be probability measure and T be an operator from a Banach lattice X into Lp(#), 0 < p <~2. Then there is a change of density g and an operator T from X into L2(g d#) so that Mg,pT -- I2,pT, where I2,p is the identity mapping from L2(g d/x) into L p ( g d # ) . Moreover, Ilrll ~< KGM(2~(X)IITII if r x has full support. For the proof of Theorem 9, let {Xi}in=l be in X and assume that T X has full support. Then using a consequence of Grothendieck's inequality (see [28, Section 10]) in the first step we have
[Txi[ 2 i=l
KG II T II
(5
[xi 12
i=1
Lp(lZ)
<~ KGIITIIM(2)(X)
Ilxi II2
i=1 Now apply Theorem 5 with S the image under T of the unit sphere of X to get the inequality in the conclusion of Theorem 9. A trivial perturbation argument now gives Theorem 9 when TX does not have full support. In the general case, given e > 0, the density g can be chosen so that IITII ~< (KGM(2)(X) + e)llTII. We now deduce Theorem 7 in the range 1 < p < c~ from Theorem 9. By duality we can assume that 2 < p < cx~, and also suppose that TLp (#) has full support. If h is a density and we apply Theorem 9 to the adjoint of the operator
Lp(#)
mp,hT Ip,2 > Lp(h d#) ~ L2(h d#)
(2)
we get a density g so that for every f in Lp(#),
f
lTfl2h p-2)/p d# ~ K2 IITII2
f
Ifl2g (p-2)/p d#.
(3)
Finite dimensional subspaces of L p
843
Set go :-- 1 and get densities gl, g2 . . . . so that for each n, (3) is satisfied with h :-- gn and g "-- gn+l " Define ~ "-- Y-~-~=0 oc 2 - n - l g ( p - 2 ) / p . This series is absolutely convergent in L(p-Z)/p(#) to a function whose norm is at most one, so the function g : = IlgllLj(~)g is a density with g ~> 1/2 which satisfies, for arbitrary f in Lp(#),
J'lvii 2g(p-2)/p
d/z ~< 2 K 2 IITII2
f
Ifl2 g fp-2)/p dlz.
This gives Theorem 7 with a slightly better constant when T L p ( # ) has full support and hence also T h e o r e m 7 as stated in the range 1 < p < oc. For 0 < p ~< 1, T h e o r e m 7 follows via interpolation from T h e o r e m 10. THEOREM 10. If 0 < p <. 1, a > 1, # is a probability measure, and T is an operator on L p ( # ) , then IlMg,pTMg-,lp " L ~ ( g d # ) -+ L ~ ( g d#)ll ~< alITll for some density g. To prove T h e o r e m 10, it is enough to find a strictly positive function ~l/p in L p ( # ) so that T maps the order interval [_~l/p, ~l/p] into [-al]Tll~, 1/p, allT]l~,l/p]. The main point is that every operator T on L p (#), 0 < p ~< 1, has a modulus ITI which satisfies IllTlll = IITII and ITfl <. I T l l f l for all f in L p ( # ) (see, e.g., the remark preceding T h e o r e m 3.2 (DO --/7 in [36]). One then defines ~l/p _ 1 + Y-~.n=l a [ITII-/TITI/71. See [27] for details when p--1. We conclude this section with a change of density l e m m a due to Pisier [51] which, except for constants, improves T h e o r e m 5. T h e o r e m 11 will be used in Section 2.1. THEOREM 1 1. Let # be a probability measure, 0 < p < q < ~ , and S a subset of L p ( # ) .
The following statements are equivalent. (i) There is a constant C1 and a density g so that for all measurable sets E and x in S, II1ExlILp(~) <<.e l ( r E gdlz) 1/p-1/q. (ii) There is a constant 62 and a density g so that g g , p S C C2BLq.~(gdlz). (iii) There is a constant C3 so that for all finite subsets So of S and subsets {ax: x E So } of[O, 1], II SUPx~so laxxlllL~,(~) <<.C3(~x~So [a~lq) 1/q. Moreover, there is a constant C = C(p, q) so that in the implication (i) =:~ (j), Cj CCi. For a proof of T h e o r e m 11, see [51 ]. This paper also contains a nice proof of T h e o r e m 5.
2. Subspaces of gpH 2.1. Fine embeddings of subspaces of L p into s Let X be a k-dimensional subspace of L p, 0 < p < oo, and let e > 0. What is the smallest n such that X (1 + s ) - e m b e d s in s n . That is, what is the smallest n such that there is a k-dimensional subspace Y of ~ Hp and an isomorphism T" X ----> Y with IIT IIIIT-1 [[ ~< 1 + e ?
844
W.B. Johnson and G. Schechtman
Let us denote this n by Np(X, e) and the maximal Np(X, e), when X ranges over all k-dimensional subspaces of Lp, by Np (k, e). Fixing a basis in X and approximating each of its members by an appropriate simple function, one sees that Np (X, e) < c~ for every k-dimensional X and e. Moreover, it depends on X only through its dimension k so that Np(k, e) < c~. However, one gets that way a (larger than) exponential in k bound on Np(k, e). In this section we shall review results which give much better bounds, close to the best possible ones. The case p = 2 is of course trivial and one can take n -- k even for e = 0. The case p = 1 has a nice geometrical interpretation: The unit ball of the dual to a k-dimensional subspace of s is easily seen to be (isometric to) the Minkowski sum of n segments in R k (centered at 0) and visa versa. Consequently, the n sought after is the smallest n such that every (centered at zero) body K in R ~ which is the Minkowski sum of arbitrarily many segments (or the limit of such bodies - these are called zonoids) can be e-approximated by a body Z which is the sum of n such segments in the sense that Z C K C (1 + e)Z. The history of this problem can be traced to the offsprings of Dvoretzky's theorem as discussed in [24]. There the case of X being a k-dimensional Hilbert space (which embeds isometrically in all the L p spaces) is treated and solved quite satisfactory: For some absolute constant C, Np(~ k, e) is at most c e - z k for p < 2 and Ce-2pk p/2 for p > 2. This is best possible except that it is unknown whether the factor e -2 can be replaced by a smaller function of e. Notice the following nonintuitive special case: The k-dimensional Euclidean ball can be approximated by a body which the sum of a constant (depending on the degree of approximation) times k segments in R ~. The first result in this direction which did not involve Euclidean spaces was proved k (which is known to in [31]. There it was shown that, for 0 < p ~< 1 and p < q < 2, ~q embed isometrically into L p whenever 0 < p < q ~< 2) (1 + e)-embeds into s n f o r s o m e n <, C(p, q, e)k. Later the second named author found some initial results indicating in particular that the dependence of n on k in the general problem stated at the beginning of this section is polynomial rather than exponential as one is first tempted to believe. The first proofs were quite complicated and worked only for p < 2 as they used fine properties of qstable random variables. Later a much simpler method was introduced in [54]. Assuming, as we may, that X is already a subspace of s U for some finite N, pick randomly a "few" coordinates and hope that the natural projection onto these coordinates, restricted to X, is a good isomorphism. If we do it with no additional preparation this cannot work. Indeed, X may contain a vector with small support (say one of the unit vector basis elements of X), in which case the chance that a coordinate in its support is picked is small; of course, if no such coordinate is picked, the said projection cannot be an isomorphism on X. The point is that one wants to change X first to another isometric copy of X in which each element of X is "spread out". This can be done by a change of density. The method of [54] was used with other tools in [8,58], and some other papers to produce the best known results. In these results it is not known what is the right dependence of n on E and we shall not try to emphasize what are the exact estimates one gets from the proofs. However, the dependence of n on k is best possible except for log factors in some places; we shall pay more attention to this in the sequel. We now state the best known results.
Finite dimensional subspaces of L p
845
THEOREM 12. (i) For p > 2, Np(k, e) <~C(p, e)k p/2 logk. (ii) For 1 < p < 2, Np(k, ~) <~ C(e)klogk(loglogk) 2. (iii) For p = 1, N1 (k, e) <~C(e)klogk. (iv) ForO < p < 1, Np(k,e) <~C ( p , e ) k l o g k ( l o g l o g k ) 2. Under some conditions ensuring that X does not contain good copies of e mp spaces, one gets better results for p < 2. Recall that a quasi-normed space X is of type p with constant C for some 0 < p ~< 2 provided
21J211
1/p
E
~-~~iXi i=1
C ~
Ilxill p
i--1
for all finite sequences x l . . . . . x~ of elements of X. The best C is denoted Tp (X). The space L p, 0 < p ~< 2, is of type p. Recall also that K (X) denotes the K-convexity constant of X, i.e., the norm of the Rademacher projection in L2(X). See [28] for a brief discussion of these notions (although it is restricted to the normed spaces, which always have type p ~> 1) and [45] for a more comprehensive discussion. THEOREM 13. (i) Let 0 < p < q < 2 and let 0 < e, C < oo. Then for some constant C ~ =
C~(p, q, e, C ) a n d a l l k-dimensional subspace X of Lp with Tq(X) <~C, Np(X, e) <~C'k. For p -- 1 we have a quantitatively better estimate: (ii) For all k-dimensional subspaces X of L1, N1 (X, e) <<,C(e)K(X)Zk. Theorem 13(i) was proved for p ~> 1 in [8]. [32] contains the full statement with a different proof. [8] also contains Theorem 12(i) and somewhat weaker versions of Theorem 12(ii), (iii), and Theorem 13(ii). The exact Theorem 12(ii) is contained in [59] while Theorem 13(ii) is the main result of [58]. Theorem 12(iii) follows from it since it is known ([49]) that K ( X ) <~ Cl~/i-~k for every k-dimensional subspace X of Ll" see L e m m a 17. Finally, Theorem 12(iv) was proved only recently [56,64] after noticing its omission while writing this survey. Before describing the proofs, we mention that there are several unsettled problem related to Theorems 12 and 13. The most important one (or at least the one that attracted the most attention) is whether the various log factors and the dependence on the type and Kconvexity are really needed. It is strange that the constants in the proofs blow up when X contains gpm spaces. Actually, as we shall see below, in at least some of the proofs the k Another problem is the determination of worst case occurs when X is isometrically gp. the dependence of N (.) on e. Scant attention has been given to that in the published work. A problem we find particularly interesting is whether there is an "isomorphic" (as oppose to "almost isometric") version to some of the results here. Here is an instance of this problem: is it true that for all 1 < p < 2 and all )~ > 1 there is a positive constant C = C(p, )~) such that whenever n < )~k, ~kp C-embeds into e nI 9. Some progress on this problem has recently been achieved in [48].
W.B. Johnson and G. Schechtman
846
Next we would like to sketch some of the ideas involved in the proofs of some of the statements of Theorems 12 and 13. As we already indicated above, a common feature of all the proofs we shall sketch is that, using a change of density, we first find an isometric copy of X with some additional good properties. We delay stating theses properties and the actual change of density that ensure them until later and denote the new space by the same notation X. We may also assume without loss of generality that X lies in a finite (but large) dimensional s m space, say L mp (/z), where # is a probability measure on {1 . . . . . m}. We denote #i = # ( { i }). We would like to show that the restriction operator to a set of relatively few of the m coordinates is a good isomorphism on X. We prefer to do it iteratively by first showing how to find a subset of cardinality at most m/2 such that the restriction operator is a (very) good isomorphism on X, provided m is much larger than k. We shall then show how to iterate this procedure. The choice of the subset will be random; for example, it is enough to show that, for an appropriate e(k, m), m
AveA
sup 2 ~ I.Li IXi [P -- ~ #i [Xi Ip x~X, Ilxll~ 1 icA i=1 m
= E
sup Y~ei#ilxil p <~e(k, m). x~X, Ilxll~<1 i = 1
(4)
Here AveA denotes the average over all subsets of {1 . . . . . m } while E is the expectation with respect to the natural product measure on { - 1, 1 }n. Indeed, if this is the case then for l+e(k,m) 1/p_ some A the restriction operators to both A and the complement of A are (1-e(k,m~) isomorphisms and of course either A or its complement A c is of cardinality at most m / 2 . Alternatively, in order to find such an A, it is enough to show that m
P(
sup ~ ei lzi lxi lp > e(k, m)) < 1. xEX, Ilxll~1 i = 1
(5)
In both cases after iterating we get that as long as
l (l§ 9 1 - e(k, m2 -i)
~< 1 + e,
(6)
X must (1 + e)-embed into s n for some n ~< m2 - l - 1 So this approach reduces the problem to finding good bounds on the quantity in (4) or (5). For technical reasons involving splitting of atoms, as explained in L e m m a 14 below, we may need to enlarge m to at most 3 m / 2 before making the random choice in each step. p3m/4 This does not effect the process significantly; after the random choice we end up in ,.p and this just means that instead of (6) we shall need to ensure that
l (l+e(k'm(3)i)) 9
1 - e ( k , m ( 3 ) i)
~< 1 + e.
(7)
Finite dimensional subspaces of L p
847
Before proceeding further, we would like to point out why there is very little hope of cn log n eliminating the log factors altogether using this approach. Consider, in g l , n vectors of the form xi -- (c logn) -1 ff-~jE~ri ej with 00i, i -- 1 . . . . . c logn, disjoint sets of cardinality c log n (assuming it is an integer). It is easy to calculate (and appears in many probability books, sometimes as the coupon collector's problem) that, if c is small, a random choice of ~nClog n of the coordinates will most likely miss at least one of the 00i's. Thus the restriction c to a "random" subset of cardinality ~n log n will not be an isomorphism on X - span{x/} (which incidentally is isometric to g~). We did not change the density here, but it is not hard to see that a change of density is not going to help to reduce the minimal m for which this procedure works below cn log n for some absolute c > 0. We now continue sketching the idea of the proofs. We first sketch a version of the simple argument of [54] which only gives
Np(k, s) <~ C(s)k 2
for0
Np(k, ~) <~ C(a)k (p+2)/2
for p > 2.
and (8)
By Lemma 4 there is a change of density on L mp so that for all x e X, Ilxll~ ~ kl/Pllxllp
for 0 < p < 2
IIx II~ ~ k l/21lx lit
for p > 2.
and (9)
Splitting the atoms of this change of density we may assume in addition that #i ~ 4 / m (paying by enlarging the original m to 3m/2). This follows from the following simple lemma. LEMMA 14. Let lZ be a probability measure on { 1, 2 . . . . . m}. Then there is an m <, M <, 3m/2, a probability measure v on { 1, 2 . . . . . M} and a partition {001,o~ . . . . . 00m } of { 1, 2 . . . . . M} satisfying (1) v({i}) ~< 4/m, i = 1 . . . . . M, and (2) Y~ie~j v({i}) -- #({j}). The proof of the lemma is very simple. Split the atoms of # which have mass larger than 4 / m into pieces each of size larger than 2 / m . This does not add more than m / 2 atoms to the original ones, thus ending the proof. Of course L p (#), and thus also X, naturally embeds into Lp(v) in particular the estimates in (9) still hold for the image of X in Lp(v). This justifies the statement in the paragraph preceding the statement of the lemma. Fixing an x e X of norm one, we get by classical deviation inequalities (see for example Proposition 5(iii) in [55]) that for 0 < p < 2,
eilzi Ixil p > t
P i--1
<~ K exp - 6 t 2 /
lZi [xil p i----1
max lZilXi, p) ~< K e x p ( -(~t2/l~i<~m
<~ K e x p ( - 6 ' t 2m / k).
848
W.B. Johnson and G. Schechtman
Let 0 < t < 1 and pick a t-net A/" in the unit sphere of X which has at most (3/t) k elements (see, e.g., [47, p. 7] for an easy proof of the existence of such a net). Then as long as k 2 < ct 2 log- 1(1 / t)m (for some absolute constant c > 0), m
P ( sup ~_~ei#ilxi[ p > t ) ~< K e x p ( - 6 ' t 2 m / k ) , i=1
from which it is easy to get that, as long as k <<,cm 1/2, m
P(
sup
~x~X, Ilxll~1
ff-~ ei#ilxil p > C k / m 1/2) < 1.
(10)
i=1
This and (7) implies the desired result (8) for 0 < p < 2. The treatment for p > 2 is similar. [3 This sketch of the argument of [54] was given only to illustrate the basic method involved. We now continue to sketch the arguments which give the stronger statements of Theorems 12 and 13. The point is to find estimates for the quantities in (4) or (5) which are better than (10). We first relate the second quantity in (4) to a similar one involving Gaussian variables. Let gl . . . . . gm be independent standard Gaussian variables. Then
/-y
m
E
m
E sup sup ~gi#ilxi[ p ~_~ei#ilxil p ~< ~/-gVL xeS, Ilxll~
This version of the "contraction principle" is easy to prove: Replace each of the gi on the right by ei Igi l, where {8i} is independent of {gi }, and replace E with the successive application of the two expectations E, Eg. Now push the expectation ~']g inside the outer 1. ] and use the fact that Elgll = ~/-2/7r. The problem now reduces to evaluating the quantity /77
E
sup ~-"~gi#ilxil p x~X, Ilxll~<1 i=1
(11)
Set Gx - ~ ~ i % 1 gi#ilxil p" Then {Gx}x~X, Ilxll~l is a Gaussian process indexed by elements of B(X) and we are required to estimate the expectation of its supremum. This is a well studied area in probability theory (see, e.g., [41 ]), related to the continuity of Gaussian processes. This quantity (or the similar one involving the Rademacher functions in (4)) is evaluated by different means in the proofs of the different parts of Theorems 12 and 13. We shall first present a PROOF OF THEOREM 13(ii) (Sketch). Consider the process {nx}x~S, llxll~l where Hx = ~--~im=lgi #i Xi. Then
-EH
EC, Cy
Finite dimensional subspaces of L p
849
for all x, y e X of norm at most 1. Slepian's lemma (see, e.g., [41, p. 75]) implies now that m
m
E
~< E
sup ~~gi#ilxi[ x~X, Ilxll~1 i=1
sup Zgi#iXi xcX, Ilxll~1 i=1
We shall show that after a change of density (and possibly enlarging m to 3m/2), m
E
CK(X)
sup Zgi#ixi xeX, Ilxll<~1 i--1
k) -m
/2 (12)
.
Then one concludes the proof of Theorem 13(ii) by applying (7). To prove (12) we shall use the following two propositions. PROPOSITION 15. Let X be a k-dimensional subspace of L"~ (lZ) with lZ a probability measure and let f l . . . . . fk be an orthonormal basis of X (considered as a subspace of Lr~(lz)). Then, one can split some of the atoms of lZ to a total of at most M - 3m/2 atoms getting a new probability measure v and a natural embedding I" Lr~ (lZ) --+ L M (v) satisfying M
E
sup ZgiviYi ycY, IlYlI~<1 i=1
~< 2m-1/2E
gifi
9
i=1
Here Y - I X and II" II* is the dual norm to that of X where duality is given by (~-~ ai fi, ~ bi fi) -- Y~ ai bi. PROPOSITION 16. With the notation of the previous proposition,
1/2 ~gifi i--1
~< x/'2K (X) OO
We now conclude the proof of Theorem 13(ii) by applying Lewis' change of density, Theorem 2 to get an orthonormal basis for (an isometric copy of X) satisfying Vq II( ~ i -k1 fi2)l / 2 l ] ~ < n 1/2 . Before sketching the proofs of Propositions 15 and 16 we would like to deduce Theorem 12(iii). This follows from the following lemma (first observed in [49]). LEMMA 17. Let X be a k-dimensional subspace of L1. Then K ( X ) ~ Cx/logk for some absolute C. PROOF 9 Using Lewis' change of density we may assume Ilxll~ ~ kllxlll for all x e X 9 Then easily Ilxlll ~< Ilxllp ~< k(P-1)/Pllxlll for all x e X. Letting Xp denote X with the Lp norm we get that
W.B. Johnson and G. Schechtman
850
K ( X ) <~k(p-1)/PK(Xp) <~k(p-1)/PK(Lp) <~ Ck (p-1)/p V / p / ( p -
1),
where the last inequality follows from the easy fact that K ( L p ) = K ( L p / ( p _ l ) ) and a (not entirely obvious) application of Khinchine's inequality, with the best order of the constant, in L p / ( p - l ) . Picking p with p / ( p - 1) = logk we get the result. N We now turn to the PROOF OF PROPOSITION 15 (Sketch). Using the contraction principle in the first inequality, we get m
E
sup Egi#ixi x~X, Ilxll~ 1 i=1 m
~< max /z/'/2E ~ sup l <~i<~m x~X, Ilxll~<1 =
=
1/ ~ , max #i .2]~
l<~i<~m
1
max # i / 2 E
l<~i<~m
E i /J2xi i=1
sup gi#il/2ei,x xcX, Ilxll~<1 i--1 sup [[Y~aj fj [[~<1 j = l
i=1
Put hj = (~im__1 gi# i |/2ei, fj }. Then hi . . . . . hk are standard Gaussian variables which are easily seen to be independent (check that E h j h l - •jl). Thus
E
sup
= E
[[~-~aj f jl[<~l j=l
sup
~_~ aj gj
IIEajfjll~ 1 j=l
i--1
= E
gifi
9
(13)
i=1 It remains to see that splitting the atoms we may also assume that ~i ~ 4 / m , but this follows from the splitting of atoms Lemma 14. E] PROOF OF PROPOSITION 16. The proposition follows easily from [ 17, Lemma 1] but we n,k present a different proof. Let {Sij}i_l,j__ 1 denote nk independent Rademacher functions (i.e., the 8ij are the coordinate functions in the product probability space { - 1 , 1 }nk). In the first inequality below we use the central limit theorem and in the last Khintchine's inequality (with the best constant) [28].
/ igifi i=1
k
n
[ lim E j--1 i=1
, 2 ) 1/2
Finite dimensional subspaces of L p
gijfj, f
= lim r/-1/2 sup
851
; IlfllL2(X) <<.1
n-----~ o o
j----1 i----1
(eijfj,xij);xij EX, E
~< K ( X ) lim n - 1 / Z s u p n -----~(N2)
j=l i=l
~< K (X) lim sup
n-l#
x2 9
n---> oo
9 .
Zi,j EijXij
1, j
L1
L~
j----1 i=1
2 E Z P~ijXij i,j = K (X)
__
;E
sup
j=l
Ll
L~
Z
gij Xij
i,j
<~ x / 2 K ( X ) L~
We now sketch a proof of T h e o r e m 13(i). We have chosen a version of the proof of [32] since it is the shortest one. However, this proof does not give a good dependence on e. m The conPROOF OF THEOREM 13(1) (Sketch). We assume first as we may that X C gp. dition of T h e o r e m 1 l(iii) is easily seen to be satisfied for So = (any finite subset of) the unit ball of X and C3 = Tq (X). So we can deduce from that theorem that, without loss of generality, X C L m p (#) for some probability measure # and Ilxllq,~ ~< for all x e X and some C' which depends only on p, q and Tq (X). By T h e o r e m 1 l(i) it is also
C'llxllp
easy to see that we may assume that #i ) 1/2m (replace g with g+---!l). Splitting the large atoms o f / z as in L e m m a 14 and changing m to 3m/2 we may assume in addition that #i ~ 4/m for all 1 <~ i <<.3m/2. Put r = q / p and note that for x r X with Ilx IIp ~< 1,
3m/2 = max jl/r I{lZilxilP}i-1 Ir,~x~ l<~j<~3m/2
(~jlxjlp),
= m a x t ( # { i ; #ilxil p >~ t})1/r
(14)
t>0
with {(#jlxjlP) * } denoting the decreasing rearrangement of {(#jlxjlP)}. Using the fact that/z i is of order 1 / m and relating the quantity in (14) to
IlXllq,~ -
maxt(#({i; t>O
[xil >~t})) '/q
maxt(m-l#{i; t>0
Ixil ~ t})l/q
(15)
W.B.Johnsonand G. Schechtman
852
we get that the quantity in (14) is at most Tq(X). We now use the inequality
P-1 Cm -4
for some C depending only on p, q and
• 8iai > t) ~ 2exp(-6(t/ll{ai}llr,~) s)
(16)
i=1
which holds for all t > 0, 1 < r < 2 and all sequences of scalars {ai }. Here s -- r and is a positive constant depending only on r. The inequality is a special case of a martingale inequality of Pisier (see [47, p. 45] for a proof or [55], Proposition 5 for a discussion of this and other similar inequalities). Note that we may assume that r = q / p < 2. Using (16) we get from (14) that for all x E X with IIx IIp = 1,
3m/2 Z 8i#i Ixi Ip > t) <. 2exp(-6'tSm) i=1
for some 6 ~ depending only on p, q and ment leading to (10),
P(
sup xcX, Ilxll~ 1
Tq(X).
From this we get, as in the standard argu-
3m/2 Z 8i#ilxilP > t) <. 2exp(klog(3/t)- 6"tSm).
(17)
i=1
It follows that, as long as 6'tSmlog-l(3/t) > 2k, We can find a set of at most 3m/4 coordinates for which the restriction operator is an (1 § 2t)-isomorphism. Choosing m 1/s we get that there is a set of at most 3m/4 coordinates for which the t ~ ( ~ log-k-) restriction operator is an (1 + C ( ~ log ~)l/s)l/p-isomorphism where C depends only on p, q and Tq(X). Iterating, we get that as long as
-I i=1
1 § C m(3/4) i log
m(3/4)i)l/s) 1 / p k
~
X must (1 + e)-embed into g np for some n ~ m(3/4) l+l Note that in each step of the iteration we get a space which is at most 2-isomorphic to X. Since the new space has type q constant at most twice that of the original space we can continue the iteration. Now it is easy to get the conclusion. []
2.2.
Natural embeddings of g'~rinto g.p
The methods embeddings" explicit good be to produce
/1
described in Section 2.1 as well as all other methods for producing "tight are probabilistic and as such are not constructive and do not produce an embedding. The most basic question concerning explicit embeddings may a specific good embedding of g~ into gin with m proportional to n (m ~< 2n,
Finite dimensional subspaces of L p
853
say). We remark in passing that for p an even integer there are specific embeddings (even isometric ones) of ~ into g mp with the relation between m and n close to the optimal one (in particular, for p = 4, m ~ n2). See [38] for that. A natural approach to get an explicit embedding of ~ into ~'~ is to fix a natural subspace Xn of L1 which is well isomorphic to ~ , for example the span of n independent standard Gaussian variables or n independent Rademacher functions, and find a subspace Ym with Xn C Ym C L1, Ym well isomorphic to ~ , and m small. However, this fails in a very strong sense: under the requirements above, the smallest m can be is C n for some C > 1 depending only on the distance of Ym to gin. Here is a somewhat stronger theorem from [21]. Recall that for an operator T : X --+ Y, Y l ( T ) = infllullllvll, where the inf is taken over all L1 (v) spaces and all operators v : X ~ L 1(v), u : L1 (v) --+ Y, satisfying T=uv. THEOREM 18. For every 0 < K < cx~ there exists a ~ = 6 ( K ) > 0 such that if X is the span, in L 1, o f n independent Gaussian variables or n independent Rademacherfunctions, X C Y C L l, and the inclusion J" X --+ Y satisfies Y1 (J) <~ K then, f o r some m ~ e ~n, ~'~ is 2-isomorphic to a subspace o f Y. In particular dim Y ~> e an. The proof of this is rather technical and we shall not reproduce it here. [21,22] contain many refinements and variations of this theorem. Also, [60, p. 201 ] contains an exposition of the proof of the simplest instance of this class of results; namely, the statement in the "in particular" part of Theorem 18 for X being the span of n independent Rademacher functions.
2.3. s k subspaces o f m-dimensional subspaces and quotients Of ~p This short section deals with the question of what is the largest k such that ~kp well embeds into any m-dimensional subspace X of s as well as some related questions. We shall not present any proofs but only summarize what is known on this subject. Note that, since X -- ~ well embeds into ~ /7p, p < 2, for m proportional to n, the answer to the question above for p < 2 is not very interesting (i.e., k must be bounded) unless n - m -- o(n). We shall say something about this case latter. For 2 < p < c~ the following theorem of Bourgain and Tzafriri [12] basically solves the problem. Let k = k p ( X , K ) be the maximal dimension of a subspace Y of X which is K-isomorphic to s k THEOREM 19. Let 2 < p < cx~ and E > O. Then there are positive constants c-= c(p, ~), C - C ( p , e) such that f o r all m < n and every m-dimensional subspace X o f s
kp(X, 1 + e) ~ c m i n { m p* /2, (m /n2/P) p/(p-2) }. The result is best possible in the sense that f o r each m < n there exists a subspace X with k p ( X , 1 + ~) <<,C min{m p* /2, (m/nZ/p)p/(p-2)}. (p, - p / ( p - 1).) As can be suspected from the statement, the proof of this result is quite involved and very technical. It uses ideas from the work of Bourgain and Tzafriri concerning restricted
W.B. Johnson and G. Schechtman
854
invertibility (some of which is surveyed Section 4.1 below) as well as from Bourgain's work on A p sets [6]. We think it worthwhile to find a simpler proof. As we said above, there can be no similar theorem in the range p < 2. However, one can prove a similar theorem for quotients of g np, 1 ~< p < 2. This was done by Bourgain, Kalton and Tzafriri in [7]. THEOREM 20. For each 1 <<.p < 2 there is a constant Cp > 0 such that if X is an m-dimensional quotient space of g np then
kp(X, Cp) >~ Cp 1(mP/n 2(p-l)) 1/(2-p), for p > 1, while kl(X, C1) ~ C-llm(/1 + logn/m). Except for the constants involved the results are best possible. Note that for m proportional to n the resulting dimension of the contained g p space is also proportional to n. For p - 1 this case was observed earlier in [15]. Note also that the conclusion of this theorem is "isomorphic" rather than "almost isometric". We do not know if one can replace the constant C p in the left hand side of the inequalities by 1 + e (of course paying by replacing the constant in the right hand side by one depending on 8). There is also a version of Theorem 19 for p -- cx~" If m ~> n ~, with 3 > 0, then every mdimensional subspace X of ~ contains a well isomorphic copy of ~ with k >~c(S)m 1/2. This was proved in [20] for m proportional to n, and in [5] in general. When m is very large there is also a version of Theorem 19 for p = 1. It was proved in [25] that for every m-dimensional subspace X of ~ , k, (X, K) ~> c min{ (n/(n - m)) log(n/(n - m)), n }. K and c are universal constants.
3. Finite dimensional subspaces of
Lp
with special structure
3.1. Subspaces with symmetric basis In this section we treat the classification of the finite symmetric basic sequences in L p, 1 ~< p < c~ and to some extent also the classification of the finite unconditional basic sequences in Lp, 1 ~< p < 2. Recall that a sequence x l . . . . , xn in a quasi normed space X (over R) is said to be K-symmetric if for all scalars {ai }, all sequences of signs {8i } and all permutations Jr of {1 . . . . . n}
• i=l
ai xi
~
• i=1
siazr(i)xi
Finite dimensional subspaces of L p
855
If we require the inequality only for the identity permutation the sequence is called Kunconditional. The article [61 ] treats the classification of symmetric basic sequences in L p, p > 2 so we only state the result (from [29]; see [61, Theorem 4.4]). THEOREM 21. For every 2 < p < ec and every constant K there is a constant D such that any normalized K-symmetric basic sequence in L p is D equivalent to the unit vector basis o f N n with the norm
]]{ai} 11--max[ ( Z
]ailP) ' / p , w ( Z
]ai 12) ./2 }
(18)
f o r some w ~ (0, 1).
Of course, since g~ isometrically embeds in L p, any norm of the form (18) embeds, with constant 2, into L p. For 1 ~< p < 2 the structure of the symmetric sequences in L p is more involved. Let M be a Orlicz function (see [28, Section 5]) and g i the associated Orlicz sequence space. It turns out that the space ~M embeds isomorphically into L p if and only if the unit vector basis of g i is p-convex and 2-concave and this happens if and only if M(]t] I/p) is equivalent to a convex function and M ( t |/2) is equivalent to a concave function on [0, ec) (see [14]). Recall that two functions M1, M2 :R --~ [0, co) are equivalent (at 0) if there exist constants K l, K2, )~, # and xo > 0 such that for all Ix l < xo KIM2()~x) <~ M | ( x ) ~< K2M2(lZX).
With the right quantifiers, a similar statement holds also for finite dimensional Orlicz spaces, g~4" The embedding, when it exists, is as a span of independent, identically distributed symmetric random variables. It follows that if {Mj }jm__l is a collection of Orlicz functions such that ~ K-embed in Lp for all j and ~j > 0 for all j then also IR" with (
m
n
the norm I]" l] -- }-~'~j=l)~J ]]" IleM~)
1/p
K-embeds in L p. The converse is also true.
THEOREM 22. For every constant K and f o r every 0 < p < 2 there is that given any normalized K-symmetric basic sequence {fi }in=1 in Lp there are m symmetric functions M j :IR --+ [0, cxz), j = 1 . . . . . m, f o r Mj(]t] I/p) are convex and M j ( t 1/2) are concave on [0, ec) and f o r {3~ }in=l is D equivalent to the unit vector basis o f N n with the norm
m )l/p I1.11- ~)~jll.ll p j=l ~Mj
a constant D such there is an m and which M j (0) = O, some weights )~j,
(19)
This theorem is a consequence of the following very nice inequality of Kwapiefi and Schtitt [39].
W.B. Johnson and G. Schechtman
856
THEOREM 23. Let {ai,j}in,j=l E R n2 and denote by of {ai,j}. Then, for any 1 <~p < oc, (~-~
) 1/p lai,rc(i)]p
Avert
1~
<~ -n i=1
i=1
{a*}n21 the decreasing
rearrangement
( , n ~1 ) ,p lip a i 4- -ai n i=n+l (~ )l/p lai,:r(i)lp
~< 5Avejr
.
(20)
i=1
Here Avert denotes the average over all permutations of 1 . . . . . n. The case p = 1 of Theorem 22 appears in [39]. The proof for the other values of p is quite similar and we shall sketch it below. The starting point is a lemma which gives another equivalent expression to the ones in (20). For 1 ~< p < ~ put
M ( t ) - Mp(t)- { (p
-
if [tl ~ I/n,
~nl)-lnp-l[t[P
-
(p--~nl)-l(p[t[--~n
1)
(21)
if Itl > 1/n.
Note that Mp is an Orlicz function and that Mp(t 1/p) is a concave function on [0, oc). LEMMA 24. Let 1 <, p < oc and let a l >/a2 >/... >/an2
>/O. Then
1/p 1
}n 2
II a/ /-lll
<~ -
ai 4- -
n i=1
}/72 ~<~P ]]{a//=, ]]e~.
ai
PROOF. Note that M ( p t ) >~p M ( t ) for all t. Consequently, if
(
p
It follows t h a t
~ai
ai
and
4- n p-1
n
n
~--~mi>l/nai ~ ~-~mi>l/n(pai<2
n 2 IIgM ~< 1 then, II{ai}i=1
1)z
pai
n
(22)
n i=n+l
ap
ai~l/n @)
<
) 1 <. --. p
(23)
1 so that
#{i; ai > 1/n} <. n.
i=1
n2 n It now follows from (23) that Y~i=l ai + n(n -1 ~,i=n+l aP) 1/p ~ 3 n
n2
To prove the other side inequality assume Y~i=l ai 4- n(n -1 ~-~4=n+l aP) 1/p ~ 1. Then
(n2
riP-1 ~ alp <~np-1 ai ~ 1/n 1
n
and, since an+l ~< n ~ i = l
Z
aip +
i=n+ 1 1
Z
aip
)
ai <~1/n,i ~n n
ai <~ -if, ~ai>l/nai <~ Z i = I ai <~ 1 .
<~2
Finite dimensional subspaces of L p
857
It follows that (for n ~> 2) n2
ZM(ai)_
(p
i=1
~<
p-1)-l(
n
(
1),
p
Z
(
ai>l/n
pai
H
+ n p-1
Eai ai~l/n
p
)
(p + 2) ~< 4.
T/
Since M(4t) ~> 4M(t) this concludes the proof.
D
We now turn to a m PROOF OF THEOREM 22 (Sketch). Let {f/}in__l be a 1-symmetric basic sequence in lp. Then up to a universal constant,
(m
~-~ ai fi
E
i=1
k=l
=
Z k=l
(azr(i)fi(k)) 2
Avejr i=1
(a~r(i)fi(k)/1.k) 2
1.~Averr i=1
where )~kp -- II{f/p (k)}i=l n IleM~/ 9 By Theorem 23 and Lemma 24 the last expression is equivalent, with constants depending on p only, to
m
1 Z )~p n k I{aflf~(~)l p/)p , :<}/,j-, ~,~/,,
)l/p (24)
k--1
Put, for k = 1 . . . . . m,
Nk(t)
-- ~ M2/p(tl3~(k)lP/)~f). i=1
It is easy to check that the Ark are Orlicz functions w i t h Nk(t p/2) concave and that 17 [l{a;]f j(k)] p/)~p n ] --]]{aP}i=l]f / k }i,j=l eM2/p
Nk
--
]{ai} n/ = 1 ]•mk' P
Where Mk(t) = Nk(ltlP). From this and (24) it is easy to conclude the proof. The main result of [46] states that every unconditional basic sequence in L p, 1 <~p <~ 2, is equivalent to a block basis of a symmetric basic sequence in L p. (The block basis can be chosen to be with equal coefficients and consequently the embedded space is also complemented. This will not be used here.) Of course if the unconditional basic sequence
W.B. Johnson and G. Schechtman
858
is finite also the containing symmetric sequence can be taken finite and the constants (of the symmetricity and of the equivalence) can be controlled by the unconditional constant. "" F/ Recall that given a sequence of n Orlicz functions M -- {Mi }i=1 the modular space ~ is R n with the norm ]]xl]~ - inf{t > 0; ~--~in_l Mi(xi/t) ~ 1}. Using [46] and Theorem 22 one can now easily prove the following theorem. THEOREM 25. For every constant K and f o r every 1 <<.p < 2 there is a constant D such that given any normalized K-unconditional basic sequence {fi }in1 in Lp there is an m, m Orliczfunction sequences M j -- { M j i } i n l , j -- 1 . . . . . m, and positive constants ~j and # j i such that Mji([t[ 1/p) are convex and M j i ( t 1/2) are concave and {fi }in__l is D equivalent to the unit vector basis o f R n with the norm
Ilxll
-
m )l/p ~_~.jll{xi.ji}i~=lll p~j
(25)
j=l Of course if m j i ( t 1/p) are K-equivalent to convex functions and m j i ( t 1/2) are Kequivalent to concave functions then any norm as in (25) embeds into L p with constant depending on K only.
3.2. Subspaces with bad gl constant Recall first that the Gordon-Lewis constant, gl (X), of a Banach space X is defined to be gl(X)-
sup{Vl(T)" T ' X -+ ~2, yrl(T) <~ 11,
where yrl denotes the 1-summing norm (see [28, Section 10]) and v I ( T ' X --+ Y ) = inf{llA ]] liB ]]}. Here the inf is taken over all L1 spaces L and over all decompositions T -- A B with B " X--+ L, A" L -+ Y. An easy but very useful theorem of Gordon and Lewis ([26] or [60, p. 260]) says that the unconditional constant of every basis of X is at least gl (X). n By "bad" we mean When 2 < p ~< ~ there is an abundance of "bad" subspaces of gp. here lacking good unconditional bases or even the weaker property of small Gordon-Lewis constant. Recall that L np is L p over the measure space consisting of n points and endowed with the uniform probability measure. THEOREM 26. There are positive constants c, c t such that if X is any subspace o f L np, 2 < p <~ cx~, satisfying d i m X ~> cn and [[xllL~ <~ 2[[XOIL~ f o r all x E X then g l ( X )
ctnl/2-1/P. Theorem 26 was proved by Figiel and Johnson in [20]. A somewhat weaker theorem, still ensuring the abundance of subspaces with large g l ( X ) was proved earlier by Figiel, Kwapiefi and Petczyfiski [23]. We refer to [60, p. 261] for the proof of Theorem 26 for p -- cx~. The case 2 < p < c~ follows easily since the Banach-Mazur distance between L np and L ~ is n 1/p.
859
Finite dimensional subspaces of L p
Of course, for some c > 0, a random subspace X of L ~ of dimension cn satisfies the second assumption of Theorem 26, i.e., ]IxIIL~ ~< 2]]XI]L~ for all x e X (see [47] or [24]). This is why we claim that there is an abundance of subspaces of L f/p satisfying the conclusion of Theorem 26. For 1 ~< p ~< 2, g l ( X ) is uniformly bounded for any subspace X of Lp. Nevertheless, there are finite dimensional subspaces of L p which have only bad unconditional bases (see [37]). This implies that sup{ubc(X); X C ~ } --+ ~ as n --+ oo but no estimates are known.
4. Restricted invertibility and finite dimensional subspaces of distance to Euclidean spaces
Lp
of maximal
4.1. Restricted invertibility o f operators on s Hp Motivated by some problems about the structure of finite dimensional subspaces of L p that will be discussed in Section 4.2, in [10] Bourgain and Tzafriri [10] proved Theorem 27 about the restricted invertibility of operators on ~ . Qualitatively, this result says (or rather implies) that a bounded operator on s n which has ones on the diagonal must be invertible on a coordinate subspace of proportional dimension (even after projecting back into the coordinate subspace). In order to state Theorem 27 we introduce the following notation. Given a subset cr of {1, 2 . . . . . n }, let s be the span in g np o f the unit vector basis vectors O" {ei" i E cr } and let R~ be the natural coordinate projection from g 17p onto gp. THEOREM 27. Let 1 <~ p <~ cx~. For each e > 0 there is 6p(e) > 0 so that if T is an n by n matrix, considered as an operator on s np, with zero diagonal, then f o r each e > 0 there is a subset rr o f { l , 2 . . . . . n} o f cardinality at least 6p(e)n so that IIR~TR~IIp <~ ellTIIp. 1
Consequently, ifeliTllp < 1, then [I(R~(I + T)Ro-) -1 lip ~< 1-ellTllp
.
The case p -- l, as well as the case p = oo, which follows by duality, of Theorem 27 was proved earlier by the second author [30] and, independently, by Bourgain [4, p. 113]. Bourgain's argument gives more than what is stated in Theorem 27; namely, that there exists a splitting cq, a2 . . . . . ak of {1, 2 . . . . . n} into k ---=k(e) disjoint sets so that for each 1 ~< i ~< k, ]]R~/T R~/]]1 ~< e ]]T Ill. Whether this strengthening of Theorem 27 remains valid for other values of p is open. For p = 2 this matrix splitting question is particularly interesting because it is equivalent to the Kadison-Singer problem [35] whether every pure state on goc has a unique extension to a pure state on B(~2). For a discussion of the matrix splitting problem on ~2 and more on the connection between the Bourgain-Tzafriri work and the Kadison-Singer problem see [ 16]. For a proof (due to K. Ball) of the matrix splitting result for e~ which gives the estimate k(e) <, 2 / e see [ 11 ]. Notice that in the case of s there is no loss of generality in treating only matrices with nonnegative entries because as operators on ~ , T and IT] have the same norm. Berman, Halpern, Kaftal, and Weiss [3] independently used ideas similar to those used by Ball to prove an g~ splitting result for matrices with nonnegative entries (which of course does not give a splitting result on s for general matrices). It is amusing
860
W.B. Johnson and G. Schechtman
and instructive to note that the matrix splitting result for g~ formally implies the matrix splitting result for nonnegative matrices on g /7p for all 1 <~ p < cx~ via a change of density argument of Weis [62]. Here is the idea. First, after a change of density, a positive operator T on L p ( # ) (# a probability) is nicely bounded on L ~ ( # ) . This follows from the fact that T maps the order interval [ - f , f ] for some f > 0 into the order interval [ - ( 1 + e)llTlif, (1 + e)llTllf], which in turn follows via iteration and summing (as in the proof of 7) from the inclusion T [ - g , g] C [ - T g , Tg], which is valid for all g ~> 0 because T is a positive operator. Secondly, by working both with T and T*, one can get that, after a suitable change of density, a positive operator T on L p ( # ) (# a probability) is nicely n bounded on both on L1 (#) and on L ~ (#). Now specialize this to a positive operator on gp. The above discussion shows that this operator can be modeled as a (positive) operator T on /7 /7 /1/ L p ( # ) for some probability on {1, 2 . . . . . n} in such a way that ]IT'L 1 (#) --+" L 1 (/z) ][ and /7 /7 i]T" L ~ (#) --+" L ~ (#)][ do not exceed (2 + e)]] T" L p (#) --+" Lp (#)][. Since we know the splitting result for operators on g~ and ~ , the result for g np follows by interpolation. The proof of Theorem 27 for g /7p u s e s e /71 in a more serious way. By duality, it is enough to prove the case 1 < p < 2, so we restrict to this range of p. The "natural" approach to prove Theorem 27 is to show that if the set cr is chosen at random from among the subsets of { 1, 2 . . . . . n} having cardinality 6n for small enough 6 -- 3(e), then with big probability ]]R~ T R~ ]]p ~< e ]]T 6]p. This is, unfortunately, obviously wrong (consider the fight shift operator). However, regarding gpn as Lp/7 (so that the injection Ip, l ' L p/7 ~ L~ has norm one), it is true that for most such choices of cr the operator W := Ip, 1R~ T Ro has norm not e x c e e d i n g eS 1/p* IITIIp (the factor ~l/p* is natural; it goes away when we regard W as an operator from Lp into L~). For a proof, which is nice, and not particularly difficult, see [ 10, Proposition 1.10]. What is remarkable is that the p = 2 case in Theorem 27 then follows immediately by an application of Grothendieck's inequality via Proposition 1 ! Indeed, W* maps L n into a Hilbert space and thus has 2-summing norm at most KG IIW]I, where KG is Grothendieck's constant (see [28, Section 10]. Applying Proposition 1, we conclude that W = D U for some operator U on L /72 of norm at most eKG[]TiI2 and some norm one diagonal operator D ' L /72 --+ L /71. The operator D is multiplication by some function g which has norm one in L n2, hence U is defined by the formula U f = R~ VR~ g f
,
whence
11R~TRcrf 112 ~ eKGIITIIzIIflI2 for all f in L~. Since 11gl12 -- 1, I?[Igl >~ x/~-/6] ~< 6/2; that is, Ig(j)l ~> x/2-/6 for at most (6n)/2 coordinates j. Throwing away any of these which are in the set cr and calling the resulting set cr ~, we have that cr ~ has cardinality at least (6n)/2 and IIe~,ze,~, f ll2 <~~/2gG~llZll2. That completes the outline of the proof of Theorem 27 in the Hilbertian case p -- 2, which by itself has many applications (see [10,11]). It is, however, the other cases of Theorem 27 that have applications to the structure theory of finite dimensional L p spaces. The proof we sketched for p = 2 does not carry over because operators from Lec into Lp, need not be p*-summing when 1 < p < 2 (they are s-summing for all s > p*, but this is not sufficient). Lemma 28 is used to get around this problem. LEMMA 28. Suppose 1 < r < p < 2, #, v, r are probabilities, X is a subspace of L p ( # ) , T : X --+ L1 (v) and U : X --+ L p ( r ) are operators so that IITxlll ~< IIUx[lr,
x E X.
(26)
Finite dimensional subspaces of L p
861
Then 7rp,(T*) <~ y(r, p)llUII, where y(r, p) is the Lr-norm of a p-stable random variable which is normalized in the L 1-norm. In view of Corollary 6, to prove L e m m a 28 it suffices to verify the estimate M (p) (T) <~ V (r, p)II U I[, which obviously follows from the the following inequality, valid for all finite sets of vectors xi in X"
[Txi[ p i
)
1/p o
(27)
p
To prove (27), write (Y~i ITxilP) 1/p -- El ~'~i fi Txil where the fi are independent p-stable random variables with EIj~ J -- 1 and interchange the expectation and the L1 norm to see that this is estimated from above by
r) 1/r (28) 1
But (El ~J-~if i U x i l r ) l/r - v(r, P ) ( ~ i ]UxiJP) I/p, so (28) is dominated by the right side of (27). The main tool for proving Theorem 27 is: PROPOSITION 29. For each O < e < 10 -2 there is p - - p(e) > O so that if l < r < 2 a n d S is an operator on L 2n (n ~ n(e, r)) with ]ISII2 <~ p, then there is a subset ~ of { 1 . .2. . . . n}
of cardinality at least en so that for all x, IIR~Sxlll ~ CE(llxllr -+-IlSxllr),
(29)
where C is a numerical constant. Here we are using as usual the L np normalization. The important thing is that the factor
Ce in (29) is a gain over the trivial factor e l/r* . The p < 2 case of Theorem 27 follows easily from L e m m a 28, Proposition 29, and the p = 2 case of Theorem 27. Indeed, Corollary 8 says that to prove Theorem 27 it is enough to consider norm one operators on L /7 p which have norm at most 2Jp as operators on L~ The p - 2 case in Theorem 27 then says that it is enough to consider norm one operators onLpn which have norm at most p - p(e) as operators on L~ . Now if S is an operator on Lp/7 with IISllp = 1 and IlSll2 ~< p, apply Proposition 29 to get or. Define T = R ~ S R ~ , considered as an operator from L /7 p into L ~p (so that both the domain and range are L p spaces of a probability), and set U - 2C(R~ | S R~), considered as an operator from L /7p into
2
l~p
LZnp ~ LpOpLp.n/7 Then for, say, r -- ~ , we have from Proposition 29 that for all x in ~n ]ITX]IL 7 <~ JGUx[IL2~. L e m m a 28 then gives that yrp,(T*) <~ y(r, p)liU[ip <~ 4 C y ( r , p). /7 Changing back to the L /7p normalization (that is, regarding R~ SR~ as an operator on Lp),
862
W.B. Johnson and G. Schechtman
we have that Jrp,(R~rS* R~) <<.e l / P 4 C y ( r , p). The completion of the proof of Theorem 27 is now just as in the p = 2 case. For a proof of Proposition 29 see [ 11, Section 5]. Here is the idea: By the same kind of reasoning that works in the first part of the proof of Theorem 27, most choices of a subset cr of {1 . . . . . n} of cardinality en make it true that IlR~Sxlll <~ CellSxllr for all vectors x whose support has cardinality at most r n (an estimate for r = r (e) comes out of the proof). For a general vector x, apply this to 1AX, where A is the set of the r n largest coordinates of Ixl, and use the smallness of IISII2 to take care of x - lAx. Corollary 30 gives a method, to be exploited in the next section, for building complemented copies of g np in subspaces of L p /7 COROLLARY 30. Let 1 <~ p <<.cx~ and 1 > e > O. If T is a norm one operator from g.p into L p ( # ) and there are disjoint #-measurable sets Ai so that II1aiTeillp >/e f o r each 17 1 <<.i <<.n, then there is an operator S from L p ( # ) into g~p with IISII ~< 2/~ and a subset cr of{ 1 . . . . . n} of cardinality at least 6p (e/2) so that STle~p is the identity of g~p. Consequently, TRotS is a projection of L p ( # ) onto span{Tei}iE~r.
Here the function 3p(.) is taken from Theorem 27. For the proof of Corollary 30, take norm one vectors hi, supported on Ai, so that ( T e i , h i ) --IITeillp. Define S from Lp(lZ) n to ~pn by "S" f - E i = l II1Ae Tei lip - 1
Finite dimensional subspaces of L p
863
1 ~< p < 2; see [10, Section 4] (as noted in [10], the p = 1 case is essentially done in [20]). For 2 < p < cx~ it is open whether there is an n-dimensional subspace of L p(lZ), not isometric to gnp, which has B a n a c h - M a z u r distance n I1/p- 1/21 to g~ We turn now to the isomorphic theory. Here and in the sequel we shall speak qualitatively about finite dimensional spaces, similar to the way one speaks about infinite dimensional spaces. In order for our statements to have content, statements should be quantified so as not to depend on dimension (but there may be dependence on other parameters). It takes only one example to illustrate why this convention is followed by Banach space local theorists. One of the main results we shall discuss is the following: THEOREM 32. If X is an n dimensional complemented subspace of Lp(#), 1 < p < e~, whose distance to g.~ is of maximal order, then X contains a subspace Y of proportional k and complemented in L p (lz). dimension (say, k) which is isomorphic to g~p The meaning of this statement is that there are functions f (p, 6, C) > 0 and g(p, 6, C) < cx~ (defined for 1 < p < cxz, ~ > 0, and C < cx~) so that if X is an n-dimensional Ccomplemented subspace of Lp(#) for some n and p, and d(X, g~) >~ Sn I1/p-I/21, then k <<, for some k ~ f ( p , ~, C)n there is a k-dimensional subspace Y of X with d(Y, gp) g(p, 6, C) and Y is g(p, 6, C)-complemented in Lp(#). Theorem 32 is also true for p = 1 [30], but in a stronger form - the complementation assumption is not needed. Given the method of [30], the main step in the proof of Theorem 32 which we sketch below is Theorem 27. For p > 2 the complementation assumption in Theorem 32 is essential; this follows easily from Theorem 26. Whether the complementation assumption is needed when 1 < p < 2 is open. A criterion for building a complemented copy of gp in a subspace X of Lp(#) was given in Corollary 30. The idea for getting an operator T taking values in X which satisfies the hypothesis of the corollary is first to find vectors fi, 1 ~< i ~< n, in the unit sphere of X for which the norm of the square function $2 ({fi }i=l)n (where Sr({fi }i=l)n . _ (Zi=ln [filr)l/r for r < cx~ and S~({f/'}inl)"--maxl~ ~ n l / p - 1 / 2 ( ~ = 1 ]]j~ 112)1/2. Set 0 -- p / 2 and f / - - jS/llj{ II. Then, using the factorization
$2 ({j~ }in=l) -- ~
If/IPlIJ~ ll2-plf/12-p
i=1
•n 1/p-1/2
IIJ~ II2
IIs2(r
n/=l)] p
i=1
n )o n (1-o) <~ IlSp({fillf~ll<2-P~/Pli=, soo ({f/}i=1) n [Isp({SIISII<2-P /P},=,)II
IIs~({fi} ~/=~)llp<~-o>
864
W.B. Johnson and G. Schechtman
=
n
IIj~ II2
(1-0)
i=1
So~({j~ }/n=l)] and setting Aj - Aj ,.~ Ui<j 1~i, we get disjoint sets Ai so that 62/(2-p)nl/p <~ II Y~.inl 1Aifillp. Thus II1Aifi lip is larger than 62/(2-p)/2 for at least (62p/(2-p)/2)n values of i. Next suppose that 2 < p < oo and that j~, 1 ~< i ~< n, are norm one vectors in Lp(#) for which ]IS2({j~ }inl)ll ~< Cn 1/p. Set 0 - 2/p. Then Letting Aj = [ I f j l -
'/'
-
[Is,({y/}"
,:,)11 ~ [Is2({fi },,=1) 0 soo({j~ }ni:1)
iis2({fi}i%l)llOlla
({fi}n i--1) II,-o
C~ ~
,-0
II
IIS ({Ii }.,_,)I1 ,-o
Thus there are disjoint sets Ai so that [[1Ai f / l i p is larger t h a n 1/(2C 2/(p-2))) for at least (2cZP/(P-Z))-ln values of i. To derive Theorem 32 from Corollary 30 we need the sequence {j~ } for which l[ 1ai ft" [[p is bounded away from zero also to be dominated by the e p basis. This can be accomplished in the following way. Let 2 < p < cx~ and let P'Lp(lZ) --+ Lp(#) be the projection with range X. Assume that, for some k proportional to n, {j~ }~1 is a sequence in the ball of the range of P* in Lp.(#) satisfying [llaifillp > 6 > 0 for some disjoint sets {Ai}kl . Put gi -P ([j~ Ip*-I sgn(fi) 1ai )" Note that {gi }~1 is dominated by the epk basis and in particular the norm of S2({gi }) is dominated by k 1/p and that the norms of the gi-s are bounded away from zero since Ilgi II >/f gi fi >/6 p. It follows that for a proportion of the gi-s there are disjoint Bi-s for which Ilgi 18i II are bounded away from zero. Now Corollary 30 applies to produce a subspace of X of proportional dimension isomorphic to e p of its dimension and complemented in Lp(#). The case 1 < p < 2 follows by duality. We remark that this part of the argument, which is what we had in mind when we wrote [30], is considerably simpler than the one presented in [ 10]. Finally, we briefly indicate how to find in any n-dimensional subspace X of Lp(#) of maximal order distance from ~ a sequence {j~}in_l with Ilg2({f/})ll of order nl/p-1/2( E i =n l II~ II2 ) 1/2 . Assume for example that 1 < p < 2. By a theorem of Kwapien, [60, Theorem 13.15] the type 2 constant of X is of order n 1~p-l~2, and by a theorem of Tomczak [60, Theorem 25.6], this type 2 constant is attained, up to a universal constant, on n vectors; i.e., there are vectors {f/}in 1 in X satisfying IlS2({fi})l[ ~> ~nl/p-1/Z(Ein=l Ilf/ll2) 1/2-
5. Complemented subspaces n 5.1. Fine embeddings of complemented subspaces of L p in g.p
Let X be a k-dimensional subspace of L p, 1 ~< p < oo, and assume there is a projection of norm K from L p onto X. Let s > 0 and denote by Pp (X, K, s) the smallest n such
Finite dimensional subspaces of L p
865
that X (1 + e)-embeds in s /7 as a (1 + e)K-complemented subspace; i.e., the smallest n such that there is a k-dimensional subspace Y of g Hp, an isomorphism T ' X ~ Y with IIr IIIIT - 1 II ~ 1 + e, and a projection of norm at most (1 + e) K from s /7 onto Y. Also, denote by Pp(k, K, e) the maximal Pp(X, K, e) when X ranges over all k-dimensional subspaces of L p. If 1 < p < oc, looking at the case of X = s shows that, at least for some K depending on p, Pp(k, K, e) >~ c(p) max{kp/2, kp*/2}, where p* = p / ( p - 1). It turns out that, except for logarithmic terms, one can achieve this bound. THEOREM 33 ([8]).
Pp(k, K, e) <<,C(p, K, e) max{kP/2, kP*}(logk) c(p).
For p = 1 one gets a better result THEOREM 34. P1 (k, K, e) <~ C(K, e)k log k. The proofs follow the method of Section 2.1. Given a k-dimensional X C s m 1 < p < cx~, and a projection P on s m with range X, we let Y -- P*gp, and following the scheme of Section 2.1 we find a subset S-2 C {1, 2 . . . . . m } of the right cardinality such that
Ixil p -
mlS21-1
I
mlS21-1 ~
Ixil p < ~,
lyilP* - Z
lyiIp* < e,
y E Y, Ilyllp* -- 1,
(31)
i=l
xiyi iEs
(30)
i:1
iES-2
mlX?1-1
x E x , Ilxllp = 1,
m
iEs
xiyi < e,
x E X, y E Y, IIx lip -
Ilyllp* = 1.
(32)
i=1
For p = 1, (31) causes of course a problem. Fortunately enough, in this case the relevant inequality (stating that the restriction to S2 is a good isomorphism when restricted to Yc s follows immediately from (30), (32) and the fact that a restriction to a subset is a norm one operator on ~ . Finally, it is easy to see that (30), (31) and (32) imply the desired result. 5.2. Finite decomposition and uniqueness of complements The class of finite dimensional well complemented subspaces of L p is a rich class, at least in the range 1 < p r 2 < cx~. It follows from the previous section that the same is true for the well complemented k-dimensional subspaces of g np as long as k is smaller than a specific power of n. The situation for larger k is not known. In particular, it is not known whether a well complemented subspace of gp of proportional dimension is well isomorphic to a s k space. It is also not known whether s n is "primary"" i.e. whether g p - X @ Y implies that either X or Y is well isomorphic to some s k (here we mean of course that the isomorphism constant should depend on the norm of the projection on X with kernel Y (and p) only). It is true however that if Y is "small" enough then X is well isomorphic to an ~pk space. There are two known instances of this statement with two different notions
866
W.B. Johnson and G. Schechtman
of smallness. These were proved in [33] following somewhat weaker results in [9] for the first theorem and [8] for the second.
THEOREM 35. A s s u m e 1 < p < oe, e np _ X + Y, d (Y, g~) <~ K, and there is a projection P f r o m g.p onto X with kernel Y and Ilell <<. t . Then d ( X , g np-k) <
Let Z i , i -- 1 . . . . . m, be Banach spaces and, f o r i = 1 . . . . . m, let Yi be a K - c o m p l e m e n t e d subspace o f Zi. A s s u m e also d(Yi, Yi-1) <~ L, i = 1 . . . . . m. Then Yo G ~-~im=l G p Z i is CKL-isomorphic to Ym G Y~im=l G p Z i . In parsi i = 1 . . . . . m, and s - - ~i=1 m s is CKL-isomorphic to ticular, if Zi -- g.p, s i then Yo G gp Ym @ U p"
The proof is very simple. Let Xi be the complement of Yi in Zi, i - 1 . . . . . m. Then, m
YO E]~~
GpZi ~
Gp
9
~]~p
0
Gp X1
Gp
Gp X2
~]~p "" Op
Gp Xm
9 (33)
We now shift the top row to the right
@p
@p
X1
@p
Op.-.@p
X2
@p
Xm
@p
@p
(34)
0
to get m
m
Yo~ Z ~pZi ~ Ym E]~~--~~pZi. i--1
(35)
i=1
This picturesque proof can easily be justified. To prove Theorem 35 we assume that p > 2 and k - 2 m. We build a path Yo . . . . . Ym of spaces connecting Y0 - ~ with Ym - e kp with d (Y i, Yi-1) ~< 2. This can be done in several ways but, since we also want Yi to be well isomorphic to a well complemented subspace of e~ with si as small as possible, we take Yi to be the e p sum of 2 i copies of e 2m-i .
_2m-i
Given any 8 > 0, g-e embeds as a complemented subspace, with constants depending only on p and S, into e~ with ri - - [ ~ 2 ( m - i ) p / 2 ] . The fact that this holds for some 8 < oe is
Finite dimensional subspaces o f L p
867
exposed for example in [24]. To get it for all 6 > 0 represent s as the ep sum of u copies of ~2v (introducing of course a constant depending on u and p), embed each summand complementably in an appropriate s space, and take the s sum of these u spaces as the containing gp space. It follows that Yi well embeds complementably into s for si - [62 i2(m-i)p/2]. Lemma 37 implies then that, for s - Ziml si, ~k (~ e Sp ~ ~,p+k. The assumption that s is well complemented in s implies that n -- y k p/2 for some ), bounded away from zero (this is a result of [2]). It is now easy to see that, with the right choice of 6, s -- n - k and s G s ~ s It remains to show that if also s 2 @ X ,~ s np then X ~ s n-k . This follows from the following simple "uniqueness of complement" result of [9] in the form proved in [33]. In the statement " + " denotes a direct sum of subspaces and the isomorphism constants implicit in the notation " ~ " of the conclusions depend only on the constants for the " ~ " and the projections in the hypotheses. PROPOSITION 38. Assume Z = Y + X = H + G with H C X, and assume H ~ Y 9 W. Then X ~ G @ W. Inparticular, if f o r i = 1,2, Z = Yi + X i = Hi-+-Gi with Y ~ Yi, G ~ Gi, H ~ Hi C Xi and H ~ Y G W. Then X1 ~ X2. To end the proof of Theorem 35 we only have to show that X, the complement of Y ~ s in gp, contains a subspace well isomorphic to gk2 and well complemented. Since n is of order at least k p/2, the general theory of Euclidean sections as exposed in [24] implies that, for some 6 > 0, X contains a subspace U1 well isomorphic to s k. U1 is automatically well complemented (see [44] or [18, p. 46]). Now find another copy of s of U1 in X and iterate.
in the complement
PROOF OF PROPOSITION 38 ([33]). It is very easy: Put F = X N G. Then Z = Y + X = Y+F-+-HandconsequentlyG~YGF. ThusX=F+H~FGYGW~GOW. D PROOF OF THEOREM 36. It follows the same outline as that of the proof of Theorem 35. Given a k-dimensional well complemented subspace Y0 of L p, 2 < p < cx~, we find a path Yo, Y1 . . . . . Ym of well complemented subspaces of Lp with d(Yi, Yi-1) < 4, Ym well isomorphic to ~ , and m -- [log 2 k]. We then use Theorem 33 to embed each of Yi in a well complemented fashion in a low dimensional g p space and continue in a way similar to the proof of Theorem 35. To build the path of spaces Yi, apply first the change of density of Theorem 2 and then that of Theorem 7 to get that without loss of generality IlYllp ~< 2nl/2-1/Pllyllz for each y 6 Y0 and the projection P from L p onto Y0 is also well bounded with respect to the L2 norm. Put Yi--{(y,2iy);
Y6Yo}CLpGpL2,
i - - 1 , 2 . . . . . m.
868
W.B. Johnson and G. Schechtman
T h e n c l e a r l y d ( Y i , Y i - 1 ) < 4, i = 1 , 2 . . . . . m, Ym is w e l l i s o m o r p h i c to ~ a n d it o n l y r e m a i n s to s h o w that e a c h o f Yi is w e l l i s o m o r p h i c to a w e l l c o m p l e m e n t e d s u b s p a c e of Lp. S i n c e P is b o u n d e d in b o t h L p a n d L2, Yi is w e l l c o m p l e m e n t e d in Zi - {(z, 2iz); y E L p } C L p O p L2. Finally, b y a t h e o r e m o f R o s e n t h a l [52] (see also [1]) Zi is w e l l i s o m o r p h i c to a w e l l c o m p l e m e n t e d s u b s p a c e o f L p. D
References [ 1] D.E. Alspach and E. Odell, L p spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 123-159. [2] G. Bennett, L.E. Dor, V. Goodman, W.B. Johnson and C.M. Newman, On uncomplemented subspaces of Lp, 1 < p < 2, Israel J. Math. 26 (2) (1977), 178-187. [3] K. Berman, H. Halpem, V. Kaftal and G. Weiss, Matrix norm inequalities and the relative Dixmier property, Integral Equations Operator Theory 11 (1) (1988), 28-48. [4] J. Bourgain, New Classes of s Lecture Notes in Math. 889, Springer-Verlag, Berlin (1981). [5] J. Bourgain, Subspaces ofl N, arithmetical diameter and Sidon sets, Probability in Banach Spaces, V, Medford, MA, 1984, Lecture Notes in Math. 1153, Springer, Berlin (1985), 96-127. [6] J. Bourgain, Bounded orthogonal systems and the A(p)-set problem, Acta Math. 162 (3-4) (1989), 227245. [7] J. Bourgain, N.J. Kalton and L. Tzafriri, Geometry offinite-dimensional subspaces and quotients of L p, Geometric Aspects of Functional Analysis (1987-88), Lecture Notes in Math. 1376, Springer, Berlin (1989), 138-175. [8] J. Bourgain, J. Lindenstrauss and V.D. Milman, Approximation of zonoids by zonotopes, Acta Math. 162 (1989), 73-141. [9] J. Bourgain and L. Tzafriri, Complements of subspaces of l~, p >>,1, which are uniquely determined, Geometrical Aspects of Functional Analysis (1985/86), Lecture Notes in Math. 1267, Springer, Berlin (1987), 39-52. [ 10] J. Bourgain and L. Tzafriri, Invertibility of "large ""submatrices with applications to the geometry of Banach spaces and harmonic analysis, Israel J. Math. 57 (2) (1987), 137-224. [11] J. Bourgain and L. Tzafriri, Restricted invertibility of matrices and applications, Analysis at Urbana, Vol. II (Urbana, IL, 1986-1987), London Math. Soc. Lecture Note 138, Cambridge Univ. Press, Cambridge (1989), 61-107. [12] J. Bourgain and L. Tzafriri, Embedding lkp in subspaces of Lp for p > 2, Israel J. Math. 72 (3) (1990), 321-340. [13] J. Bourgain and L. Tzafriri, On a problem of Kadison and Singer, J. Reine Angew. Math. 420 (1991), 1-43. [14] J. Bretagnolle and D. Dacunha-Castelle, Application de l'dtude de certaines formes lindaires aldatoires au plongement d'espaces de Banach dans des espaces LP, Ann. Sci. t~cole Norm. Sup. (4) 2 (1969), 437-480 (French). [15] B. Carl and A. Pajor, Gelfand numbers of operators with values in a Hilbert space, Invent. Math. 94 (3) (1988), 479-504. [ 16] K.R. Davidson and S.J. Szarek, Banach space theory and "local" operator theory, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 317-366. [17] W.J. Davis, V.D. Milman and N. Tomczak-Jaegermann, The distance between certain n-dimensional Banach spaces, Israel J. Math. 39 (1-2) (1981), 1-15. [18] J. Diestel, H. Jarchow and A. Tonge, Absolutely Summing Operators, Cambridge Studies in Advanced Math., Vol. 43, Cambridge University Press, Cambridge (1995). [19] L.E. Dor, On projections in L1, Ann. of Math. 102 (3) (1975), 463-474. [20] T. Figiel and W.B. Johnson, Large subspaces of 1n and estimates of the Gordon-Lewis constant, Israel J. Math. 37 (1-2) (1980), 92-112.
Finite dimensional subspaces o f L p
869
[21] T. Figiel, W.B. Johnson and G. Schechtman, Factorizations of natural embeddings of l~ into Lr, I, Studia Math. 89 (1) (1988), 79-103. [22] T. Figiel, W.B. Johnson and G. Schechtman, Factorizations of natural embeddings of l~ into Lr, H, Pacific J. Math. 150 (2) (1991), 261-277. [23] T. Figiel, S. Kwapiefi and A. Petczyfiski, Sharp estimates for the constants of local unconditional structure of Minkowski spaces, Bull. Acad. Polon. Sci. Srr. Sci. Math. Astronom. Phys. 25 (12) (1977), 1221-1226. [24] A.A. Giannopoulos and V.D. Milman, Euclidean structure infinite dimensional normed spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 707-779. [25] E.D. Gluskin, N. Tomczak-Jaegermann and L. Tzafriri, Subspaces of l N of small codimension, Israel J. Math. 79 (2-3) (1992), 173-192. [26] Y. Gordon and D.R. Lewis, Absolutely summing operators and local unconditional structures, Acta Math. 133 (1974), 27-48. [27] W.B. Johnson and L. Jones, Every Lp operator is an L 2 operator, Proc. Amer. Math. Soc. 72 (2) (1978), 309-312. [281 W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84. [29] W.B. Johnson, B. Maurey, G. Schechtman and L. Tzafriri, Symmetric structures in Banach spaces, Mem. Amer. Math. Soc. 19 (217) (1979). [3O] W.B. Johnson and G. Schechtman, On subspaces of L 1 with maximal distances to Euclidean space, Proceedings of Research Workshop on Banach Space Theory (Iowa City, Iowa, 1981), Univ. Iowa, Iowa City, IA (1982), 83-96. n , Acta Math. 149 (1982), 71-85. [311 W.B. Johnson and G. Schechtman, Embedding lpm into 11 [32] W.B. Johnson and G. Schechtman, Remarks on Talagrand's deviation inequality for Rademacher functions, Functional Analysis (Austin, TX, 1987/1989), Lecture Notes in Math. 1470, Springer, Berlin (1991), 72-77. [33] W.B. Johnson and G. Schechtman, On the distance of subspaces of l~ to Ikp, Trans. Amer. Math. Soc. 324 (1) (1991), 319-329. [341 W.B. Johnson and G. Schechtman, Computing p-summing norms with few vectors, Israel J. Math. 87 (1-3) (1994), 19-31. [351 R.V. Kadison and I.M. Singer, Extensions ofpure states, Amer. J. Math. 81 (1959), 383-400. [36] N.J. Kalton, The endomorphisms of Lp (0 <<,p <~i), Indiana Univ. Math. J. 27 (3) (1978), 353-381. [37] T. Ketonen, On unconditionality in L p spaces, Ann. Acad. Sci. Fenn. Ser. A I Math. Dissertationes 35 (1981). [38] A. Koldobsky and H. Krnig, Aspects of the isometric theory of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 899-939. [39] S. Kwapiefi and C. Schtitt, Some combinatorial and probabilistic inequalities and their application to Banach space theory, Studia Math. 82 (1) (1985), 91-106. [40] H.E. Lacey, The Isometric Theory of Classical Banach Spaces, Die Grundlehren der Mathematischen Wissenschaften, Band 208, Springer-Verlag, New York (1974). [41] M. Ledoux and M. Talagrand, Probability in Banach Spaces, Isoperimetry and Processes, Springer-Verlag, Berlin (1991). [421 D.R. Lewis, Finite dimensional subspaces of Lp, Studia Math. 63 (2) (1978), 207-212. [431 D.R. Lewis, Ellipsoids defined by Banach ideal norms, Mathematika 26 (1) (1979), 18-29. [44] B. Maurey, Thdorkmes de Factorisation pour les Opdrateurs Lindaires gt Valeurs dans les Espaces L p, Astrrisque, No. 11, Socirt6 Mathrmatique de France, Paris (1974) (French). [45] B. Maurey, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [461 B. Maurey and G. Schechtman, Some remarks on symmetric basic sequences in L 1, Compositio Math. 38 (1) (1979), 67-76. [471 V.D. Milman and G. Schechtman, Asymptotic Theory of Finite-Dimensional Normed Spaces, Lecture Notes in Math. 1200, Springer-Verlag, Berlin (1986). [481 A. Naor and A. Zvavitch, Isomorphic embedding ofg.np, 1 < p < 2, into ~I l+~)n , Israel J. Math., to appear.
870
W.B. Johnson and G. Schechtman
[49] G. Pisier, Un thdorkme sur les opdrateurs lindaires entre espaces de Banach qui se factorisent par un espace de Hilbert, Ann. Sci. l~cole Norm. Sup. (4) 13 (1) (1980), 23-43 (French). [50] G. Pisier, On the dimension of the Ipn-subspaces of Banach spaces, for 1 ~< p < 2, Trans. Amer. Math. Soc. 276 (1) (1983), 201-211. [51] G. Pisier, Factorization of operators through L p ~ or Lpl and noncommutative generalizations, Math. Ann. 276 (1) (1986), 105-136. [52] H.E Rosenthal, On the subspaces of L p (p > 2) spanned by sequences of independent random variables, Israel J. Math. 8 (1970), 273-303. [531 G. Schechtman, Almost isometric Lp subspaces of Lp(O, 1), J. London Math. Soc. (2) 20 (3) (1979), 516528. [54] G. Schechtman, More on embedding subspaces of Lp in In , Compositio Math. 61 (1987), 159-169. [551 G. Schechtman, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [56] G. Schechtman and A. Zvavitch, Embedding subspaces of L p into gpN, 0 < p < 1, Math. Nachr. (to appear). [57] S.J. Szarek, Computing summing norms and type constants on few vectors, Studia Math. 98 (2) (1991), 147-156. [58] M. Talagrand, Embedding subspaces of L ! into 1N, Proc. Amer. Math. Soc. 108 (1990), 363-369. [59] M. Talagrand, Embedding subspaces of L p in l N , Geometric Aspects of Functional Analysis (Israel, 19921994), Oper. Theory Adv. Appl., Vol. 77, Birkh~iuser, Basel (1995), 311-325. [60] N. Tomczak-Jaegermann, Banach-Mazur Distances and Finite-Dimensional Operator Ideals, Pitman Monographs and Surveys in Pure and Applied Math., Vol. 38, Longman Scientific & Technical, Harlow (1989). [61] L. Tzafriri, Uniqueness of structure in Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [62] L. Weis, Integral operators and changes ofdensity, Indiana Univ. Math. J. 31 (1) (1982), 83-96. [63] P. Wojtaszczyk, Banach Spaces for Analysts, Cambridge Studies in Advanced Math., Vol. 25, Cambridge University Press, Cambridge (1991). [641 A. Zvavitch, More on embedding subspaces of L p into gnp, 0 < p < 1, GAFA Seminar 1996-2000, Lecture Notes in Math., Vol. 1745, Springer (2000), 269-280.
CHAPTER
20
Banach Spaces and Classical Harmonic Analysis
S.V.
Kislyakov*
POMI, Fontanka 27, 191001 Saint Petersburg, Russia E-mail: skis @pdmi. ras. ru
Contents 1. Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
873
2. Averaging. Complementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
874
3. Invariant projections in H I(T ') . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4..........................................
875 ................
875
5.......................................................... 6. p - s u m m i n g operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
876 877
7. Grothendieck theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
878
8. Invariant Grothendieck theorem
.......................................
879
9..........................................................
880
10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
880
11. A counterexample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
882
12. Stein theorem
883
.................................................
13. Ap-sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14. Sidon sets
...................................................
15. Gordon-Lewis local unconditional structure
................................
16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
883 885 886 887
17. Quasi-Cohen sets vs quasi-Marcinkiewicz sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
889
18. Sidon sets and arithmetic diameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
891
19. Sidon sets and cotype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
892
20. Multipliers on spaces of vector-valued functions 21. B-convexity and K-convexity
..............................
........................................
893 894
22. UMD-spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
894
23 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
895
24 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
895
References
896
.....................................................
*Supported in part by the Russian Foundation for Basic Research, Grant. no. 96-01-00693. H A N D B O O K OF THE G E O M E T R Y OF B A N A C H SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 871
872
S. V. Kislyakov
Abstract Some points of contact of the two fields are discussed, specifically: projections onto translation invariant subspaces, Cohen's theorem and related results, multipliers of H 1, the use of invariant means, p-summing and p-integral operators, the vicinity of the Grothendieck theorem, some consequences of the Maurey-Nikishin-Rosenthal factorization theorem, Ap-sets and Bourgain's solution of the Ap-problem, translation invariant subspaces without GordonLewis local unconditional structure, Sidon sets, multipliers on spaces of vector-valued functions, specific spaces (related to harmonic analysis) in the general theory. Some proof are indicated or even exposed in detail, in case they are not technical and help to better illustrate the interplay between the fields in question.
Banach spaces and classical harmonic analysis
873
Many classical Banach spaces admit a natural action of some group, and many specific operators commute with translations. This features can be used in the study of such spaces and operators. Reciprocally, sometimes the techniques of the Banach space theory apply in harmonic analysis. This is nearly all that can be said about the subject of this paper in general. Furthermore, the manifestations of this relationship are scarce and heterogeneous. Invoking the polemical metaphor of Kahane and Salem for their classical book [ 17] on Fourier analysis ("il peut ... ressembler en quelque sorte h un herbier" [17, Pr6face]), it can probably be said that not only may the present text resemble somewhat, but it really is a "herbarium", much smaller and less systematized than [ 17], representing a far less explored taxon, and incomplete even in the known part of the latter. In other words, the subsequent discussion can be viewed simply as a collection of examples. I hope, however, that the reader will find some intrinsic logic in them, and that at least sometimes he will be amused by the interplay of the two fields mentioned in the title.
1. Basic definitions Throughout, "a group" means "an Abelian group". Any group G acts on functions on it by translations f w-> f x , where f x ( y ) = f ( y + x), x, y c G. A linear space X of functions on G is said to be translation invariant if with every function f it contains all translates of f . Let X and Y be two translation invariant spaces on G, and let T :X ----> Y be a linear operator. Then T commutes with translations if T ( f x ) = ( T f ) x for f 6 X, x 6 G. If G is compact (which will be assumed in what follows unless otherwise is claimed explicitly), we denote by dx the normalized Haar measure on G, by F the dual group, and by f the Fourier transform of a function f 6 L I(G) (the same notation is used for the Fourier transform of a measure). Mainly, we shall deal with the classical spaces C (G), LP(G) (1 <, p <<,oc), M ( G ) = C(G)*. Let X be any of these spaces, and let E C F. Then def
XE -- { f E X" f (v) -- O for v 9 E}
is a translation invariant subspace of X. If X = C ( G ) or X = LP(G) with 1 ~< p < ec, all translation invariant subspaces are such. Next, the spaces L~c (G) and M E ( G ) are the only w*-closed translation invariant subspaces of L ~ (G) and M ( G ) , respectively. In an obvious way, the symbol XE makes sense for other spaces X, and we shall use this notation without further explanations. Now, let each of X, Y be either LP, 1 ~< p < ec, or C, and let El, E2 C F . Suppose we are given a bounded linear operator T :XEj ---> YE2. Then T commutes with translations if and only if it is representable in the form (T f ) A -- m . f ,
(1)
where m is a bounded function on F vanishing on the complement of E1 CI E2 (note that, unless X -- Y = L 2, not every bounded m gives rise to a bounded operator via (1)). Often,
S. V. Kislyakov
874
m is called the symbol of T. A similar characterization holds if the spaces L ~ or M are involved, under the assumption of w*-continuity of T. We write T = Tm if T acts in accordance with (1); also, T is called the multiplier with
symbol m. Among specific examples of translation invariant spaces, we mention the Hardy spaces
H p (~n), where ~n is the n-dimensional toms. The dual group of 'Tn is Z n, and H p (~n) is simply the space L(z+)nP(Tn), 1 ~< p ~< c~. The space C(z+)n(T n) is called the polydisk algebra (the disk algebra if n = 1). All these spaces have a well-known interpretation as the traces on the distinguished boundary ~n of certain spaces of functions holomorphic in the polydisk.
2. Averaging. Complementation A most usual idea in the study of translation invariant subspaces is to average something. We start with simple and old examples; further, we shall come across several instances in which the realization of this idea is more intricate. Again, let each of X, Y be one ofthe spaces LP(G) (1 ~< p < c~) or C(G), and let El, E2 C F . For every bounded linear operator T :XE1 --+ XE2, we put
-
T f --
L (T fx)_x dx.
Then T commutes with translations. The mapping T ~ T is a norm 1 projection, and T can be referred to as the invariant part of T. Clearly, T inherits many properties of T, but it may happen that T - 0 . The oldest and most well-known applications in which T is quite substantial are related to projections. Supposethat X -- Y and E2 C ~ El. If P is a projection of XE1 onto XE2, it is easily seen that so is P. The symbol m of P is none other than the characteristic function of E2. In particular, it follows that the subspace L p (1 ~< p < c~) (respectively, CE) is complemented in L p (in C) if and only if the multiplier with the symbol XE is bounded on L p (on C). PROPOSITION 1. For 1 <<.p < cxz, p ~ 2, and G compact infinite, some of the subspaces
LP (G) are uncomplemented in L p (G). PROOF. By the preceding discussion, otherwise the characters of G form an unconditional basis in L P (G). From the Khinchin inequality it is easy to deduce that this is not so if p ~ 2. [2 In the case of L1 (G), much more can be said. We refer the reader to [13] for the proof of the following statement.
THEOREM 2 (Cohen's idempotent theorem). L I ( G ) is complemented in L I ( G ) if and only if E is in the coset ring of F.
Banach spaces and classical harmonic analysis
875
By definition, the coset ring of F is the smallest system of sets containing all cosets of subgroups of F and closed under finite unions and intersections and under complementation. It is known that translation invariant operators of L 1(G) or C (G) into itself are precisely the operators of convolution with a finite measure. The projections among such operators correspond to the idempotent measures/z (/z 9 # - #, or (/2) 2 --/2). So, Theorem 2 describes also the idempotent measures. Now, it is clear that the complemented translation invariant subspaces of C (G) admit a characterization similar to Theorem 2. A measure # on G is said to be quasiidempotent if ]/2]2 ~> ]/2l (i.e., for each V 6 F either/2(V) -- 0 or ]/2(V)] ~> 1). A set E C F is called a quasi-Cohen set if there exists a quasiidempotent measure # such that E -- {V" /2(9/) ~: 0}. This notion was introduced in [23] and will be used later on in this paper.
3. Invariant projections in H 1(T) The subspace H ~ ( T ) may be complemented in H 1(T) for some E C Z+ not belonging to the Boolean ring generated by the arithmetic progressions and singletons (i.e., here the picture is different from that described in Theorem 2). The simplest example is E = { 1 , 2 , 4 . . . . . 2 n . . . . }. THEOREM 3 (Paley; see, e.g., [16,38]). For f c H i ( T ) we have
~< cllflll. n~>0 A similar inequality is valid for any Hadamard lacunary set E (we recall that E C Z+ is said to be Hadamard lacunary if E = {ml, m2 . . . . } with mk+l ~> Xmk for a fixed constant X > 1). From the properties of lacunary trigonometric series, it follows that such an inequality means precisely that TxE is a bounded operator on H 1. THEOREM 4 (see [22]). The subspace H~ (T) is complemented in H i ( T ) if and only if E is in the Boolean ring of subsets of Z+ generated by singletons, arithmetic progressions,
and Hadamard lacunary sequences.
0
Now we discuss the complemented translation invariant subspace of L 1(G) for G locally compact but noncompact. H.E Rosenthal showed (see [33]) that for such a space X the set
hX -- {v r F: f (y) - O for all f e X} must be in the coset ring of Fd, i.e., of the group F dual to G and endowed with the discrete topology. This is, however, rather far from a complete description. We quote a final result for the group R (see [1 ]).
876
S. V. Kislyakov
THEOREM 5. A translation invariant subspace X is complemented in L I(R) if and only if h X = U 1~i ~n (OliZ -'~ ~i) \ F, where F C R is a finite set, {xi } and {~i } are two finite collections o f reals, and the numbers t~i are pairwise rationally independent. The above general result of Rosenthal is again based on averaging, and we show how this procedure applies this time. Let P be a projection onto a translation invariant subspace X of L 1(G). Then P* : X* --+ L ~ (G) is a simultaneous extension operator, i.e., P* extends each bounded linear functional on X to a bounded linear functional on L I(G) (and the latter is identified with a function in L ~ (G), as usual). Now, we define a simultaneous extension operator C :X* --+ L ~ (G) commuting with translations as follows:
(CF, g) - mx(F, ( P ( g x ) ) _ x ) ,
F E X*, g ~ L1 (G).
Here m is a fixed invariant mean on l ~ (G), i.e., a linear functional satisfying m(1) -- 1, m ~ O, a n d m ( f x ) = m ( f ) for f ~ l ~ ( G ) , x ~ G (the subscript x in the notation mx shows that m is applied "in the variable x"). For G commutative, such a functional always exists, see, e.g., [14]. Now, we see that the operator A : F ~ F - C ( F [ x ) is a projection commuting with translations and taking L 1(G)* = L ~ (G) onto X • = {F 6 L 1(G)*: F [ x = 0}. Eventually, it turns out that Cohen's theorem can be applied to deduce Rosenthal's result, but some additional analysis is needed to do this. The crucial step is the observation that on the Bohr compactification G of G we can find a measure/z such that A f -- f 9 lz for every continuous almost periodic function f on G. Since the dual group of G is precisely Gd, we easily arrive at Rosenthal's result with the help of Theorem 2. See [33] for more details.
0
Averaging against invariant means can also be used to show that, strikingly, certain statements involving the uniform structure of a Banach space can be linearized. Let X be a Banach space, and Y its closed subspace. Assume that there exists a linear operator v mapping Y* to the space of uniformly continuous functions on X such that v f is an extension of f for every f E Y*. Next, assume that v is continuous for the topology of uniform convergence on bounded sets. THEOREM 6 (see [28, Remarks to w Proposition A]). Under the above assumptions, there exists a bounded linear operator w : Y* -+ X* extending each y* ~ Y* to the entire space X. An operator v as above exists if, e.g., Y is a uniform retract of X (we may put v f = f o p if p : X --~ Y is a uniform retraction). We infer that if, say, Y is reflexive and uncomplemented in X, then Y is not a uniform retract of X. In particular, no infinite dimensional reflexive subspace of C ( K ) is a uniform retract of C ( K ) .
Banach spaces and classical harmonic analysis
877
PROOF OF THEOREM 6. The operator w is given by the formula
(wf)(z)-
fx
f [(vf)(x+ y + z ) -
(vf)(x + y)]dydx,
where by f r " .dy and f x ' " .dx we denote some fixed invariant means on l ~ ( Y ) and 1~ (X) (X and Y are regarded as Abelian groups under their own operations of addition). The definition of w is consistent because the "integrand" is bounded (note that, separately, the terms in the "integrand" are unbounded). That w has the required properties is easy. The external "integration" is responsible for the additivity of (wf)(.), while the internal one for the relation ( w f ) [ y = f . D
6. p-summing operators Let G be a compact space, X a subspace of C ( G ) , and T : X --+ Y a p - s u m m i n g operator (1 ~< p < oc). By the Pietsch theorem (see "Basic concepts"), the latter is equivalent to the existence of a probability measure/z on G such that
IITfll <<.c P R O P O S I T I O N 7.
Ifl p d#
,
f E X.
(2)
If G is a compact group, X is translation invariant, and T satisfies
1[Tfx [1 = [[T f I[ ( f e X, x E G), then in (2) we can take the Haar measure of G as #. PROOF. This is again done by averaging: we substitute fx for f in (2), raise to the power p, and integrate in x against the Haar measure. D The condition on T in Proposition 7 is fulfilled if, say, Y -- L r (G) or Y = C ( G ) , and T commutes with translations. PROPOSITION 8. Let 1 <. p < oo, and let E C F. The identity embedding i: CE(G) --+
LP (G) is p-integral if and only if the multiplier with symbol XE is bounded on LP (G). PROOF. The "if" part is clear. We prove the "only if" part in the case where p > 1 (only slight complications arise for p = 1, but we do not dwell on this). If i is p-integral, we have the following factorization:
C(G) *-~ CE(G)
c~(~)
i
id
> LPE(G)
> Le (~),
where u, v are bounded linear operators,/z is a probability measure, and id is the identity inclusion. By the extension property of L ~ , there exists an extension w of u making the
S. V. Kislyakov
878
diagram commutative. Thus, the operator T -- vid w is a p-summing extension of i to C (G). It is easily seen that the averaged operator T, N
T f --
-
L
(T fx)-x dx,
is a p-summing extension of i to C(G) that commutes with translations. Now, Proposition 7 shows that T acts in fact from LP(G) to LP(G); consequently, the invariant projection of LP(G) onto LP(G) is bounded. E] By Proposition 1 we now conclude that, if G is infinite, some of the embeddings
C E(G) --> L P(G) are not p-integral (though, clearly, all of them are p-summing). This observation was made in [27]. Trace duality then yields examples of quasi-p-nuclear operators that are not p-nuclear.
7. Grothendieck theorem
This theorem says that any bounded linear operator T" 11 __+ 12 is 1-summing. By the lifting property of l 1, to prove this it suffices to exhibit a single 1-summing operator onto 12. Such an example can be provided by harmonic analysis (loosely, there is a multiplier with this property). This way to the Grothendieck theorem was indicated by Petczyfiski and Wojtaszczyk around 1977. PROPOSITION 9. The operator r "C7;+ (T) --+ 12 given by r f -
{f(2 ~)}n/>l is 1-summing
and onto. PROOF. The fact that r is 1-summing is a consequence of Theorem 3. To show that r is onto, it suffices to prove the estimate IIr*x II >/cllx II, x ~ 12, which, by the F. and M. Riesz theorem (see, e.g., [16]), is equivalent to the estimate Z
~ c(~
xn~2n -+- p ( Z )
lXnl2) 1/2
LI(T)
n>/1
for every analytic polynomial p. We use the classical fact that the multiplier with symbol Xz_ maps L 1(T) to Lr (,~) for every r 6 (0, 1). This yields _2 n
Z XnZ + p(z) n>/1
L 1(T)
• n>/1
Xn Z 2n Lr
It remains to refer to the classical properties of lacunary series. Another curiosity of similar nature is the operator acting by the same formula as r, but defined on the space CE(T), where E -- Z_ U {2n}n~>0. It maps CE(T) onto 12 and
Banach spaces and classical harmonic analysis
879
is r-summing for every r > 0. This yields a direct proof (again via harmonic analysis) of Maurey's extension of the Grothendieck theorem: every operator from 11 to 12 is rsumming for every r > 0. See the survey [20] for the details.
8. Invariant Grothendieck theorem We may state the Grothendieck theorem like this: If H is a Hilbert space, then for every operator u : H --+ C(K) the adjoint u* is 1summing. The problem of replacing C(K) in this statement by some other spaces has received considerable attention. In particular, it is reasonable to ask about the spaces of the form CE(G) that verify this theorem. We mention some cases in which this is indeed so. (i) F \ E is a Ap-set for some p > 1 (equivalently, the space L l \ E is reflexive" Apsets will be discussed later in this article). (ii) G = qF, E = Z+ (then CE(G) is the disk algebra). (iii) G : qrn, E = Z n \ (Z_) n . (iv) G = ~', E = Z+ U {-2n}n~>0. (v) G -- •, E -- Z+ U (Uk~>0[-22k+l, --22k]). Statement (i) is due to the author and Pisier; see, e.g., [10,31 ] for the proof. More generally, if Y is any reflexive subspace of L I(K, lz), then the space Y• = { f 6 C ( K ) : f g f d # -- 0 for all g E Y} verifies the Grothendieck theorem, so that translation invariance is in fact irrelevant here. Statement (ii) is a celebrated result of Bourgain, the proof of which is based on fine Hardy space theory; see [11,20] for more information and references. Statement (iv) is easy modulo (ii), see the survey [20]. Statements (iii) and (v) were proved, respectively, in [37] and [21]. Curiously, the arguments are somewhat similar. They are based on the ideas of the proof of (ii), but involve some additional techniques (real variable Hardy spaces, Littlewood-Paley decompositions, etc.). We see that, in spite of the (accidental) involvement of translation invariant subspaces, the methods leading to (i)-(v) can hardly be classified as deserving a more detailed exposition in this paper (at least, they deviate too much towards only one of the two items mentioned in the title, whereas the paper is about the vicinity of the conjunction "and"). The situation becomes different if we restrict ourselves to operators commuting with translations. DEFINITION (see [23]). A set E C F is called a Marcinkiewicz set (respectively, a quasiMarcinkiewicz set) if the multiplier with symbol )(E is of weak type (1,1) (respectively, there exists a bounded function x on F such a that ~(Y) = 0 for V r E, [ot(~,)[ ~> 1 for V 6 E, and the multiplier with symbol ot is of weak type (1,1)). We remind the reader that an operator T is of weak type (1,1) if meas{ITgl > t} ~< ct -lllgllt~, for every g in the domain of T.
t > O,
880
S. V. Kislyakov
THEOREM 10 (see [23]). I f E is a quasi-Marcinkiewicz set, then f o r every operator u : L2(G) ---->CE(G) commuting with translations the adjoint u* is 1-summing. The proof will be given in Section 10. It is well known that Z+ is a Marcinkiewicz set for G = q[', but then Theorem 10 gives much less than statement (ii) above. The theory of martingale transformations yields many examples of Marcinkiewicz sets in the dual of the group Fin Zkn (see, e.g., [ 15]). Let, for instance, kn = 2 for all n = 1, 2 . . . . (the dyadic group D). We agree that Z2 = { - 1 , 1 } (written multiplicatively). Let 8j be the j t h coordinate function of D; then all characters of D except the function 1 are of the form ej~ e j 2 . . . ejN, where N = 1, 2 . . . . . jl < j2 < 9" < jN. Fixing an arbitrary set B of integers, consider the set E of the characters for which jN ~ B in the above representation. Then E is a Marcinkiewicz set.
0
It is interesting that Theorem 10 can be adapted to the space of smooth functions C (k) (Tl). Neglecting the constants, we may endow this space with the norm Ilfll~ = m a x [ D ' f [ Isl--k equivalent to the standard one (as usual, D s denotes the partial derivative corresponding to a multiindex s). Thus, in a natural way, C (k) (qFI) can be regarded as a subspace X of
c(T') e...
9c(T')
(the number of summands is equal to card{s: Isl = k}), and X is invariant under simultaneous translation on all copies of T 1. Moreover, the orthogonal projection onto the closure of X in L2(T l) 9 G L2(qFl) is of weak type (1,1). In [23] it was shown that, using this projection, it is possible to work with X as if we had a space of the form CE(G) with a quasi-Marcinkiewicz set E. (The reader may trace this inspecting the proof in Section 10.) In particular, it follows that any operator u : L2(q['l) ---> C(k)(qr l) commuting with translations has 1-summing adjoint. See [23] for the details.
10. Theorem 10 is proved by two-fold averaging. The first averaging yields a useful lemma stated below. Let each of X and Y be either L p (G) or CE (G) (maybe, with different E's), let T :X ---> Y be an operator commuting with translations, and let S: Y ---->X be a finite rank operator. LEMMA 1 1. We have trace T S - trace T S_ where "Sf = fG ( S f x ) - x dx.
Banach spaces and classical harmonic analysis
881
PROOF. Indeed, let ~rx be the translation f v-+ fx. By the invariance of T we have trace TS --trace(or7 | T~xS) --trace(TcrxS~s and it suffices to integrate in x over G. Now, we remind the reader that the trace duality identifies the ideal of operators with 1-summing adjoint as the dual of the ideal F1 of Ll-factorable operators; see [10]. In particular, for u : L2(G) --+ CE(G) we have 7rl (u*) - sup{trace(uv)" v'CE(G) --+ L2(G) is of finite rank and Yl (v) ~< 1 }. Here, as usual, Y1 stands for the norm on F1. By Lemma 11, if u commutes with translations, the supremum can be restricted to v commuting with translations. We see that it suffices to estimate the nuclear norm vl (v) in terms of Y1(v) for any such v: Vl (v) ~< Cyl (v). We show a more general statement, without the finite rank assumption. THEOREM 12. If E is a quasi-Marcinkiewicz set, then every 1-factorable operator v: CE(G) --+ L2(G) that commutes with translations is nuclear. PROOF. Again, we use an averaging procedure, but this time it is intricate. Let T be a multiplier of weak type (1,1) the symbol ot of which satisfies or(y) = 0 for y r E and Iot(?')l ~> 1 for Y c E. Clearly, T maps L I(G) to LI/2(G). Next, since v is L|-factorable, from the Maurey extension of the Grothendieck theorem (see the end of Section 7) we deduce that v is 1/2-summing. Writing the definition of a 1/2-summing operator in the integral form, we see that if (12,)~) is a measure space and F : C2 --+ CE(G) is a reasonably good function, then
f
llvF(~o)lll2/2 d~. (o)) 7rl/2(v)l/2suplJ; I(F(o)),/1)l'/2d)~(o)): /1 e M(G), Iltzll ~<
1}.
As (S-2,)0 we choose the group G with the Haar measure, and we put F(o)) -- T(f~o), co E G, where f is a trigonometric polynomial (a finite linear combination of characters) on G. Clearly, F indeed maps G to CE(G), and we obtain, using invariance:
]]vTfl] 2 ~< Yrl/2(I))
sup
](Tfo), lz)l 1/2dO)" /1 6 M(G), ]l/ill ~< 1 .
Putting/2@) = / 1 ( - e ) and using the fact that T maps L ] to L |/2, we see that f c I(Tfo~, ~)1 ~/2 d ~ -
f c I T ( f */2)11/2 ~< CIIf * ~1111/2 ~< CIIfll~1/2
S. V. Kislyakov
882
It follows that vT is a bounded operator from LI(G) to L2(G). Now, let b be the symbol of v. We have
.__, y6E
<. c
III
for every f E L l(G). Letting f run through some approximate identity, we infer that ~y~F Ib(y)l 2 < c~. Thus, v is the convolution with an L2-function V the spectrum of which is contained in E. So, v factors as follows:
C(G)
l
ce(a)
id > LI(G )
l
~> L2(G),
where fi is again convolution with V. Since id is 1-integral and fi maps to a reflexive space, v is nuclear. D
11. A counterexample Returning to the beginning of Section 8, let us agree to say that a space X verifies the Grothendieck theorem if every operator from 12 to X has 1-summing adjoint. From the discussion in "Basic concepts" (see [ 10] for more details), it can be seen that this happens if and only if every operator from X to I 1 is 2-summing. From (ii) and (iv) in Section 8 we see that the spaces Cz+ ('IF) and CE (ql"), where E = Z+ U {-2n}n~>0, verify the Grothendieck theorem. However, their injective tensor product does not. Indeed, this injective tensor product is none other than Cz+ xe (~'2), and the latter space contains a complemented copy of 11 (then, surely, the projection to this copy is not 2_2 n summing). This copy is spanned by the functions {zZ"z2 }n~>0.To exhibit a projection, we observe that the characteristic function of {(k, - k ) : k E Z} C Z 2 is the Fourier transform of a measure on qr x "IF.The required projection is given by convolution with this measure. It is known that the dual spaces of Cz+ and C E are of cotype 2 (see, e.g., the survey [20]). The above construction can be adapted to show that the projective tensor product of two cotype 2 spaces may fail to be of cotype 2, moreover, it may contain a complemented copy of co. The conjectures disproved by the above were perceived as natural for some time in the past. A "more radical" answer to these and many other questions is given by Pisier's celebrated counterexamples (see, e.g., [31 ]). I have included the above material (first published in [20]) because of its transparence and relevance to harmonic analysis.
Banach spaces and classical harmonic analysis
883
12. Stein theorem Seemingly, in the proof of Theorem 12 we did not use the full-scale assumption that E is a quasi-Marcinkiewicz set (we needed "only" the (L 1 _ L 1/2) continuity of the corresponding multiplier). The result of Stein stated below shows that this is not so. We denote by S(G) the space of all measurable functions on G endowed with the (metrizable) topology of convergence in measure. THEOREM 13. Let A C F, and let T'L1A (G) --+ S(G) be a continuous linear operator commuting with translations. Then T is of weak type (1,1). PROOF. The point is that something similar can be said if T does not necessarily commute with translations. Specifically, by the Nikishin theorem (see [25]), in the latter case for every e > 0 we can find a set X-2eC G with IG \ X-2cI < e such that I{t ~ X2s" I(rf)(t)[ > ~.}l ~< Cs)~-1 ]If Ill,
f ~ Llx, X > 0.
(3)
Now, if T does commute with translations, it only remains to average. We fix, e.g., e = 1/2, and put s = 1-21/2. Substituting fx for f (x 6 G) and using invariance, we rewrite (3) in the form
f xY2-xX{[Tf[>)~} ~ C~.-~ Ilfll~; integrating in x over G, we get
IX21 [{ITfl > 9~}] ~ C)~-I Ilfll~. 13. Ap-sets Let 1 <~ p < oo. A subset E of F is called a Ap-set if in L~ (G) the norm convergence is equivalent to the convergence in measure. This is the same as if we demand the equivalence of the L p- and Lq-norms on L~ for some q < p. A subspace X C L 1(#) (where # is a finite measure) if reflexive if and only if the norm convergence on X is equivalent to the convergence in measure; see [18]. Thus, E is a A 1-set if and only if L 1 (G) is reflexive. The most well-known examples of Ap-sets (for all p at once) are Hadamard lacunary sets of integers (here, of course, we have in mind the circle group T as G). In general, for a fixed E the set {p: E is of type A p } is an interval. In the pioneering paper [34] of Rudin it was shown that for any integer n ~> 2 there is a A2n-Set not of type A2n+e for any e > 0. The main idea of the construction is that I ~ • a• yl2, expands explicitly as a linear combination of products of the characters in E and their complex conjugates, so that we may play with arithmetic conditions on E to ensure huge cancellation after integration.
S. V. Kislyakov
884
This leads to nontrivial examples of A zn-sets, some of which turn out to be not of type AZn+e, 6 > O. The question as to whether similar examples exist for p ~: 2n, n >~ 2, n 6 Z, was often referred to as the A p-problem. For p < 2, the solution of this problem has turned out to be easy (at least formally). Namely, if p < 2, then every Ap-set is a A p+~-set for some e > 0. As in the preceding section, the reason is in the Rosenthal-Nikishin-Maurey factorization theorem saying, in particular, that if X is any subspace of L P with p < 2 on which the topologies of LP and S are equivalent, then, after a change of density, X becomes a subspace of Lr with some r > p:
(f
x i o)1'
4c
(/
x
lip
xEX,
where a is a positive weight, f a = 1. See [25] for the details. Now, if X is translation invariant, we easily get rid of the density by averaging, which implies our claim. This observation was made in [2]. However (in reality, so-to-say), no specific examples of Ap-sets with p < 2 that are not Az-sets seem to be known (this is the precise meaning of the above remark on the formality of the solution in question). For any p > 2, Bourgain [5] proved the existence of Ap-sets that are not of type Ap+~. (Thus, the Ap-problem remains unresolved for p -- 2 only.) Bourgain's proof is probabilistic and combinatorial in nature. Technically, it has little in common either with harmonic analysis or with Banach space theory, though philosophically the result may be linked with the latter. It is a known fact of the finite-dimensional theory of Banach spaces that if X is a space of dimension n and its cotype p constant is at most C, then a typical subspace of X of dimension [C1 n z/p] is 2-distant from the Hilbert space of the same dimension. Here C1 depends only on C. See [26, Theorem 9.6]. Thus, in any n-dimensional subspace X of L p (p > 2) there are many nearly Hilbertian subspaces of dimension proportional to n z/p. Now, suppose that X is spanned by characters 9/1. . . . . Yn. If we succeed in finding a subspace Y of X also spanned by characters, of dimension roughly n z/p, and such that the LP-norm is equivalent to the LZ-norm on Y with a constant independent of X, we are very close to the desired example. (In other words, we need to know that subspaces with "typical" behavior occur even among quite specific spaces, namely, among those spanned by characters.) To understand why this finite-dimensional statement suffices, we note that in many cases the exponent 2 / p in the above discussion is optimal. If we restrict ourselves only to the spaces X having this property, then on the above Y the best constant of equivalence of the L p- and L P+~-norms must tend to infinity as n ~ c~, for every e > 0. Then we attach such spaces Y to one another to obtain an infinite set that is precisely of type A p. In [5] an infinite set of this sort was constructed by considering the spaces span{z 2n,..., z 2n+l } on ~' in the role of X; the "attachment" procedure consisted in applying the LittlewoodPaley decomposition. This is easy. It is the above finite-dimensional statement that is really difficult. Bourgain showed that, in fact, translation invariance is irrelevant in it.
Banach spaces and classical harmonic analysis
885
THEOREM 14 (see [5]). Let q91 ..... q3n be mutually orthogonal functions on a probability space(S2, v), let lqgil <. l for all i, and let 2 < p < cx~. Then there is a subset S C {1 . . . . . n} of cardinality at least n 2/p satisfying
)1/2 i~Esai q)i
<~ C (p)
Z
lai 12
iES
for all scalar sequences {ai }. The constant C (p) depends only on p. In fact, in a due probability sense, most subsets S of {1 . . . . . n} possess the required property 9 Later, Talagrand gave a different proof of this remarkable theorem (see [36]). Talagrand's proof is somewhat simpler than Bourgain's, but it leans upon finer probabilistic background. Also, there is some more philosophy of Banach space geometric nature around Talagrand's proof. We quote an intermediate statement occurring in [36]. In the case of X = LP it is weaker than Theorem 14; however, it is applicable in a more general setting. The norm of a Banach space X is said to be 2-smooth if IIx + y II + Ilx - y II ~< 2 + C Ily II2 with C independent of x and y. It is known that for p ~> 2 the norm of L P is 2-smooth. Now, let {xi } 1<~i<~n be a collection of vectors in a real Banach space X, and let
r -- sup[ Z f(xi)2 9f E X*, Ilfll ~< 1 }. l<~i<~n Talagrand showed that if the norm of X is 2-smooth and m = [n 1 - e / r ] (e E (0, 1) is an arbitrary fixed number), then there exists a subset I of cardinality rn in { 1 . . . . . n } such that
i~Elai Xi
K
Z
iEI
lai
) 1/2 [2
(In fact, "most" subsets I of cardinality m are such.) Here K depends only on e and on the constant C in the definition of 2-smoothness. We observe that if X = L P, then r <. n 1-Z/p, so that the statement is "e-close" to Theorem 14. The reader is referred to [36] for the proof of the above result, and also for the procedure of eliminating e in the case of the LP-norm, p > 2. By the way, the method generalizes to some other norms (an analog of Theorem 14 for the Lorentz space L p,1 was presented in [36]). We do not have the possibility of entering into the technicalities of either Bourgain's or Talagrand's proof.
14. Sidon sets A subset E of F is called a Sidon set if Y~• ] f ( g ) ] ~< c]]f]]~ for every f E L ~E (G). The smallest constant c is called the Sidon constant of E and is denoted by S(E). Again,
S. V. Kislyakov
886
the most well-known example is given by the Hadamard lacunary sequences (for G = qF). Also, the coordinate functions of the dyadic group (viewed as characters) constitute a Sidon set. Basically, the latter sequence is none other than the sequence of Rademacher functions on the segment [0, 1]. It is well known that every Sidon set E is a Ap-set for all p 6 [ 1, ~ ) and, moreover, Ilfllp ~< c~fffllfll2
for f ~ L 2 ( G ) , p ~ 1;
(4)
see, e.g., [24]. A deep result of Pisier [29] shows that, in fact, (4) is a characterization of Sidon sets. We shall discuss Sidon sets in more detail later on. Here we mention only a recent result of Kalton and Pdczyfiski [19] who showed that if E C F is a Sidon set, then L l \ E ( G ) is not an El-space (the question about this had been circulating for some time before that). In fact, again, the group structure has turned out to be irrelevant here. THEOREM 15. I f IZ is a finite measure and Q is a surjection of L l(#) onto a space conmining a copy of co, then X = Ker Q is not complemented in X** and the Grothendieck theorem fails f o r X (in the sense that there is an operator from X to 12 that is not 1summing). Consequently, X is not an s and is not isomorphic to a Banach lattice. See [ 19] for the proof and related results.
15. G o r d o n - L e w i s local unconditional structure
We refer the reader to "Basic concepts" for the precise definition, recalling only that a Banach space X possesses Gordon-Lewis local unconditional structure (G-L 1.u.st.) if and only if X* is a complemented subspace of a Banach lattice. Harmonic analysis provides many examples of spaces without local unconditional structure. THEOREM 16 (Kwapiefi and Petczyfiski [23]). If E is a quasi-Marcinkiewicz set and there is a bounded function r ~ 12(E) such that the multiplier T~o maps L1E to L 2, then CE(G) fails to have G-L l.u.st. THEOREM 17 (Pisier). If E is a A2-set, then E is a Sidon set if and only if CE(G) has G-L l.u.st. THEOREM 18 (Pisier). Let E be a A2-set, and let p > 2. Then E is a Ap-set if and only if L PE(G) has G-L 1.u. st. REMARK. The "only if" parts of Theorems 17 and 18 are trivial because C E ~ l l(E) for every Sidon set and L p ~ 12(E) for every Ap-set if p > 2. The proofs will be given in the next section, along the lines of the paper [23]. Here we make some comments. In Theorem 16, let G = qF, E = Z+. Then E is a Marcinkiewicz
Banach spaces and classical harmonic analysis
887
set. As q), we may take either X{2 n. n)0} (see Theorem 3), orthe function n w-> (n + 1) -1/2, n e Z+ (then the continuity of T~0"H 1 --+ H 2 is a consequence of the classical Hardy inequality y~n>~o(n + 1 ) - l l f ( n ) l ~< cllflll, f e H 1" see [16]). So, we recover the wellknown fact that the disk algebra fails to have G-L 1.u.st. Also (compare with Section 9), the proof of Theorem 16 can be adapted to show that the space c(l)(qI 'k) does not have G-L 1.u.st. if k ~> 2 and l ~> 1. Clearly, it suffices to prove this only for 1 -- 1, k - 2. Then the role of T~0 in Theorem 16 can be played by the Sobolev embedding operator W~ 1) (qI'2) ~ LZ(T2). In Section 9 it was already explained that C(1)(ql"2) is similar to a space of the form CE(G) with E a quasi-Marcinkiewicz set. The adjustment of the details is left to the reader (or see [23]). Passing to applications of Theorem 17, we take E -- {2n" n ~> 0} C Z (this is a Sidon set for qr) and consider its square E l - - { ( 2 k,2l)" k , l ~ > 0 } C Z 2. From condition (4) for E it is easily seen that E1 is a A p-set for •2 for every p < oc. However, it is also easy to observe that in the analog of (4) for E1 the constant grows as p, i.e., faster than ~/-fi. Thus, E~ is not a Sidon set. By Theorem 17, CEI (q,2) fails to have G-L 1.u.st. (we note that CE~ (~2) is isomorphic to the projective tensor product of 11 by itself). The domain of applicability of Theorem 18 is outlined in Section 13. We refer the reader to [23] for more examples.
16. Theorems 16, 17, and 18 are proved by similar methods. For Theorem 16, we consider the following operators:
CE(G) id>
L~(G) T~>L2E(G) ~> CE(G).
Here ~p is an arbitrary function in 12(E). By Theorem 12, the adjoint T~ is 1-summing. If CE(G) has G-L 1.u.st., the identity embedding id factors through L 1 (because id is 1summing; see [10]). So, we have come across a composition of operators belonging to mutually adjoint operator ideals, hence it follows that
vj (T~oid T~) <~cllZ~011 II~IIz2(E) (vl stands for the nuclear norm), or
I~(y)ll~o(y)l ~ C'II~IIz2(E). FeE
This contradicts the condition q) r 12(E).
E]
888
S. V. Kislyakov
For Theorem 17, we consider the following operators"
CE(G)
i d > L1E(G)
k > L2E(G)
T> CE(G).
(5)
Here k is the formal identity; k is continuous because E is a A2-set. Next, T is the convolution with a function x ~ C E(G). We claim that T* is 1-summing. Indeed, T* is an operator of convolution with x ( - t ) ; convolving this function with a measure, we obtain a function in C - E ( G ) , which does not depend on a particular choice of this measure in a fixed coset modulo CE(G) • Hence, T* acts as follows:
T* 9CE(G)*
(.),x(-t)
> C-E(G)
idl > LI_E(G)
kl > L 2 E ( G ) ,
where, as in (5), idl is the identity embedding and kl is a formal identity; thus, T* is 1-summing. Now, in (5) id is 1-summing, so that, if CE(G) has G-L 1.u.st., as in the preceding proof we obtain
v l ( k i d T ) <~ CIIxll~,
E CE(G),
x
or
I~(y)l ~ CIIxll~,
x E CE(G).
vcE
Thus, E is a Sidon set. For Theorem 18, we let q-1 + p-1 _ 1 and f ~ L q (G). Then, regarded as a mapping from L P ( G ) to L Z ( G ) , the operator S of convolution with f factors as follows:
S 9LPE(G)
r > L ~E (G)
id > L1E(G)
k > L2(G).
Here r is again the operator of convolution with f , and k is the formal identity 9 We see that S is 1-summing. If LPE(G) has G-L 1.u.st., then S factors through L 1(v) for some measure v. Next, for p > 2 every operator from a subspace of L p to L 1(v) factors through a Hilbert space H (see [31, Chapter 3]). This implies the following factorization for S"
S.LP(G )
u > H
v > LI(v)
w > L2(G).
By the Grothendieck theorem, w is 1-summing. Hence, wv is Hilbert-Schmidt, with Hilbert-Schmidt norm not exceeding Cllf]lq. Since (wv)* is also Hilbert-Schmidt with
Banach spaces and classical harmonic analysis
889
the same norm, it follows that
) 1/2 IIS* (z)ll 2
<<.Cllfllq,
f E L q (G),
yc-E
or, equivalently,
( ~~' If(y)12) 1/2
<. CIIfllq,
f E Lq(G).
ycE
A simple duality argument shows that the latter condition is equivalent to the fact that E is a A p-Set. E]
17. Quasi-Cohensets v s quasi-Marcinkiewiczsets It turns out that in Theorem 16 the condition of the existence of q) admits a nice reformulation. See the end of Section 2 for the definition of a quasi-Cohen set. THEOREM 19 (Kwapieti and Petczyfiski [23]). The following properties of a set E C F are equivalent. (i) E is a quasi-Cohen set. (ii) For every multiplier T~'L1E(G) -+ L 2 ( G ) we have r ~/2(E). (iii) There is a constant K such that for every trigonometric polynomial p satisfying fi(y) ~ Ofor y E E, we have ~-'~• fi(Y) <<-KIIPlI~. PROOF. (i) =, (ii). Let # be a measure satisfying/2(?,) = 0 for y r E, IZ%(Y)I ~ 1 for y E E, and let S be the operator of convolution with #. Then T~S maps L 1(G) to L 2 ( G ) , i.e.,
(
),/2f
ISl,
s~
y~E
Letting f run through some approximate identity, we infer that
(
(
) 1/2
C/> Z If6(Y)12lqg(Y)12 /> Z Iq)(Y)[2 yEE
y~E
(ii) =:~ (iii). L e t / 3 ( y ) / > 0 for y e E. We introduce the multiplier
S. V. Kislyakov
890
To estimate the norm of this operator, for f 6 L1 we write
~_. "---
"(y)[) 2 -- f]__f (-x)(p , f)(x)dx <~IIflll
Ilf
,J (_i
gEE
pilot
~< Ilfll2llpll~,
hence l I T / ~ 1 1 ~ Ilpll 1/2. From (ii) we deduce that (iii) =~ (i). In the space
w {P
C(G),
C(G)"
E
~Z~• P(Y) <~CIIPlI~.
we introduce two convex sets:
/3(y)/> 0 for y E E, ~ / 3 ( y )
/ -- K | , J
ycE
U -- {ql nt- q2" Ilqlll~ < 1, 02(?/) ~< 0 for y ~ E}.
Here K is the constant occurring in statement (iii). Clearly, W is closed and U is open. Also, W A U = 0. Indeed, if p E W and p = ql + q2 as in the definition of U, then (/3 - qz)lE ~> 0 and K - ~/3(y)
~ ~
y~E
(/3(9/) - 02(y)) ~
flip -
q211~ -
Xllq~ I1~
< K.
TeE
Thus, there is a (signed) measure # on G such that
< K
f o r p E U,
d/z(x) ~> K
f o r p 6 W.
Re
fG p(x)d/z(x)
Re
fa p(x)
We show that the measure v defined by v(A) = / z ( - A ) y 6 E, then K y 6 W, hence we see that
has the desired properties. First, if
K <~RefGKyd#--KRe~z(-y), i.e., I~(y)l ~ 1. Second, if y ~ E and e is any complex number, then p = c a u s e / 3 0 0 = 0 for X E E). Thus, K > Re K (ef,(y)), and ~(g) = 0.
Key ~ U
(beV]
Now we can restate Theorem 16 as follows: if E is a quasi-Marcinkiewicz set but not a quasi-Cohen set, then C E (G) fails to have G-L 1.u.st.
Banach spaces and classical harmonic analysis
891
18. S i d o n sets and a r i t h m e t i c d i a m e t e r
For a finite set A C F , its arithmetic diameter d(A) is defined as the smallest N such that CA (G) is at most 2-distant (relative to the Banach-Mazur distance) from a subspace of the N-dimensional space l~c. If E is a Sidon set in F , then CA (G) is at most S(E)-distant f r o m / i l l for every finite A C E. Consequently, the arithmetic diameter of A must grow exponentially as a function of IAI. There are many ways to see this. For instance, we may argue as follows. First, recall that for p ~> 2 the type 2 constant of L P is of order c~/-fi. Next, it is easily seen that the BanachMazur distance between l~c and 1~gN is bounded uniformly in N, so the type 2 constant of l~c does not exceed c~/log N. Finally, if el . . . . . en are the coordinate unit vectors in In1, then the Rademacher average f01 II ~ ri(t)ei IIdt is equal to n, so that the type 2 constant of I1 is at least ,c/ft. Thus, if 11 is 2-embeddable in l~c, then ~ ~< C'~/log N, as desired. It is remarkable that the exponential dependence of d (A) on IAI in fact characterizes the Sidon sets. THEOREM 20 (Bourgain [4]). I f l o g d ( A ) >~6lAIfor everyfinite A C E, then E is a Sidon set. Moreover, we have S(E) >~c6 -11, where c is a universal constant. The proof of this Banach-geometric characterization of Sidon sets can hardly be called "geometric"; rather, it is combinatorial, because the only geometric notion involved is that of entropy. In general, in a metric space with metric p, the e-entropy of a set F is the logarithm of the smallest cardinality Np (e) = Np (F, e) of an e-net for F. The idea of using entropy in the theory of Sidon sets was originally exploited by Pisier (see, e.g., [30]). Later, Bourgain revised Pisier's work on Sidon sets, replacing some fine probability methods by an elementary random choice combined with entropy combinatorics and with harmonic analysis arguments. We refer the reader to the Bourgain's survey [4] (and to the references therein) for this. The following statement is [4, Corollary 8]; this is a slight improvement of Pisier's original entropy characterization of Sidon sets (see [30, p. 941 ]). For a subset A of F, we define a metric pA o n G as follows:
pA(x, y) = sup Iv(x) -- y ( y ) [ . yEA
THEOREM 21. Let E C F. l f for every finite set A C E and some r > 0 we have
NpA (G, r) > 2 ~IAI, then E is a Sidon set and S(E) <~ Co--l~
2 r)
10
"
We show how Theorem 20 is deduced from Theorem 21. Suppose that E C F is such that l o g d ( A ) / > 3[AI for all finite sets A C E. For any finite A, along with pA we introduce another (greater) metric x A on G:
)cA(x, y ) - s u p { I f ( x )
- f(y)l
f e CA(G), Ilfll~
~ 1}.
S. V. Kislyakov
892
It is easily seen that for the arithmetic diameter d(A) we have d(A) <, NxA (G, 1/3). Indeed, denoting the latter quantity by N, we find a (1/3)-net Xl . . . . . XN for G in the metric x A . Then the mapping f w-~ ( f (x l) . . . . . f (XN)) is a good embedding of CA (G) into 1~. Next, fixing a finite set A C E, we clearly have
.~B(X, y) <~S(A)p B (x, y) for all B C A, yielding
Np8
G, 3S(A)
>~Nx8 G,
>, d(B) >~2 ~IBI.
Now, Theorem 21 implies the inequality
S(A) <~C6-1~ which does not allow the quantity S(A) to grow infinitely as A expands. More precisely, we obtain S(A) <. Ct~ -11 for every finite A C E, whence S(E) <. Ct6 -11. [:]
19. Sidon sets and cotype THEOREM 22 (Bourgain and Milman; see [8,4]). If CA (G) is of cotype q for some q < cx~,
then A is a Sidon set. Conversely, if A is a Sidon set, then CA (G) ~ l 1(A), which is of cotype 2. Recalling the well-known characterization of the spaces with finite cotype, we can restate Theorem 22 in the following way: either A is a Sidon set, or CA contains the spaces l ~ uniformly. Theorem 22 can be deduced from Theorem 20 and the following subtle fact of the theory of finite-dimensional Banach spaces. LEMMA 23. Let X be an n-dimensional normed space 2-isomorphic to a subspace of l~. Then n <. C[Cq(X) logCz(X)] q logN. Here Cr (X), 2 ~< r < cx~, is the cotype r constant of X. We refer to [4] for the proof of this lemma. Now, we prove Theorem 22. Let E C F, and let C E(F) be of cotype q, i.e., Cq(CE(G)) < oo. We fix a finite set A C E and obtain an a priori estimate for S(A) (like that at the end of the preceding section). If B C A, then C2(CB(G)) <<,S(A). By Lemma 23 IB[ ~< C[Cq(CE(G))]q[logS(A)] q logd(B),
Banach spaces and classical harmonic analysis
893
where, as before, d (B) stands for the arithmetic diameter of B. Now, Theorem 20 implies that
S(A) <~C'[Cq (CE(G)) log S(A)] 11q, which does not allow the numbers S(A) to be unbounded. Thus, E is a Sidon set.
[J
It should be mentioned that the case of q -- 2 in Theorem 22 is less involved (this case had been analyzed independently by Pisier and by Kwapiefi and Petczyfiski prior to Theorem 22). We can argue nearly as in the proof of Theorem 18. Indeed, let CE(G) be of cotype 2. For a function x ~ CE(G), the operator r of convolution with x ( - t ) can be viewed, in a natural way, as a mapping from CE(G)* to C-E(G). So, the domain of r is the conjugate of a cotype 2 space, and its range is a cotype 2 space. By Pisier's factorization theorem (see [31, Chapter 3]), r factors through a Hilbert space H. This factorization r = riot is shown in the following diagram: j~<
L2E(G) *
> CE(G)*
r
> C-E(G)
k > L2_E(G)
H Here j and k are identity embeddings. Now, k is Considering the adjoints, in a similar way we quently, the composition kfi@* is nuclear with sition is again the operator of convolution with
1-summing, hence kfl is Hilbert-Schmidt. see that orj* is Hilbert-Schmidt. Consevl (kfi@*) <~Cllxll~. Since this compox ( - t ) , we obtain Y~'~• I~(y)l ~< CIIxll. D
20. Multipliers on spaces of vector-valued functions Let X be a Banach space" then the spaces of X-valued functions LP (G" X) (1 ~< p ~< cx~) or C(G, X) are defined as usual. As in the scalar case, for E C F the spaces L p (G; X) and CE (G; X) are distinguished by the condition
f(Y) de__f~ ffi _ 0
f o r v ~ E.
Now, assume we are given a bounded multiplier of scalar spaces Tm" L pl El (G) --+
LP2 E2(G), acting in accordance with formula (1). The expression m 9f makes sense also for X-valued functions f , so it is natural to ask about the description of the spaces X for which (1) generates a bounded operator from L p~EI(G,X) to L P2Ez(G,X). In fact, for each particular multiplier Tm this question presents a mystery, which can be clarified only rarely, and by dissimilar techniques. Below we briefly discuss two important
894
S.V. Kislyakov
cases. Beyond this, we mention the paper [3], in which the vector-valued analogs of the Paley inequality (see Theorem 3) and the Hardy inequality ~(n n>~O
+ 1 ) - l l f ( n ) l ~< CIIflll,
f E H 1(qF),
as well as some related questions were discussed. In spite of a quite considerable bulk of information presented in [3], no complete description of the corresponding classes of Banach spaces X is available in these cases as yet.
21. B-convexity and K-convexity Here our basic multiplier is the orthogonal projection of L2(D) (D is the dyadic group, see Section 8 for the definition) onto the subspace generated by the coordinate functions of D. (Identifying D with [0, 1] in the usual way, we arrive at the Rademacher projection.) The spaces X for which this projection extends in a natural way to LZ(D; X) are said to be K-convex. The notion of K-convexity is quite useful in the theory of Banach spaces, primarily due to the fact that it is intimately related to the duality between type and cotype. Remarkably, K-convex spaces admit a complete characterization. THEOREM 24 (Pisier; see [32]). A Banach space X is K-convex if and only if X does not contain the spaces lln uniformly. The spaces with the latter property are called B-convex. A space is B-convex if and only if it is of nontrivial type; see "Basic concepts".
22. UMD-spaces Here our basic multiplier is the Hilbert transformation 7-[ on L 2 (T). We have 7-[ = Tm, where m(k) = - i sgnk, k E Z. It is really quite useful to know for which Banach spaces X the Hilbert transformation (more generally, an arbitrary Calder6n-Zygmund singular integral operator) acts on the LP-space of X-valued functions. The general theory reduces the case of any p E (1, cx~) to the case of p = 2; see, e.g., [35]. Surprisingly, the class of spaces X for which 7-/acts on L2(T, X) admits a complete description. THEOREM 25 (Burkholder, McConnel, Bourgain). A Banach space X has the above property if and only if there is a biconvex function ~ : X x X --+ IR such that ~ (0, O) > 0 and (x, y) ~< IIx + y II if IIx II - IlY II - 1.
Banach spaces and classical harmonic analysis
895
The same condition is equivalent to the continuity of standard martingale transformations on LZ(x). That is why the spaces X such that 7-/acts on L2(X) are called UMDspaces (UMD is for "unconditionality of martingale differences"). The proof of Theorem 25 is indirect and passes via this statement on martingales. We refer the reader to Burkholder's survey [9] and to the references in it for the details. It should be noted that the condition formulated in Theorem 25 is difficult to work with. Basically, the only way to prove the existence of the above ~ on a Banach space X is to verify the continuity of 7-/(or of the martingale transformations) on L 2 ( X ) directly. The Hilbert space seems to be the only one presenting a simple possibility of exhibiting ~ (for instance, we may put ((x, y) = 1 + Re(x, y)). Direct verification of the continuity of 7-/shows that the spaces L P with 1 < p < cx~ and the Shatten-von Neuman classes Cp (again with 1 < p < oc) are UMD-spaces. Next, the property of being a UMD-space is inherited by the subspaces and quotient spaces. This is a superproperty (if X is finitely representable in Y and Y is UMD, then so is X). A UMDspace must be superreflexive, but not all superreflexive spaces are UMD. As before, we refer the reader to [9] and the references therein. Again, the proofs of the above statements do not use Theorem 25.
23. The above discussion should be supplemented with the following result due to Bourgain. THEOREM 26. I f tx is a finite measure and 0 < r < 1, then the Hilbert transformation 7-[ is bounded f r o m L2('1F; L 1(Ix)) to L2(~I'; L r (Ix)). We refer the reader to the survey [20] for the proof and related material. The result can be used to verify statement (ii) in Section 8. This idea is Bourgain's; the details of this verification can also be found in [20].
24. Returning once again to statements (i)-(v) in Section 8, we note that we may ask any Banach space theory question about any specific space arising in harmonic analysis. This will yield an incontestable point of contact of the two fields, but rarely will this show a real interplay between them. In many cases, a pure problem of hard analysis (even without the adjective "harmonic") arises in this way, as is described in Section 8 after statements
(i)-(v). Some exceptions of these "rule" were, however, discussed above. Another one is presented by the still mysterious space U of uniformly convergent Fourier series on the unit circle. Bourgain was the first to prove that U* is weakly sequentially complete, and his proof involved difficult techniques of hard analysis; see [6]. But later it was discovered that the statement can be verified almost entirely within the ("soft") methods of functional analysis. We refer the reader to the paper [11 ] in this collection for a discussion and references.
S. V. Kislyakov
896
Having mentioned the space U, we probably cannot avoid considering the spaces of trigonometric polynomials on the circle. Let 7Jnp -- span{l, z
. . . . .
s
with the metric of L p (Tf), 1 <<,p <<,cx~. In a way (and from the harmonic analysis viewpoint), these spaces may be regarded as "elementary (n + 1)-dimensional blocks" building the Hardy classes H p (T). For 1 < p < oo this is emphasized by the fact that the 7anp are c o m p l e m e n t e d in HP uniformly in n. However, for p -- 1 or p -- oo the norm of the invariant projection of H p onto 7)np grows as c log n (and, by averaging, the norm of any other projection cannot be smaller). The following observation (folklore) was made by Bourgain and Petczyfiski about 20 years ago. PROPOSITION 27. The spaces 7)1n (respectively, T~n~ ) can be embedded uniformly complementedly in H 1 (respectively, H ~ ) . PROOF. We treat only the case of H l, the other one being similar. Considering the subspaces generated by the odd or by the even powers of z, it is easy to deduce that H 1 ~ H 1 • H 1. Thus, instead of H l, we embed 7~l into ~ 1 The embedding in question is given by the formula
In'p~--~ (~n+lp, p),
G H 1 -- L 1_ (~') G L l + (~').
pETgln.
Now, we define an operator Jn "L l_ G Ll+ ~ p l by the formula
Jn ( f , g)
= z n+l
Kn * f + Kn * g,
where Kn is the nth Fej6r kernel. It is easily seen that Jn In -- idpl. Without entering into the details, we refer the reader to [7] for a construction of a good basis in 7)n~ , and to [ 12] for evaluation of various constants (such as G-L 1.u.st., B a n a c h Mazur distances to various spaces, etc.) for the spaces 7)p. Let us stop at this point.
References [1] D.E. Alspach and A. Matheson, Projections onto translation-invariant subspaces of L 1(R), Trans. Amer. Math. Soc. 277 (2) (1983), 815-823. [2] G.E Bachelis and S.E. Ebenstein, On A(p) sets, Pacific J. Math. 54 (1) (1974), 35-38. [3] O. Blasco and A. Pelczyfiski, Theorems of Hardy and Paleyfor vector-valued analyticfunctions and related classes of Banach spaces, Trans. Amer. Math. Soc. 323 (1) (1991), 335-367. [4] J. Bourgain, Subspaces of L~, arithmetical diameter, and Sidon sets, Probability in Banach Spaces, V, Medford, MA, Lecture Notes in Math. 1153, Berlin, Springer (1984). [5] J. Bourgain, Bounded orthogonal sets and the A(p) set problem, Acta Math. 162 (1989), 227-246.
Banach spaces and classical harmonic analysis
897
[6] J. Bourgain, Quelques propriYt~s lin~aires topologiques de l'espace des s~ries de Fourier uniformement convergentes, S6minaire Initiation d'Analyse, G. Choquet, M. Rogalski and J. Saint Raymond, eds, 22e ann6e, Expos6 no. 14, Univ. Paris-6 (1982-83). [7] J. Bourgain, Homogeneous polynomials on the ball and polynomial bases, Israel J. Math. 68 (3) (1989), 327-347. [8] J. Bourgain and V. Milman, Dichotomie du cotype pour les espaces invariants, C. R. Acad. Sci. Paris, S6r. 1300 (9) (1985), 263-266. [9] D.L. Burkholder, Martingales and Fourier analysis in Banach spaces, Lecture Notes Math. 1206, Springer, Berlin (1986), 61-108. [ 10] J. Diestel, H. Jarchow and A. Tonge, Absolutely Summing Operators, Cambridge University Press (1995). [11] T.W. Gamelin and S.V. Kislyakov, Uniform algebras as Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 671-706. [ 12] Y. Gordon and S. Reisner, Some geometrical properties of Banach spaces of polynomials, Israel J. Math. 42 (1-2) (1982), 99-116. [13] C.C. Graham and O.C. McGehee, Essays in Commutative Harmonic Analysis, Springer, Berlin (1979). [14] F.E Greenleaf, Invariant Means on Topological Groups and their Applications, Van Nostrand-Reinhold Company, New York (1969). [15] R.F. Gundy, In~galit~s pour martingales ~ un et deux indices, Lecture Notes Math. 774 (1980), 251-334. [16] K. Hoffman, Banach Spaces of Analytic Functions, Prentice-Hall, Englewood Cliffs, NJ (1962). [ 17] J.E Kahane and R. Salem, Ensembles Parfaits et S~ries Trigonom~triques, Hermann, Paris (1963). [18] M.I. Kadets and A. Petczyfiski, Bases, lacunary sequences and complemented subspaces in the spaces L P, Studia Math. 21 (1961/62), 161-176. [19] N.J. Kalton and A. Pdczyfiski, Kernels of surjections from 121-spaces with an application to Sidon sets, Math. Ann. 309 (1) (1997), 135-158. [20] S.V. Kislyakov, Absolutely summing operators on the disk algebra, St. Petersburg Math. J. 3 (4) (1992). [21] S.V. Kislyakov, Some more Banach spaces satisfying Grothendieck's theorem, St. Petersburg Math. J. 7 (1) (1996). [22] I. Klemes, Idempotent multipliers of H 1, Canadian J. Math. 39 (5) (1987), 1223-1234. [23] S. Kwapiefi and A. Pdczyfiski, Absolutely summing operators and translation invariant spaces of functions on compact abelian groups, Math. Nachr. 94 (1980), 303-340. [24] J.M. Lopez and K.A. Ross, Sidon Sets, Lecture Notes in Pure and Appl. Math. 13, Marcel Dekker, New York (1975). [25] B. Maurey, Th~orkmes de factorization pour les op~rateurs lin~aires ~ valeurs dans les espaces L p, Ast6risque 11 (1974). [26] V.D. Milman and G. Schechtman, Asymptotic Theory of Finite Dimensional Normed Spaces, Lecture Notes Math. 1200, Springer-Verlag, Berlin (1986). [27] A. Pdczyfiski, p-integral operators commuting with group representations and examples of quasi pintegral operators which are not p-integral, Studia Math. 33 (1969), 63-70. [28] A. Pdczyfiski, Linear extensions, linear averagings, and their applications to linear topological classification of spaces of continuous functions, Dissertationes Mathematicae (Rozprawy Matematyczne) LVIII, PWN Warszawa (1968). [29] G. Pisier, De nouvelles caract~risations des ensembles de Sidon, Advances in Math., Supplementary Studies, Math. Analysis and Appl. (part B) 7, L. Nachbin, ed. (1981). [30] G. Pisier, Conditions d'entropie et caract~risations arithm~tiques des ensembles de Sidon, Topics in Modern Harmonic Analysis, Proc. Seminar held in Torino and Milano, May-June 1982, Vol. II, Instituto Nazionale di Alta Matematica Francesco Severy, Roma (1983). [31] G. Pisier, Factorization of Linear Operators and Geometry of Banach Spaces, Reg. Conf. Ser. in Math. no. 60, Amer. Math. Soc., Providence, RI (1986). [32] G. Pisier, Holomorphic semi-groups and the geometry of Banach spaces, Ann. Math. 115 (1982), 375-392. [33] H.E Rosenthal, Projections onto translation-invariant subspaces of LP (G), Mem. Amer. Math. Soc. 63 (1966). [34] W. Rudin, Trigonometric series with gaps, Math. Mech. 9 (1960), 203-227. [35] E.M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press (1970).
898
S.V. Kislyakov
[36] M. Talagrand, Sections ofsmooth convex bodies via majorizing measures, Acta Math. 175 (2) (1995), 273300. [37] Q. Xu, Some properties of the quotient space L 1(qFn)/H 1(Dn), Illinois J. Math. 37 (3) (1993), 437--454. [38] A. Zygmund, Trigonometric Series, Vols I-II, Cambridge University Press (1959).
CHAPTER
21
Aspects of the Isometric Theory of Banach Spaces Alexander
Koldobsky*
Department of Mathematics, University of Missouri, Columbia, MO, USA E-mail: [email protected]
Hermann K6nig Department of Mathematics, University of Kiel, Kiel, Germany E-mail: hkoenig @math. uni-kiel,de
Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Extension of isometries in Lp-spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Positive definite functions and isometric embedding of normed spaces in L p-spaces . . . . . . . . . . . 4. The Busemann-Petty problem on sections of convex bodies . . . . . . . . . . . . . . . . . . . . . . . . 5. Approximation of zonoids by zonotopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Exact estimates for projection constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
901 902 906 918 924 928 935
*The first author was supported in part by the NSF Grants DMS-9531594 and DMS-9996431, and by the UTSA Faculty Research Award. H A N D B O O K OF THE GEOMETRY OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 899
This Page Intentionally Left Blank
Aspects of the isometric theory of Banach spaces
901
1. Introduction
The isometric theory of Banach spaces was born and developed in inseparable connection with other areas of the Banach space theory. Many articles of this Handbook contain isometric and almost isometric results and problems. A remarkable result of Ball [4] on the maximal volume of hyperplane sections of the unit cube in R n, Burkholder's martingale inequalities with exact constants [15], exact probabilistic estimates are just a few examples. Isometric and almost isometric operators appear in numerous situations. In spite of this close connection with the rest of the Banach space theory, the isometric theory has its specific features and methods. The goal of this article is to show how these methods work in different settings and to emphasize connections between the isometric theory and other areas of mathematics, such as Fourier analysis, geometry, probability and combinatorics. The classical direction, initiated by the work of Banach in the early 30s, is the characterization of the isometries of Banach spaces. Many people have contributed to this direction, and we refer the reader to a comprehensive survey [25] for numerous results and references. Surjective isometries of Banach spaces are now well understood in a long sequence of results culminating in complete characterization of such isometries for all rearrangement invariant spaces by Zaidenberg [130] (complex case) and Kalton and Randrianantoanina [49] (real and complex cases). The characterization of injective isometries is far from being complete, though recent results by Carothers, Haydon and Lin [17] (for Lorentz spaces) and Randrianantoanina [114] (Orlicz spaces) can be considered as a successful start. Another important goal of this direction of the isometric theory is to generalize to other classes of spaces the extension method for Lp-isometries discovered in the 70s by Plotkin [105-109] and, independently, Rudin [117] and Hardin [40]. In Section 2, we describe this method and its connections with Fourier analysis, potential theory and probability. In Section 3, we present some other results that belong to the intersection of the isometric theory, harmonic analysis, probability and combinatorics. The connections between positive definite functions, stable measures and isometric embeddings of Banach spaces into Lp were discovered in the 30s by P. L6vy. Since then, these connections have been under intensive investigation. We discuss recent developments, in particular, a solution to Schoenberg's problem on positive definite functions and its generalizations. This section also includes recent results explaining the difference in the isometric structure of the spaces Lp and lp, and showing the special place of the spaces Lp with p being an even integer. We explain recent results of Delbaen, Jarchow and Pe~czyfiski [ 18] when subspaces X of Lp imbed isometrically into lp or Ip, U. the results differ for p E 2N and p r 2N. N with small In the case of X - l~ and p c 2N, we study concrete imbeddings into Ip, N -- N(n, p). Recently discovered connections between convexity and the Fourier transform are subject of Section 4 and are explained there through a complete analytic solution to the Busemann-Petty problem on section of convex bodies, which was considered one of the most important isometric questions in convexity. An almost isometric counterpart of the problem of imbedding of subspaces of L p into Ip studied in Section 3 is considered in Section 5 for p = 1: Given e > 0, when can finite
902
A. Koldobsky and H. Krnig
dimensional spaces X __cL1 be (1 + e)-isomorphically imbedded into 1N, N = N ( X , e)? This is the dual setting of the problem of approximating zonoids by zonotopes which was studied by various authors, e.g., Bourgain, Lindenstrauss, Milman, Matou~ek, and Talagrand. The best known estimates for N ( X , e) as a function of e are given in Section 5. In Section 6 we derive precise upper bounds for the projection constants of general finite dimensional Banach spaces which improve the square root of dimension estimate of Kadec-Snobar. In the case of symmetric spaces better bounds hold. The examples, which show that these estimates are almost exact, are related to the techniques used to imbed l~ into 1u .
2. Extension of isometries in Lp-spaces
The extension method for L p-isometries provides an effective tool for characterization of the isometries of subspaces of L p, and it also has different applications to other areas of functional analysis, probability and harmonic analysis. This method is based on three principles: the Uniqueness theorem for measures, the Equimeasurability theorem and the Extension theorem. The first versions of these principles appeared in early 70s in the papers of Plotkin [105-108], and in 1976 all three results were published in their final form in [ 109]. The Uniqueness and Equimeasurability theorems were discovered independently by Rudin [ 117] in 1976. The Extension theorem was proved independently by Hardin [40] in 1981. We start with a simple proof of the Uniqueness theorem. THEOREM 1 (Uniqueness theorem). Let p > O, p q~ 2N, C E R, # a n d v are finite Borel m e a s u r e s on R so that, f o r every a E R,
fRl
x - al p d/z(x) + C = fR Ix -- al p dr(x) < cx:~.
(1)
Then lz -- v a n d C = O.
PROOF. Let 4~ E S(R) be a test function supported outside of the origin. Then f a q~(t) d t = 2yr4~(0) --0. Since p is not an even integer, for every fixed x E R the Fourier transform (in the sense of distributions) of the function a w-~ Ix - al p is equal to (Ix - a l P ) A ( t ) = C p l t 1 - 1 - p exp(-itx), where Cp = ( 2 p + l y r l / 2 F ( ( p + 1 ) / 2 ) ) / F ( - p / 2 ) (see [31, p. 173]). Using this fact and Fubini's theorem, we get
= Cp f R I t l - l - P f t ( t ) q b ( t ) d t .
(2)
Aspects of the isometric theory of Banach spaces
903
Integrating both sides of (1) with respect to ~(a) da and using (2), we get
s
ltl-l-P fz(t)~(t) dt -- s
Itl-l-P~(t)r
dt
for every test function 4~ supported outside of the origin. Therefore, continuous functions and ~ are equal everywhere on R, which implies that # - v. [] If p is an even integer the Fourier transform formula used in the proof does not hold. In this case, the statement of the Uniqueness theorem is no longer true, because the equalities (1) imply only that a fnite number of moments of the measures # and v are equal. The proofs of the Equimeasurability and Extension theorems below are essentially the same as in [109]. Recall that the distribution of a function f ~ Lp(f2, 0") is the measure # o n R defined by # ( A ) = 0-{co E Y2: f(co) E A} for every Borel set A C R. THEOREM 2 (Equimeasurability theorem). Let p > O, p q~ 2N, (121, al) and (.('22, 0"2) be probability spaces, H be a subspace of L p(S-2t, 0"1) containing the constant function 1(o9) = 1, and T : H w+ Lp(S-22, 0"2) be a linear isometry such that T1 = 1. Then every function f E H and its image T f have equal distributions with respect to the measures 0"1 and 0"2, correspondingly. PROOF. Let/z and v be the distributions of f and T f . Since T 1 -- 1, for every a E R
fs2 If(~
- alp do-1 ( c o ) - fs? [T f ( c o ) - a P da2 (co).
1
2
Making the changes of variables x -- f (co) and x - T f (co) in both integrals, we get (1). By the Uniqueness theorem, # -- v. [2 In order to prove the Extension theorem, we need the following LEMMA 3. We remain under the conditions of the Equimeasurability theorem, except for
T 1 -- 1 which is no longer required. Then for every function f ~ H, the support of T f is contained in the support of T 1. Besides, for any k ~ N the joint distribution of any functions f l . . . . . fk ~ H with respect to the measure (71 is equal to the joint distribution of the functions T f l / T 1 . . . . . T f k / T 1 with respect to the measure [T 1 [P d0-2. PROOF. Let f c H. Then, for every a 6 R
fs2 lf (co)
a Ip do-1 (co)
1
-
[ T f (co) - a T 1 (co) [P da2 (co) 2
I
l(co)--/:0} ~
(co) -- a
I IT
1 (CO)[P da2 (co) +
f,w
1(~o)=0}[ T f (co) [P da2 (co).
A. Koldobsky and H. K6nig
904
If we denote the second summand in the right-hand side by C, and pass to the distributions of the functions f and T f ~ T 1 with respect to the measures do'1 and iT 11p do'z, correspondingly, we find ourselves in the situation of the Uniqueness theorem. Hence C = 0, and the function T f is equal to zero o'z-almost everywhere on {T 1 (co) -- 0}, which means that the support of T f i s contained in the support of T 1. Now the operator T ' H w-> Lp.~S-22, ITllPdo'2) defined by T f T f / T 1 is a linear isometry satisfying the condition T 1 -- 1. Denote by # and v the measures on R ~, which are the joint distributions from the statement of lemma. By the Equimeasurability theorem, for every b - (bl . . . . . bk) E R ~, the functions ~ bi fi and ~ bi Tfi / T1 have equal distributions with respect to o'1 and IT 11p do'z, respectively. Therefore, the Fourier transforms of the measures # and v are equal at the point b. Since b is arbitrary, the result follows. D THEOREM 4 (Extension theorem). Under the conditions of the Equimeasurability theorem, except f o r T 1 -- 1 which is no longer required, there exists an injective linear isometry T f" Lp(S-21, .fit, o'l) ~ Lp(S-22, o2) so that TtII4 - T, where ..4 is the smallest o'-algebra of subsets o f 1-21 making all functions from H measurable. PROOF. It is easily seen that the space Lp(S21, A, o'l) is the closure in Lp(S'21, o'l) of the set of functions that can be represented in the form B (fl . . . . . f~), where k 6 N, B is a Borel function on R k and fl . . . . . fk belong to H. Define an operator T ~" Lp (~21, A, o'l) w-~ Lp (S-22, o'2) by
T ' ( B ( f l . . . . . f~)) -- T 1 . B ( T f l / T 1 . . . . . T f k / T 1 ) for every choice of k, B, f l . . . . . fk. The operator T ' is well-defined. In fact, suppose that B1 (fl . . . . . fk) - - B2 (gl . . . . . gm) for k, m E N, Borel functions B1, B2 o n R k and R m, and functions f l . . . . . f~:, g l . . . . . gm ~ H. By Lemma 3 applied to the functions
f l . . . . . fk, gl . . . . . gm,
fs?
l B l ( f l . . . . . f ~ ) - B2(gl . . . . . gm)] p do'1
1
=
]BI(Tfl/T1 ..... Tf~/T1)-
B 2 ( T g l / T 1 . . . . . Tgm/T1)IPITIIPdo'2.
2
Therefore,
T 1 . B I ( T f l / T 1 . . . . . T f ~ / T 1 ) -- T 1 . B 2 ( T g l / T 1 . . . . . T g m / T 1 ) as elements of Lp(.Q2, o'2). The same argument (with B2 - - 0 ) shows that T' is a linear isometry, and its restriction to H is equal to T. D The Extension theorem applies only to subspaces of L p containing constant functions (unital subspaces). However, the general case can be reduced to the unital case by changing the density, since every separable subspace of L p contains a function with maximal support (see, for example, [3]).
Aspects of the isometric theory of Banach spaces
905
Let us show a typical application of the Extension theorem. Let E be a domain in R n. Denote by Lph (E) the subspace of L p ( E ) (with Lebesgue measure) which consists of harmonic functions. The following result was proved in [108]. THEOREM 5. Let p > 0, p r 2N, n/> 2, E1 and E2 be two domains in R n. The spaces L hp(El) and Lph (E2) can be isometric only in one of the following situations: (i) the domains E1 and E2 are similar; (ii) p = 2 n / ( n - 2) and the domains E1 and E2 coincide up to the composition of an inversion and a similarity. PROOF (Sketch). The smallest o--algebra A, making all functions from L hp (El) measurable, is the a-algebra of all Borel subsets of El. By the Extension theorem, every linear isometry T from Lph (El) to Lph (E2) can be extended to an isometry T I from the whole space Lp(E1) to Lp(E2). Using the classical characterization of the isometries between L p-spaces (see [72]), one can show that T' is a composition operator, i.e., T'f(co) = T l ( c o ) f ( r ( w ) ) , where r is a measurable mapping from E2 to the closure of El. Now our problem is reduced to characterization of all composition operators mapping harmonic functions to harmonic functions. Direct differentiation shows that this happens only if r is a conformal mapping. Now one can apply Liouville's theorem that every conformal mapping in R n , n ~> 3, is the composition of a similarity and an inversion. It can be shown that an inversion can be included only when p - 2 n / ( n - 2). In the case n = 2, the conformal mapping r is a holomorphic or an antiholomorphic function, and the fact that r is a similarity follows from a calculation based on the fact that T 1 -- Ir'l 1/p must be a harmonic function. D The Uniqueness theorem was generalized in [37,80] to the case of measures on certain finite and infinite dimensional normed spaces. Suppose that p > 0, E is a separable Banach space, and #, v are finite Borel measures on E so that, for every a c E,
g(a) -- fE IIx -- allP d # ( x ) -- fE IIx - allP dv(x) < cx3.
(3)
We say that p is an exceptional exponent for E if (3) does not necessarily imply that # = v. The exceptional exponents for the space Lq, q > 0 are those for which p / q is an integer, and, in the case of the n-dimensional space l q, n besides the condition p / q E N one of the following must be satisfied: (i) p / q < n, (ii) q is an even integer, (iii) q and p / q - n are both odd integers. If E -- C (K), where K is a compact without isolated points, then there are no exceptional exponents. For the complex space l ~ exceptional exponents are even integers, and in the real case p is exceptional if and only if n + p is an odd integer. Other results in this direction include formulae for calculating the measure # out of the potential g (see [53]), uniqueness theorems for Gaussian measures (see [75,81,82]). Applications and generalizations of the extension method also include [2,18] (see Section 3), [30,41,42,52,56,78,79,85-87,89,95,113,116,122,124,134].
A. Koldobsky and H. KSnig
906
3. Positive definite functions and isometric embedding of normed spaces in Lp-spaces The problem of how to check whether a given Banach space is isometric to a subspace of Lp was posed by Lgvy [74] in 1937. A well-known fact is that a Banach space embeds isometrically in a Hilbert space if and only if its norm satisfies the parallelogram law [26, 47]. Neyman [103] (see also [126] and [15]) proved that subspaces of Lp with p ~ 2 cannot be characterized by a finite number of equations or inequalities. There is a close connection between isometric embeddings in L p and positive definite functions. Recall that a complex valued function f defined on R n is called positive definite if, for every finite sequence {xi }im_1 in R n and every choice of complex numbers {ci }iml, m
E
m
ECiCjf(xi
-- Xj) ~/O.
i=1 j--1 By Bochner's theorem, every continuous positive definite function on R n is the Fourier transform of a positive finite measure. It was known already to L6vy [74] that if B = (R n, I[. II) embeds isometrically in Lp, 0 < p ~< 2, then exp(-l[xl[ p) is a positive definite function, and, hence, is the Fourier transform of a symmetric p-stable random vector. The actual equivalence of the two notions was discovered by Bretagnolle, Dacunha-Castelle and Krivine [13] who proved that, for 0 < p ~< 2, a Banach space B is isometric to a subspace of Lp if and only if the function exp(-llxll p) is positive definite. For different proofs and generalizations see [13,70,1,101,46,127] and [7, Chapter 11]. Bretagnolle, Dacunha-Castelle and Krivine used their result to prove that the space Lq embeds isometrically in L p when 0 < p < q ~< 2. However, it is more difficult to check whether exp(-[Ix I[p) is positive definite for other norms. For example, the following problem posed by Schoenberg in 1938 (see [121]) was open for more than fifty years: for which p e (0, 2) is the function exp(-llx I[p) positive definite, where [Ix [[q is the norm of the space l q,n 2 < q ~< oo .9 An equivalent formulation asks whether the spaces l qn embed in Lp with 0 < p ~< 2. After Dor [21] answered the question for p e [1, 2), the complete solution (including p e (0, 1)) was given in [51] for 2 < q < oo and in [99] for q = oo: if n ~> 3 the spaces l q,n 2 < q <<.oo do not embed isometrically in L p, 0 < p < 2. For n = 2 the embedding exists if and only if 0 < p ~< 1. The fact that every two-dimensional normed space embeds in Lp for every p e (0, 1] was established much earlier (see [23,43, 76]). We now present the solution to the case 2 < q < oo of Schoenberg's problem from [51 ]. Parts of this solution will also be used in Section 4. The solution is based on the following Fourier transform characterization of subspaces of Lp (see [54,57] for details). In [20] this criterion was also applied to Lorentz spaces. THEOREM 6. Let p > 0, p ~ 2N and let (R n, [l" I[) be an n-dimensional subspace of Lp. Then the Fourier transform of the function F ( - p / 2 ) [ I x lip is a positive distribution on
Rn \ {0}.
907
Aspects of the isometric theory of Banach spaces
PROOF. We have to prove that, for every non-negative even test function 4) 9 ,S(R n) supported in R/7 \ {0}, (r(-p/2)(llxllP) A, 4~) ~> 0. A simple fact going back to L6vy [74] is that an n-dimensional normed space (R n, II 9 II) embeds isometrically in L p, p > 0, if and only if there exists a finite Borel measure/z on the unit sphere S n- 1 in R n so that, for every x 9 R/7, Ilxllp = fs,-~ I(x, ~)1 p d#(~). Therefore,
,
-
~(x) dx)du(~)
= (2~)n--|CP fS~_, fll Itl-l-pd#(t~)dtd#(~)>~O' where Cp is the constant from the proof of the Uniqueness theorem in Section 2, and we use the fact that (27r)nq~(t~) is the Fourier transform of the function t w-~ f(x,~)=t ~(x) dx. Since r ( - p / 2 ) C p ~ O, the result follows. Fq n 2 < q < cx~. Denote Let IlXllq -- (Ix]l q + ' - " + Ixnlq) 1/q be the norm of the s p a c e lq, by Vq the Fourier transform of the function z --+ exp(-Izlq), z 9 R. The properties of the functions 7q were studied by Polya [110]. In particular, if q is not an even integer, the function 7q(t) behaves at infinity like Itl - q - 1 . Namely (see [111, Part 3, Problem 154]),
lim tl+qyq(t)
-
-
21-'(q + 1) sin(zrq/2).
t---+ (x3
If q is an even integer, the function Vq decreases exponentially at infinity. The integral
Sq (or) - fR Itl~ yq (t) dt converges absolutely for every ot 9 ( - 1 , q). These moments can easily be calculated (see [135]; ot is not an even integer):
Sq(c~) - 2~+2rcl/ZI'(-a/q)I'((a + 1 ) / 2 ) / ( q r ( - a / 2 ) ) . Clearly, the moment Sq (or) is positive if ot 9 ( - 1 , 0 ) U (0, 2), and the moment is negative if ot 9 (2, min(q, 4)).
LEMMA 7. Let q > O, n 9 N, - n < # < qn, #/q ~ NU {0}, ~ 1 <.k <.n. Then /7
(llxll )A( )
-
q Ji~ tn+#-l 1-i y q ( t ~ ) d t . 1-,(-/3/q) k=l
=
(~1 ....
,
~n) 9 R n, ~k r 0,
A. Koldobsky and H. Kgnig
908
PROOF. Assume that - 1 < fl < 0. By the definition of the F-function (Ixll q §
§ IxnlP) ~/q
-Y - 1 - ~ e x p ( _ y q (IXl [q + ' " + l x n l q ) ) d y . _ F ( - fql / q ) fo ~176 For every fixed y > 0, the Fourier transform of the function x --+ exp(-yq(lxllq + ... + [xn[q)) at any point ~ 6 R n is equal to y-n I-I~=l y q ( ~ / y ) . Making the change of variables t - 1/y we get
(([Xl[ q § 2 4 7
[xnlq)fl/q)A(~)
q
-- F ( - f l / q )
I ' ( - fql / q )
fo
y-n-fl-1
fo
Un Yq(~k/Y) dy k=l
/7 U yq(t~)dt. k=l
(4)
The latter integral converges i f - n 3 < qn since the f u n c t i o n H ~ - - 1 yq(t~k) decreases at infinity like t -n-nq (recall that ~k # 0 for every k). If/3 is allowed to assume complex values then both sides of (4) are analytic distributions of/3 in the domain { - n < Re/3 < nq, fl/q ~ N U {0}}. These two functions admit unique analytic continuation from the interval ( - 1 , 0). Thus the equality (4) remains valid for all fl E ( - n , qn), fl/q ~ N U {0} (see [31] for details of analytic continuation in such situations). D LEMMA 8. Let q > 2. If n > 3, p 6 (--n + 3, 0), or n -- 3, p 6 (0, 2), or n = 2, p E (1, 2), then the Fourier transform of the distribution IIx IIqp is a sign-changing function on Sn-1. PROOF. By Lemma 7 and properties of the moments Sq (O/), the integral I(O/1 . . . . . O/n-l) = .fl/ [~l [Oil " " " [~n-I [Otn-1 ([[X [[P) A (~1 . . . . . ~ n - l , 1) d~l""" d~n-I = Sq(O/1)'' " S q ( O / n - 1 ) S q ( - O / 1
.....
O/n-1 - P)
converges absolutely if the numbers O/1 . . . . . O/n-l,--19/1 . . . . . O/n-1 + p belong to the interval ( - 1, q). Choosing O/~ 6 ( - 1, 0) for every k = 1 . . . . . n - 1, we have the moments Sq (o/h), k = 1 . . . . . n - 1, positive, and we can make -o/1 . . . . . O/n-1 + P equal to any number from (p, n - 1 + p) A ( - 1, q). Because of our conditions on p, this interval contains a neighborhood of 2, and, since the moment function Sq changes its sign at 2, we can make the integral I (O/1. . . . . O/n-l) positive for one choice of O/'s and negative for another choice. This means that the (homogeneous of degree - n - p) function (llx IlqP)A is sign-changing. []
Aspects of the isometric theory of Banach spaces
909
Theorem 6 and the cases n -- 3 and n - 2 of L e m m a 8 imply that the spaces l qn
with
n ~> 3, 2 < q < ec cannot embed in L p, 0 < p ~< 2, and that the spaces l q2 with 2 < q < cx~ can not embed in Lp with 1 < p < 2, which answers Schoenberg's question. Soon after the paper [51] appeared, Zastavnyi [128,129] and Lisitsky [83] independently proved a more general result that there are no non-trivial positive definite functions of the form f([] 9Ilq), where q > 2, n ~> 3. Note that for q = oc a similar result was established earlier in [99]. Zastavnyi's result is even stronger: THEOREM 9. Let E be a three-dimensional normed space with a basis el, e2, e3 such that the function x w-> Ilxel + ye2 + ze31] is differentiable at the point x = 1 for almost all (y, z) 9 R 2, and suppose that the function (y,z) ~
]]xel + ye2 +ze3llxt (1 , y, z)/llel -4- ye2 + ze31]
y, z 9 R,
belongs to the space L1 (R2). Assume that, for some function f : R ~-+ R, the function (x, y, z) ~ f (]lxel Jr- ye2 + ze3 I1) is positive definite on R 3 . Then f is constant. PROOF (Sketch). First, one can assume without loss of generality that f is differentiable on (0, ec), limt--,otf'(t) = O, and there exists a constant C > 0 so that Itft(t)l < C for every t > O. In fact, if this is not the case, one can replace f by a function F(t) = f o f ( t s ) g ( s ) d s , where g is an infinitely differentiable non-negative function with compact support in R. It is easy to check that the function
h(y,z)--
(1 -]y])(1 -Izl),
O,
max{lYl, Izl} ~ 1, otherwise
is positive definite on R 2. Let us prove that
~p(t) - fR2 f (lltel + ye2 nt- ze311)h(y,z)dydz is positive definite on R. In fact, for every e > 0
~/e(t, y , z ) -
f ( l l t e l + ye2 + ze311)h(y,z)exp(-elt])
is integrable and positive definite on R 3 (as the product of positive definite functions). Therefore, for every s E R
~ ( s , 0, 0) - fR exp(--ist) e x p ( - - e l t l ) ~ ( t ) dt >>.O, which means that the function e x p ( - e l t l ) g r ( t ) is positive definite on R. Since e is arbitrary, 7z is positive definite on R.
910
A. Koldobsky and H. K6nig
Using the fact that the derivative of the norm by the first coordinate is a homogeneous function of degree 0 and making the change of variables y = tu, z = tv, we compute the derivative of the function 7t at a point t ~ 0: /,
~r'(t) - I f ' ( l l t e l + ye2 + ze311)lltel + ye2 + ze311't(t, y , z ) h ( y , z ) d y d z JR 2 t2sign(t)
f.2
f'(Itlllel + ue2 + ve311)lltel + ue2 + re3 II't(1 , u, v)
• h(tu, tv) du dr. Multiplying and dividing the expression under the latter integral by Ilel + ue2 -+- re3 II, we can use the condition of the theorem and the properties of the function f mentioned in the beginning of the proof to show that limt~0 ~P!( t ) / t = 0. (Note that this is possible because we have t 2 in front of the integral, which is due to the fact that the dimension of the space is three. This is the point of the proof that fails in dimension 2.) This condition on the derivative of ap implies that ~p!(0) = ~p!!(0) = 0. It follows now from [87, Theorem 4.1.1] that ~p is a constant function. The same is true if we replace the function h(y, z) by the function n2h(ny, nz), where n 6 N. As n --+ cx~, the fact that the corresponding functions ~ are constant immediately implies that f is constant. [] Clearly, every normed space of dimension ~> 3, which is the q-sum with q > 2 of two non-trivial normed spaces, satisfies the conditions of Theorem 9. Another necessary condition for isometric embedding in L p, 0 < p ~< 2, was given in [62]. The proof in [62] is a modification of Zastavnyi's argument. THEOREM 10. Let X be a three-dimensional normed space with a normalized basis {el, e2, e3 } so that: For every fixed (x2, x3) E R2\{0}, the function Xl --+ Ilxlel + x2e2 + x3e311 has a continuous second derivative everywhere on R, and f!
Ilxll' (0 x2, x3)= IlXllx2(O, x2, x 3 ) - 0 X 1
'
' and IIx II'x'2 stand f o r thefirst and second derivatives by Xl, respectively. where Ilxllxl 1
There exists a constant K so that f o r every Xl E R and every (x2, x3) E R 2 with IIx2e2 + " (Xl , x2, x3)<~K x3e311 - 1, one has IlXllx~ Then, f o r every 0 < p <~2, the space X does not embed isometrically in Lp. Theorem 10 applies to some spaces not covered by Zastavnyi's result, for example, to Orlicz spaces l~ whose Orlicz function M has continuous second derivative and M'(0) = M" (0) = 0 . In 1981, Eaton [22] introduced a concept of an isotropic random vector generalizing stable vectors in the following way: a random vector X = (X1 . . . . . Xn) is called isotropic if there exists a quasi-norm I1" II on R n so that, for every a E R n, the random variables
Aspects of the isometric theory of Banach spaces
911
ai Xi and [la [IX1 have equal distributions. A simple fact is that a random vector X is isotropic if and only if its characteristic functional (Fourier transform) has the form f (11. II), where I1" II is a quasi-norm on R, and f : R w-~ R is a continuous function. In view of Bochner's theorem, the problem of characterization of isotropic vectors reduces to the study of positive definite functions of the form f(ll" II). On the other hand, if f(ll" II) is a positive definite function on R n, then f is positive definite on R, and is the characteristic function of a finite measure on R. It is easy to show that if this measure has finite moment of order p for some p > 0 then the space (R n, I1" II) embeds isometrically in Lp. These results represent another connection between probability, positive definite functions and isometric embeddings in L p. The classes of positive definite functions of the form f (11 9 II) have been studied by several authors. We refer the reader to surveys [57,100] for proofs of the results mentioned above and for numerous references. The following question is probably the most important open problem in this direction: is it true that a non-constant positive definite function of the form f (11" II) exists only if the corresponding quasi-normed space embeds isometrically in one of the spaces L p , 0 < p < 2? Isometric embeddings of finite dimensional normed spaces in L1 play important role in convex geometry, due to the fact that the unit ball of every such space is a polar to a zonoid (see Section 5 for more on zonoids). Surveys [120,35] include different characterizations of zonoids and related results. We would like to mention the following problem of Lindenstrauss and Schneider. Let d2(X) denote the Banach-Mazur distance between X and the Hilbert space of the same dimension. For each n, consider the number an -- sup{d2 (X)" X and X* embed isometrically in L1, dim X -- n }.
Is it true that lim S U P n ~ an = 1 ? Note that Schneider [118] gave the first example of an n-dimensional normed space X such that both X and X* embed isometrically in L1 but X is not isometric to l~. The spaces of Schneider are smooth small perturbations of the euclidean unit ball. This result of Schneider answered negatively a question raised by the work of Grothendieck [38] (see also Bolker [8]). Answering a further related question, Lonke [84] showed that in dimensions 3 and 4 there exist non-smooth spaces X with the same property. Schneider's construction was modified in [55] to show that there exist non-Hilbertian normed spaces of any finite dimension, which are isometric to subspaces of L p with p < 2 and of Lq with q > 2, simultaneously. Note that the question of whether every infinite dimensional space satisfying this property is isometric to a Hilbert space is still open, as well as a similar question for infinite dimensional spaces X such that X and X* both imbed isometrically in L1. In the following, L p denotes the space L p (0, 1) with respect to Lebesgue measure. If X is a subspace of L p, a change of density argument shows that there is a subspace Y of L p isometric to X and containing constant functions (unital subspace). For X C L p it is a natural question whether X embeds isometrically into l p or not. This is of interest even for X - l~. A very satisfactory answer to this problem, which depends on whether p is an even integer (p E 2N) or not, was recently given by Delbaen, Jarchow and Pe]czyfiski [18]:
A. Koldobsky and H. KOnig
912
THEOREM 1 1. Let 0 < p < c~ and X be a closed subspace of Lp. (i) If p q~ 2N, X is isometric to a subspace of lp if and only if every (equivalently,
some) unital subspace Y C Lp isometric to X consists of functions with discrete distribution. (ii) If p E 2N and X is finite dimensional X embeds isometrically in lp. For n E N and p = 2s E 2N, let NR(n, p)"--
n+p-
1)
-- 1,
N c ( n ' P ) ' - - ( n + s - 12 ) s
P
-1.
Then, more precisely, every n-dimensional subspace of Lp embeds isometrically into lpN where N <<,NK(n, p ) f o r K E {R, C}. Here the distribution )~f of a function f is discrete if )~f -- y~jENaj(~xj, where Xj, aj E K, aj >/O, y~jENaj -- 1. Following [18], we present an I D E A OF THE PROOF OF (i). Let p ~ 2 N .
(a) Assume Y C Lp is a unital subspace isometric to a subspace of lp, via an isometry w ' Y w-~ lp. Let y = w(1), S "-- supp(g). Then w(Y) can be regarded as a subspace of lpS -- {~ ~ lp ~s --0 for all s ~ S}. The formula v(B) "-- ~sEB IV(S)] p for B C S defines a probability measure on S since v(S) = I I w ( 1 ) l l p - 1. Moreover, define s
u'lp ~ Lp(v),
~ ~ (~(s)sgn(g(s))/y(S))se s.
This is an isometric isomorphism with u ( y ) = 1. Thus uw : Y ~ Lp(v) is an isometric isomorphism onto the unital subspace uw(Y) of Lp(v) with uw(1) -- 1. Take f E Y. If f = zl for z 6 K, f has discrete distribution. If not, u w maps the 2-dimensional subspace [ 1, f ] C Lp onto [1, uw(f)] C Lp(v) isometrically. By the Equimeasurability theorem (see Section 2), the distribution ~f of the function f with respect to Lebesgue measure is equal to the distribution of u w ( f ) with respect to v, which is discrete. Therefore, ~f is discrete. (b) If Y C Lp has finite dimension and f l . . . . . fk is a basis of functions having discrete distribution, we may write for n = 1 . . . . . k
fn -- Z Zj,nXC(j,n), jEJ(n)
(C(j, rt))jEJ(n) a partition of [0, 1] into Borel sets. These sets generate a purely atomic ~r-algebra generating a space isometric to Ipm for some m 6 N U {ec} containing Y. For dim Y - co, more measure theory is needed, see [1 8] for details. 71 PROOF OF (ii) FOR K - R. Let p E 2N, p - 2s. For a Banach space Z, let BMn(Z) be the set of isometry classes of n-dimensional Banach spaces [E] generated by the ndimensional subspaces E ___Z with respect to the Banach-Mazur metric. Let n E N and N -- NR(n, p) as in the claim (ii). It suffices to show that there is a subset An C BMn(Lp), dense with respect to the Banach-Mazur metric, such that
Aspects of the isometric theory of Banach spaces
913
An C BMn(IN). Continuity of the embedding and compactness of BMn(1 N) then yield BMn(Lp) C BMn(lNp), i.e., the statement we want to prove 9 ForrEN,
l e t l k ,r . _ [ k-r-l ,
ur "Lp ~ Lp,
k The averaging operator ?-].
fl---~~r(flk k=l
f ( t ) dt)Xlk,r ,r
has norm 1 and Ur(1) = 1. Let An,r be the family generated by the n-dimensional subspaces of ur(Lp) C Lp, and put An :-- UrcN An,r. Then An is dense in BMn(Lp). We will show that An C BMn (l N). The important fact is that N is independent of r. Since Ur(1) -- 1, if Y C Lp is unital, so is ur(Y). We thus start with a unital subspace Y C ur(Lp) of dimension n. Since ur(Lp) C Loc, Y has a basis {f0, f l . . . . . fn-1 } with f0 -- 1 and [Ifj 11~ = 1 for j = 1 . . . . . n - 1. Moreover, the functions in Y are constant on small intervals by our reduction to ur(Lp). Let
nl n-1
S'--
oteZ+
}
9 l~<Eotj~
ot--(otj)j=
nm
1 .
j=l
Then I S I x E R n-1
,
NR(n, p) = N. We use standard multiindex notation: for o t ~ Z~_-1 and 9Put F" __ ( f j ) jn--11 _ Then for all (bj)j-~ C R n
n x Ol.__ Ej=-ll Xj j
bjfj
.
-
bjfj
Lp
j=0
c~m~,
d)~ - co + orES
j=0
where the ca are polynomials in the coefficients (b j) and mc~ "--f01 F c~d)~. Let m -
(ma)acs E R u. We will show that m can be also obtained by integration of similar functions v c~ on a discrete set of size N. Let I = [ - 1 , 1] and L "-- {(x~)~cs" x - - ( x j ) j - l E I n - 1 1 C R N. Put x (k) "-- ( f j ( ( k - 1/2)/r))j-11 for k - 1 . . . . . r. Then (x(k)) ~ ~ L since Since the f j ' s are constant on the small intervals Ik,r,
m -- F
II/jll~
- 1.
E C " - conv,_,.(L) k=l
By (an improvement of) Caratheodory's theorem, m is a convex combination of at most N elements of L C R u (see [18]). Thus there are ~(d) 6 L and Vd ~ 0 (for d -- 1 , . . . , N) with ~ f f - - 1 V d - 1 such that m -
~ f f = l Vd~ (d) 9 Let x (d) - - t~"x j(d)~n-1 be the ( u n i q u e ) p o i n t Jj=l
of I n-1 such that ((x(d))~)~S -- ~(d) E L. Let .N" -- { 1 . . . . . N} and
vj "A/" w+ R,
_ (d)
d E ./V"w+ z j
(j -- 0 . . . . . n -
1)
A. Koldobsky and H. KSnig
914
and v0 = 1. Also, v = (Vd)ff=l defines a probability measure on A/'. Let Z-
span[v0,131
.....
Vn-1] C
N Lp(v) =~lp.
Then
Y w-~ Z,
f j ~ vj
(j - - 0 . . . . . n - 1)
extends to an isometric isomorphism: note that with V - (Vj.j=)n--11
(N
)
d-1
N
-- ~
orES
1)d~(d) = m -- (ma)~ES,
d=l
and hence
~bjvj j=0
=
Lp(v)
... j=0
r
dv
--
co z S c. +
--co+Zcama--co+Zc ~ orES Hence Y is isometric to a subspace of l pN
V c~
dv
a6S
orES
fo
Fad~:
~bjfj j=O
Lp D
In the case of embedding l~ into l N, the idea of using Caratheodory's theorem was employed by Milman [96]. The general method was already used by Kemperman [50]. COROLLARY 12. Let 0 < p < c~, p q~ 2N. Then X is isometric to a subspace of lp if and only if every 2-dimensional subspace of X is isometric to a subspace of lp. This is immediate from (i). Since no infinite dimensional subspace of lp is isomorphic to 12 for p # 2, we get COROLLARY 13. Let 0 < p < cx~, p ~ 2N. Then the 2-dimensional Hilbert space 12 is not isometric to a subspace of lp. This was a question of Pietsch motivating a part of [ 18]. It was known before that 12 does not embed isometrically into l pN i f p ~ 2 N , cf.[91]. For p - 2s E 2N, let MK(n, p) denote the minimal value of N such that (ii) of the theorem holds; thus MK(n, p) <, NK(n, p). For n -- 2, p -- 4, (ii) gives MR(2, 4) ~< 4, Mc(2, 4) ~< 8. The following beautiful result in [18] is very surprising: PROPOSITION 14. Every two-dimensional subspace Y of L4 is isometric to a subspace of 134 if K - R and of 14 if K - C. In the real case, such a subspace admits a 1-symmetric basis 9
Aspects of the isometric theory of Banach spaces
In the case of the two-dimensional Hilbert space Y = l 2 C L4 and K give an explicit embedding into 143 since
B2--
{
915 R it is easy to
3]
x--(xj)~__ 1" I l x l l 4 - 1 , Z x j - O j=l
is a circle (of radius 21/4). Of course, for all n 6 N and s 6 N, l~ embeds isometrically into Lp for p -- 2s. By HK(n, p) we denote the minimal dimension N such that l~ embeds isometrically into lpN , p -- 2s. Hence by Theorem 1 l(ii), HK(n, p) <. MK(n, p) <. NK(n, p). Let 79horn --p,n denote the space of polynomials in n variables homogeneous of degree p -- 2s restricted to the sphere S n-1 , in the real case. In the complex case, we mean all polynomials q (z 1. . . . . Zn, Z 1 . . . . . ~ n ) which are homogeneous of degree s in each of the sets of variables (Zl . . . . . Zn) and (Ti-. . . . . ~nn), restricted to S n- 1(C). Let ~r denote the normalized surface measure on S n-1 . Hilbert's formula for x 6 K n , i s n-, I(x' y ) 12s dcr (y) = Cn,s IIx 112 2s
shows that cn ,s Ilxll2s is in the (closed) convex hull of the set {l(', Y)I 2s " Y E S n-1 } ~ "Dh~ --2s,n" Using again (the improvement of) Carath6odory's theorem, one gets again the upper bound
HK(n, p) <<.dim 7--p, ")h~n
-
-
1 --
NK(n , p) ,
p = 2s ,
cf. Milman [96] for this argument. A lower bound is known too [64]. We summarize these bounds in the P R O P O S I T I O N 15. For n, s ~ N and p -- 2s, let
LR(n, p) "--
n-k-s-I) s
LR(n, p) "-- ( n + [[(S+(s + 11)/2]--)/2] 1)(n+[s/2]--l)[s/2]
Then LK(n, p) <. HK(n, p) <. NK(n, p). The upper bounds for HK(n, p) and MK(n, p) being identical, the following conjecture is made in [ 18]. CONJECTURE 16. HK(n, p) = MK(n, p) for n, s E N, p = 2s. At least it is true for n = 2, p = 4. For fixed p = 2s and n ~ oc, LK(n, p) n p/2, NK(n, p) ~ n p, i.e., both bounds differ substantially. By Dvoretzky's theorem for lp,
916
A. Koldobsky and H. KOnig
cf. [24], 12n embeds (1 + e)-isomorphically into l pN , where N - - t e n p / 2 if p > 2, e > 0. If for p E 2N, c~ could be chosen to be bounded as e --+ 0, one would achieve the order of the lower bound, HK(n, p) ~ n p / 2 . For p -- 4, this is true, see Proposition 18 below. By results of Reznick [115], Goethals and Seidel [34], and Lyubich and Vaserstein [93], the existence of isometric embeddings of l~ into l~ is equivalent to the existence of certain cubature formulas on the sphere S n- 1 with N points; this would seem to suggest that the upper bound is reasonable: PROPOSITION 17. Let n, s, N E N. Then the following are equivalent: (1) There exists an isometric embedding of l~ into lU ; (2) There are N points (x~)N1 C S n-1 and a probability sequence (/zk)N_l C R+, ~--~
y ~ N 1 #~
h~ 1 , such that for all polynomials p E ,"D2s,n N
fs
n--I
(5)
~/~p(xk);
p(y)dy-
k=l
(3) There exist (xk) and (lz~) as in (2) such that N
Cn,s - l._l l._, I(x, Y)I 2s d o t ( x ) d o ' ( y ) Jo
J~
~
2s.
(6)
k,l=l
PROOF (Sketch [64]). (1) => (3) Any embedding of l~ into 1N has the form x w-~ ((x, z k ) ) ~ with z~ 6 K n. Let x~ - - z~/llz~ll2 and #k "-IlzkllZs/~-~'~Ul IIz~II2s. Then (3) holds. (3) ==>(2) Let K - R. For x E R n , we denote by x | 6 R nzs the (2s)-fold tensor product of copies of x. Then (x | y| _ (x, y)2S. Consider N
"-- Z
[J"kx~2S
-- fs n-1 x|
dcr (x) E Rn2s .
k=l
Then Sidelnikov's inequality holds,
N o
-
2s
m Crl,s ~ O~
k,l=l hom which implies that ~ = 0: all monomials of degree 2s and hence all p E ,, D 2s,n are integrated exactly as indicated in (5). (2) =~ (1) Let x 6 K n. Apply (2) to p ( y ) - I(Y, x)12s. Then N
E. l(x,xs l
I(x, y)l 2' d o ' ( y ) - Cn,sllXll2, n-1
k=l
i.e., x ~ ((lZk/Cn,s)l/2S(x,xs))~=l defines an isometry of l~ into l~.
917
Aspects of the isometric theory of Banach spaces
Condition (3) is useful to prove the existence of certain explicit embeddings since only one e q u a l i t y - (5) - has to be checked. The lower bound is exact, L K ( n , p) -- H K ( n , p) --: N, if and only if the spherical design (Xs)sU__l is tight, i.e., the set of scalar products C = {l(xk, xl)l: k ~ 1} is a very small set, and all #k'S are equal to 1 / N . For p = 4, I C I - 1, and the vectors (xk)~_-i span e q u i a n g u l a r lines. Known cases of equality are n(n + 1) HR (n , 4) -- - , 2
, n2 Hc(n 4)-,
n--2,3,7,23,
n--2,3,8.
For p = 4 it is not known whether there are more values of n such that this holds. For p >i 6, there are only very few cases of equality, the most spectacular being HR (24, 10) -- (28), cf. [115,93,64]. For n = 2 and p E 2N, Proposition 17 and Gaussian quadrature show that HR(2, p) -- p / 2 + 1. Thus, for 2 <~ N <~ p / 2 , 12 does not imbed isometrically into lpN . In this case, the distance to an (1 -+- e)-isomorphic/Z-section of l pU has been estimated by Lyubich and Shatalova [92]. To show that H K ( n , 4) is close to L K ( n , 4), one may study systems (xk) which are "almost" equiangular. PROPOSITION cases: (a) K = C, (b) K = C, (c) K = R, (d) K = R,
18 ([64,112]). The space l~ e m b e d s isometrically into 1u in the f o l l o w i n g n n n n
= = = =
q + l, q p r i m e power, N = n 2 + 1 . q, q o d d p r i m e power, N = n 2 + n. 2 m, m integer, N = n 2 + n. 4 m, m integer, N = n(n -k- 1)/2.
Hence HK(n, 4) is of the order of the lower bound. To show, e.g., (b), define for k = (kl, k2) E { 1 . . . . . n} 2 the vectors
xk " - ~
1
t {
2rri ( k l j + k 2 j 2) exp --n--
j=l
c S n-1 (C).
Then for I = (ll, 12) r k = (kl, k2),
{ 1/x/n ilk2~12, I(xk, xl)l --
0
if k2 -- 12.
In addition to these n 2 vectors one takes the n unit vectors el . . . . . en. These n 2 -+-n vectors satisfy (5) provided all/zk = 1 / N . Case (a) is done by using Bz-sequences and exponential vectors defined by them, case (d) follows by using vectors from the Kerdock code, cf. [64]. Case (c) uses a construction with Walsh functions, as observed by Schechtman and explained in Rabinovich [ 112]. For p = 2s/> 6, the problem is completely open. PROBLEM 19. Let s E N, s ~> 3, p = 2s. Is it true that for large n HK(n, p) ~ n p/29.
918
A. Koldobsky and H. K6nig
4. The B u s e m a n n - P e t t y problem on sections of convex bodies
The Busemann-Petty problem (see [ 16]) asks the following question. Suppose that K and L are origin-symmetric convex bodies in R n such that voln-| (K n ~•
~< voln-1 (L n ~_t_)
for every ~ from the unit sphere S n-1 in R n, where ~• = {x 6 Rn: (x, ~) = 0} is the central hyperplane perpendicular to ~. Does it follow that voln(K) ~< voln(L)? The answer is negative if n >~ 5 and affirmative for n ~< 4. The solution to the problem was based on the concept of an intersection body introduced by Lutwak [90] in 1988. A star body A in R n is called an intersection body of a star body B if for every ~ 6 S n-1 the radius of A in the direction of ~ is equal to the (n - 1)-dimensional volume of the section of B by the central hyperplane perpendicular to ~: pA(eS) --II~IIA ~ - voln-1 (B nse•
(7)
where II~IIA = min{t ~> 0: ~ ~ tA} is the Minkowski functional of A, PA is the radial function, and ~ • = {x ~ Rn: (x, ~) = 0}. A star body A is called an intersection body of a star body if there exists a star body B satisfying the equalities (7). The class of intersection bodies can be defined as the closure of the class of intersection bodies of star bodies in the radial metric d ( A , B) = max~esn-1 I,OA(~) - ,OB(~)l. Lutwak [90] found the following connection between intersection bodies and the Busemann-Petty problem (the original results of Lutwak were slightly improved by Gardner [27] and Zhang [131 ]): (i) If K is an intersection body then the answer to the Busemann-Petty problem is affirmative for every star body L; (ii) If L0 is an origin symmetric convex body that is not an intersection body, then one can perturb L0 twice so that the resulting bodies K and L are convex and give a counterexample to the Busemann-Petty problem. Therefore, THEOREM 20. The Busemann-Petty problem has an affirmative answer in R n if and only if every origin symmetric convex body in R n is an intersection body. The Busemann-Petty problem has a long history. The negative answer for n >/5 was established in a series of papers by Larman and Rogers [73] (for n ~> 12), Ball [5] (n 10), Giannopoulos [33] and Bourgain [9] (n >~ 7), Papadimitrakis [104] and Gardner [27] (n >~ 5). Gardner [28] proved that every symmetric convex body in R 3 is an intersection body, and, hence, the problem has an affirmative answer in the case n -- 3. For several years the answer in the dimension 4 was believed to be negative (see [132]), until it was shown n 2 < q <~ ~ , is an intersection in April 97 (see [59]) that the unit ball of the space lq, body if and only if n ~< 4. This result disproved the counterexample from [132] and is
Aspects of the isometric theory of Banach spaces
919
a consequence of the following connection between intersection bodies and the Fourier transform established in [58]: THEOREM 21. A star body K in R n is an intersection body if and only if II 9 IlK ~ is a positive definite distribution. Recall that a distribution f 6 S' (R n) is positive definite if and only if its Fourier transform f is a positive distribution in the sense that ( f , ~) ~> 0 for every non-negative test function q5 (see [32] for definitions and details). Another consequence of Theorem 21 (see [58]) is that the unit ball of every finite dimensional subspace of L p with 0 < p ~< 2 is an intersection body. Theorem 21 was also applied in [60] to present a variety of counterexamples to the Busemann-Petty problem in dimensions ~> 5. For example, if q > 2 the unit ball of the q-sum of any normed spaces X, dim(X) ~> 4 and Y is not an intersection body. After learning the results of [58] and [59], Zhang [133] proved that the answer in the four dimensional case is affirmative. Two months later, a unified solution to the BusemannPetty problem was given in [29]. The proof in [29] is based on THEOREM 22. Let K be an origin-symmetric star body in R n with C ~ boundary, and let k E N U {0}, k g= n - 1. Suppose that ~ E S n- l, and let A K , ~ ( t ) - - v o l n _ , ( K N {x 6 R n" ( x , ~ ) = t } ) ,
t6R,
be the parallel section function of K in the direction of ~. (a) If k is even, then
(llx IIK'~-+-k-+-~)~(~) -
( - - 1 ) k / 2 z r ( n -- k - 1)A K,~.(0); (k)
(b) If k is odd, then
(llxllKn+k+~)~(~) :
(--1)(k+l)/22(n -- 1 -- k)k! x
Z-k-1
AK,~ (Z) -- AK,~ (0) -- A"K,~ (0) .2! .... _
A (k-l)
z k-1 )
K,~ (0) ( k -
1)--------~ dz,
(k) where A K,~ (0) is the derivative of the order k of the parallel section function at zero, and (llX[I Kn+k+l) A is the Fourier transform in the sense of distributions.
Theorem 22 is a direct consequence of
A. Koldobsky and H. K6nig
920
Let K be an origin-symmetric infinitely smooth star body in R n . Suppose that ~ ~ S n-l, and let AK,~ be the correspondingparallel section function of K. For q ~ C with 9~q > -1, q ~ n - 1 we have THEOREM 23.
A K,~ (q) (0) --
COS q:rr Kn+q+l) A 2 ([IX[[ (~). n'(n -- q -- 1)
(q)
Here A K,~ (0) is the fractional derivative of order q at zero, defined by
A(q) K,~
(0) --
1
1-'(-q)
foC~t_q_lAK,~(t)dt
if - 1 < 9~q < 0, and by
A (q) (0) -l fo oe ( k-1 t2Jza(2J)(o)) dt' K,~ F(-q) t -q-1 AK,~(t) -- E (2j)!'*K,~ j=0 whenever 2k - 2 < 9lq < 2k, k E N. The function q w-~ A (q) K, ~ (0) , q ~ C , can then be extended to an analytic function on all of C. Note that the function A K,~ is even and that if q is an even integer the fractional derivative of order q coincides with the ordinary derivative of the same order. We refer the reader to [31, pp. 48-56] for details. To prove T h e o r e m 23, suppose first that - 1 < q < 0. The function AK,~(Z)= f(x,~)=z x ( l l x l l ) d x is even. Applying Fubini's theorem and passing to spherical coordinates, we get
A(q) x,~(o) =
1 F
2F(-q)
Izl -q-1AK,~
(z) dz
1 fRn [(X'~)l-q-1 x(llxllg)dx
= 2F(-q) __ 2F(-q) _
r n-q-2 x(rllOIIK) dr dO
,-1 1
2(n-q-
~
1)F(-q)
n-1
1(0 , ~) [-q-1 IlOIIgn§
dO.
(q) We now consider A K,~ (0) as a function of ~ ~ R n \ {0}. Using the same argument as in the end of the proof of T h e o r e m 6, we see that, for every even test function q9 ~ S
A K,~ q) (0) , qg(~)) =
1
2(n - q - 1 ) F ( - q )
fs
n-1
IlOIIK"+q+ldOfR I(O'~:)l -q-l~O(~)d~
-
-- 4(n - q - 1 ) F ( - q ) F ( q
+ 1) sin Y-~
~-1
IlOll~n+q+l
Itlq~(tO)dtdO
Aspects of the isometric theory of Banach spaces
921
qzr
(8)
~ ( n - q - 1)
((ll ll
Kn+q+l
)A
where the last equation follows from the property F ( - q ) I ' ( q + 1) = - J r / s i n ( q z r ) of the F-function and the simple calculation
- f.o -
f,
n-1
ilOil~+q +1
fo
tq~(tO)dtdO
(note that the function IIx [[-n+q-+-I is locally integrable on R n because - 1 < q < 0). Since (8) holds for every even test function qg, Theorem 23 is proved when - 1 < q < 0. To prove the theorem for other values of q, we first observe that (llxlI-Kn+q+l) A is an analytic distribution (with respect to q) on {q E C: 9~q > - 1 }. It follows that for every even test function q9 6 S, the functions q ~
q~--~
cos 2
(re(n-q-
1)
(llxll
K
(A (q) K,~(0), qg) and
,
)
(~), ~0
are analytic on the connected region {q 6 C: 9~q > - 1 , q ~ n - 1 } (for details of analytic continuation in such situations, see [8]). These functions coincide on the interval - 1 < q < 0, so they coincide on {q E C: 9]q > - 1 , q ~ n - 1 }. Since q9 is an arbitrary even test function, we have proved Theorem 23. Part (a) of Theorem 22 immediately follows from Theorem 23 and the fact that fractional derivatives coincide with ordinary derivatives. To prove part (b) of Theorem 22, divide both sides of the formula in Theorem 23 by cos(qTr/2), and compute the limit as q --+ k, where k is an odd integer. Another short proof of Theorem 22 (with the Fourier transform replaced by the spherical Radon transform) was recently given by Barthe, Fradelizi and Maurey [6]. Let us show how the solution to the Busemann-Petty problem follows from Theorems 4.1, 4.2 and 4.3. Let n = 4 and K be any symmetric infinitely smooth convex body in R 4. Put k = 2 in part (a) of Theorem 22. By the Brunn-Minkowski theorem (see [119]), the volume of the central section is maximal among volumes of sections perpendicular to a given direction, therefore, A ttx,~ (0) ~< 0 for every ~ By part (a) of Theorem 22, we conclude that Ilxll~.1 is positive definite, and, by Theorem 21, K is an intersection body. The positive answer to the Busemann-Petty problem in dimension 4 follows now from Theorem 20. If n = 5, we have to put k = 3 and positive definiteness of the function Ilx I1-1 depends on the properties of the third derivative of parallel section functions. Since convexity does not control the third derivative, it is easy to construct a counterexample using part (b) of Theorem 22. We now present a proof of a result from [61] that generalizes the solution to the Busemann-Petty problem. This proof no longer uses intersection bodies and is based on a version of Parseval's formula. The result is as follows:
922
A. Koldobsky and H. K6nig
THEOREM 24. Let K and L be (k - 1)-smooth origin symmetric convex bodies in R n such that, f o r every ~ ~ S n-1 , ( _ 1)(k-1)/2a(k-1) (k-l) "'X,~ (0) ~< (-1) (k- 1)/2AL, ~ (0),
(9)
where k is an odd integer and 1 <~ k <~ n - 1. Then (i) if k >~ n - 3 then voln(K) ~< voln(L); (ii) if k < n - 3 then it is still possible that voln (K) > voln (L).
Clearly, the case k = 1 of Theorem 24 is the answer to the Busemann-Petty problem. Also it is enough to prove Theorem 24 in the case where K and L are infinitely smooth. The crucial point of the proof is a version of Parseval's formula on the sphere that can be proved by extending functions to R n and using the classical Parseval's formula. LEMMA 25. Let K and D be origin symmetric star bodies with C ~ - b o u n d a r i e s in R n and k E N, k < n. Then
f n,
(11x IIg ~) A(0) (11XIIOn+k) A(0) dO -- (2yr)n fan_ 1 II0 IIKk II0 IIOn+k dO.
THEOREM 26. Let k be an odd integer, 1 <<,k <<,n - 1, and let K and L be origin symmetric (k - 1)-smooth star bodies in R n. Suppose that the distributions IIx I1~.k and
IlxllL n+k --IIx II~.n+~
are positive definite. Then voln (K) ~< voln(L).
PROOF. We have
fsn_ (llXllgk)A(o)(llXll2nq-k)A(O)dO ~ fsn_1(llXllKk)A(o)(llX]lKn-q-k)A(O)O0" By Lemma 25,
fs.-, II011~kll011z~+k dO ~ fs.-' II011zn dO. By H61der's inequality and since ( l / n ) fsn_l
II011gn dO - voln (K), we have
( v o l n ( K ) ) k / n ( v o l n ( L ) ) ( n - k ) / n ~ voln(K).
THEOREM 27. Let 0 < k < n and let L be an origin symmetric convex body in R n with C ~ - b o u n d a r y and positive curvature so that the distribution IIx IIL ~ is not positive definite. Then there exists an origin symmetric convex body K in R n with C ~ - b o u n d a r y such that the distribution IlxllL n+~ -- Ilxll~ n+~ is positive definite but voln (K) > voln(L).
PROOF. Since (llxllL~) ~ is a continuous sign-changing function o n S n-l, there exists an open subset 12 in S n-1 on which (llxllL-k) A is negative. Let f ~ C c~ (S n - 1) be a nonnegative (and not identically zero) function supported in S2. One can prove (see [59,
Aspects of the isometric theory of Banach spaces
923
L e m m a 5]) that the function f(O)r -k is the Fourier transform of a function g(O)r -n+k where g ~ C ~ (S n- 1). Define a body K by -n+k
IIx II~n+~ _ IIx IIL
e
(27r)n g(x)'
where e > 0 is small enough so that the body K is convex (a standard perturbation argument is that, given an infinitely differentiable function on S n-1 , one can choose a small enough e so that the differential properties of the norm I1" IIt-nq-p equivalent to convexity of L are preserved after adding an e-multiple of the ( - n + k)-homogeneous extension of this function). We have
(IIXIILn+P) A - (IIxlI-Kn+P) A - e f (O)r -p,
(lO)
so the distribution IIx IIZ ~+p - IIx II~;~+P is positive definite. On the other hand, by (10) and L e m m a 25
~
(llxllZk)A(o)f(O)dO=(27r)n( voln(L) -- f s n-I
eF/
IIOIIZPlIOII~ n+p ) . n-1
Since the quantity in the left-hand side of the latter formula is negative, we use H61der's inequality (as in Theorem 1) to see that voln (K) > voln (L). D PROOF OF THEOREM 24. Putting k --- 2 in part (a) of Theorem 22 and using the fact that the central section of a convex symmetric body is maximal among sections perpendicular to a given direction, we conclude that the function IIx II~n+3 is a positive definite distribution for every symmetric convex body K. Similarly, IIx I1~-n+2 and IIx I1~-n+l are positive definite (put k = 1 and k = 0 in Theorem 22). Now part (i) of Theorem 24 immediately follows from Theorems 26 and 22. To show (ii), let L be the unit ball of the space with the norm Ilxllz = Ilxl14+ ~llxl12, where e > 0 and I1" Ilq stands for the norm of the space lqn. By Lemma 8, the distribution IIx 114k is not positive definite, therefore IlxllZ ~ is not positive definite for small enough e. Using this value of e in the definition of L (the perturbation of the/~-norm was made to ensure that L has positive curvature) and using Theorem 27 we get a body K giving the desired example (again use Theorem 22 to connect the Fourier transform with the derivatives of parallel section functions). D The condition that IIx IIK ~ is positive definite, that we use in the proofs, has a clear geometric interpretation. For 1 ~< k < n, let us say that a star body K in R n is a k-intersection body of a star body if there exists a star body L in R n so that, for every (n - k)-dimensional subspace H of R n, YOlk(K A H -L) -- vOln-k(L f-) H). It was proved in [61 ] that an infinitely smooth symmetric star body K is a k-intersection body of a star body if and only if IIx IIZ-~ is a positive definite distribution.
924
A. Koldobsky and H. KOnig
Finally, we would like to mention that the isomorphic version of the Busemann-Petty problem is open and equivalent to the famous hyperplane (or slicing) problem (see [97] for details).
5. Approximation of zonoids by zonotopes A zonotope in R n is a special convex polytope, namely the Minkowski sum of finitely many segments I j, j = 1 . . . . . N, in R n,
Z N -- Z
yj" yj 6 Ij, j - - 1 . . . . . N
I j --
j=l
.
j=l
By a segment we mean a compact one-dimensional convex set. For simplicity, we will assume that 0 is the center of all segments; then 0 is the center of symmetry of ZN. A zonoid B is a convex body which can be approximated arbitrarily well by zonotopes in the Hausdorff metric. For n -- 2, all centrally symmetric convex bodies are zonoids, for n ~> 3, the unit balls B pn o f I pn are zonoids if and only if 2 ~< p ~< ec. Several authors studied the problem of approximating zonoids by zonotopes: what is the minimal number N = N ( B , e) of segments lj needed to approximate a zonoid B C R n up to e > 0 by a zonotope ZN -- ~ U _ 1 lj given by N segments, i.e., ZN C B C (1 + e ) Z N ? This is of particular interest for the Euclidean ball B -- B~. Before stating the estimates for N ( B , e) known for general B and B -- B~, we explain the dual functional analytic formulation. By definition, a zonotope ZN is a linear image of the unit cube B N. Thus, if the interior of ZN is non-empty, the norm II 9II induced by ZN in R n is a quotient norm of l u . Consequently, the polar of the zonotope, P -- Z~v, is the unit ball of an n-dimensional subspace of lN with norm N
ttxtt,-
jl<x,xj)l,
x E R n.
j=l
Choosing Xj E S n-1 , the value )~j is just 1/2 of the length of Ij. Similarly, the polar B ~ of a zonoid B is the unit ball of a norm
IIx II, - fsn-1 I<x, y) Idlz(y), where # is a (positive) measure o n S n - 1 . In other words, B is a zonoid if and only if B ~ is the unit ball of an n-dimensional subspace of L 1 ( S n-1 , IX); one may also take L 1 ([0, 1], IX), cf. [8]. For B = B~, IX is multiple of the usual surface measure. Thus, using the B a n a c h Mazur distance d, the above problem can be restated as follows: Given an n-dimensional subspace X of L1 = L1 (S n-1 , Ix) and e > 0, what is the minimal number N = N ( X , e) such that there is an n-dimensional subspace Yof l~ with d (X, Y) <~ 1 + e. Here X and Y
Aspects of the isometric theory of Banach spaces
925
have as their unit balls the polars of a zonoid and the approximating zonotope. The results of Section 3 show that, e.g., X -- 12 C L1 (S n - l ) but that 12 does not embed isometrically into l l, so one may not take e = 0. In this sense, the results presented here form an almost isometric counterpart to those in Section 3. Figiel, Lindenstrauss and Milman [24] showed that N(l~, e) <, ce-Z(lne-1)n, Gordon [36] improved it to ~< ce-2n. Johnson and Schechtman proved in [45] that N(lnp, e) <<,
Cpe-P'n for 1 ~< p < 2. Bourgain, Lindenstrauss and Milman [12] then established a general estimate for all Banach spaces X8 with zonoid unit ball B, N ( X ~ , e ) <~ ce-Z(ln e-1)n(lnn) 3. Talagrand [123] simplified their proof and improved the estimate at the same time, obtaining N ( X s , e) <<,ce-Zn(lnn). The constants c, Cp here do not depend on n and e. Up to the (ln n)-terms, the estimates are clearly optimal as far as the dependence on n is concerned. However, fixing n, the dependence on e is not the best possible. Better results on this setting were obtained in a series of papers by Bourgain, Lindenstrauss and Milman, Matou~ek and Wagner, for the Euclidean ball B~ as well as for the general zonoids B. In terms of the previous notation
{
NIj, ZN
N(B, e ) - min N c N: ~Z N - - Z
C B C (1 +
6)ZN
/
j=l
the best known result is THEOREM 28. (1) Let n ~ 2. Then there is a constant c(n) > 0 such that for all e > 0
N ( B ~ , e ) ~ c(n)e -2(n-1)/(n+2).
(11)
(2) Letn ~ 2 and B C R n be a zonoid. Then there is d(n) > 0 such thatforall 1 ~ e > O, ifn~5orn-2
N(B, e) <~d(n)e -2(n-1)/(n+2),
(12)
and if n = 3, 4 N(B, e) <~d(n)(e -2 lne-1) (n-1)/(n+2)
(13)
The segments of the approximating zonotope may be taken of equal length if n ~ 5 or ( B - B~ and n ~ 2). The lower estimate (11) was proved in [12] using spherical harmonics. Bourgain and Lindenstrauss [10] proved the upper estimate (13) for all zonoids (with a slightly worse logarithmic term if n - 4) and for B - B~ in all dimensions n ~> 2. Wagner [125] had proved that the approximating zonotope for B~ may be chosen to have segments of equal length if n ~< 6. Bourgain and Lindenstrauss [ 11 ] in a second paper removed this restriction n ~< 6 for B -- B~ (with logarithmic terms). Matou~ek [94] improved these upper estimates to the form stated in (2) of Theorem 28.
926
A. Koldobsky and H. KOnig
We sketch some basic ideas of the proof of Theorem 28. (i) To approximate a zonoid B by a zonotope Z N generated by N segments up to s > 0 means by the dual formulation that, for a given probability measure tt on S n - i , one has to find a discrete probability measure )~ - (~.j) on S n-1 supported in N points ( x j ) C S n-1 such that for all x E S n - 1 N fsn_ 11(X' Y) i d # ( y ) - ~ ~,jl(x, x j ) I < s . j=l
(14)
For B -- B~ the m e a s u r e / z is the normalized surface measure. In this case f s n _ l l ( X , Y)I x d/z(y) = fin is a constant depending only on n and (14) reads N
n-- jl(x,xj)l
< s,
(15)
x ~ S n-1.
j--1
That the segments of Z N have equal length then means that )~ is uniform, i.e., all ~ , j a r e equal, )~j - - 1 / N , j -- 1, . . . , N. (ii) For the lower estimate of N ( B ~ , s ) in (11), assume (15) to hold and put h ( x ) " ~Y-1
)Hl(x,xj)l,
x ~ S n-1. Then s > Ilfln - hllLz(m), m -- normalized surface measure
on S n - i . For each k ~> 0, let (Y~j)M(I'~) be an orthonormal basis of spherical harmonics of degree k on S n-1 . Expand h into spherical harmonics ex~ M(n,k)
(16) k=l j=l where (., .) is the scalar product in L2(m). By the F u n k - H e c k e formula [102] N
N
(h, Ykj) -- ~ ~j j--1
fsn_11(2,xj)lY j(x)dm(2) E jy j(xj), j=l
where calculations show that otk ~ k -(l+n/2) for even k, or0 - 1 and ot~ - 0 for odd k E N. The generating function of the generalized Legendre polynomials and the addition formula for spherical harmonics imply that, for every 0 <<,r < 1, x, y E S n-1 , 1 - r2
~
M(n,k)
(1 + r 2 - 2 r ( x , y))n/2 - PL-~ rk ~ k=0 j=l
Ykj
(y).
Hence U
1 -- r 2
~j j=l
oo
= Z r~/~ (1 + r 2 - 2 r ( x x j ) ) n / 2 ~=0 '
M(n,k) E (h, j=l
(17)
Aspects of the isometric theory of Banach spaces
927
and using (16) in the form oo M(n,k) 82 > Ilfln - hl122(m)-
E
E
(h, Ykj)2
k=l j=l we find using c~o = 1 and (17)
II s 2 max(r k/ak) k/>l
~
N
1 - E ~J j----1
1 -
(1 -k- r 2 --
r 2
2r(.,xj)
)n/2 L2(m)
N 1 --r 4 -1 -- j E )~i ~,j ,i=l (1 + r 4 -- 2r2(xi,xj)) n/2
-
Choosing r such that (1 - r2) 1-n -- 2N, the and (18) implies the desired estimate
maxk(rk/~k)
1.
(18)
is attained for k ,-~ (1 - r2) -1
c ( n ) e 2 N (n+2)/(n-1) >/1.
For more details, see [ 12, Section 6]. (iii) To prove the upper estimate for N(B, s) in (12) and (13), given a measure #, one has to construct points (xj) c_ S n-1 and a probability measure )~ = ()~j) on them so that (14) holds. The proofs in [ 10,11 ] and [96] are technically involved. In the first deterministic procedure, the sphere S n-1 is divided into N small p i e c e s ( Q J ) Y - 1 having diameters of order of magnitude N -1~(n-l) and (approximately) the same/z m e a s u r e , Iz(Qj) : 1/N. In the second step, in each of the Qj's, (n + 2) or (n + 1) p o i n t s (Yjl)l ~ Qj and weights ()~jl)l are chosen, in a probabilistic way or by employing methods of the geometric discrepancy theory ("irregularities of distribution"); the union of the points Yjl constitutes the sequence (xj) c_ S n-1 and (~)~jl) the measure on them. In [10] the set r j of probability D X-"n +2 m e a s u r e s O'j on Qj is considered which have the form o-j A-.,/=I )~jl6yjl with Yjl E Qj such that linear functions h are integrated exactly on Qj
NfQ hdlZ=fQ h d a j J
J
)
Xjlh(yjl) , /=1
i.e., the barycenter of crj coincides with the one of the probability measures Ntt[ Qj. Using Caratheodory's theorem and the separation theorem, one shows that N#] Qj belongs to the
A. Koldobsky and H. Kgnig
928
weak* closure of the convex hull of ~v'j. Since ~v'j is compact, the Krein-Milman theorem implies that there is a probability measure vj o n ~v,j such that N # [ Qj = f r
aj dvj (crj),
1 <~ j <~ N.
J Then v := I--[Y=l vj is a probability measure on Z "= I-IY-1 ~'j and a probabilistic deviation inequality shows that the probability of the set of those rr = (Crl, . . . , rrU) ~ r , crj of the above form, so that
n-, I(x' Y ) I d # ( Y ) - ~ Z ~j,l(x, yjl)[ > e / 2 j=l 1=1 is exponentially small for a fixed x
~ S n-1 ,
(19)
namely at most
2 e x p ( - d l (n)N(n+2)/(n-1)e2), and still less than 1 for all points in a suitable (e/4)-net in
S n-1
provided that
d2(n)N(n+2)/(n-1)e2/lne -1 < 1 which is the type of the condition in (13). Hence there is cr = ( o 1 . . . . . fiN) 6 X7 SO that (19) is false for any x in the (e/4)-net on S n-1 . For this rr, the left-hand side in (19) is at most e for all x ~ S n-1 , giving a measure )~ - ()~jl) on the N(n + 2) points (Yjl) c__S n-1 with (14). It does not matter that N is replaced by N (n + 2) since we disregard constants depending on n. The measure )~ is not uniform in general, however, and so the approximating zonotope will not (yet) be generated by segments of equal length. For the use of methods of the geometric discrepancy theory we refer to [94].
6. Exact estimates for projection constants By Lindenstrauss-Tzafriri's complemented subspaces theorem [77], a Banach space Z is isomorphic to a Hilbert space if and only if every closed subspace X of Z is complemented. This means that the relative projection constant of X in Z, ~(X, Z) "= inf{llPll" p2 = p E/~(Z) is a projection onto X} is finite. The proof shows that Z is isomorphic to H if and only if there is a constant c > 0 such that )~(X, Z) ~< c for every finite dimensional subspace X of Z. Conversely, if Z is not Hilbertian, there will be a sequence of subspaces Xn of Z such that )~(Xn, Z) tends to infinity together with dim Xn. If there is no projection in s from Z to X, like in the case X = co, Z = l~, put ~(X, Z) -- oo. The absolute projection constant is defined as ~(X) "= sup{)~(X, Z): Z is a Banach space containing X as a subspace}.
929
Aspects of the isometric theory of Banach spaces
Let n := dimXn < o~. Then X(Xn) < oo. In fact, by Kadets and Snobar [48], X(Xn) <<, in this n-dimensional case. In this section, we give some general (almost) exact estimates for absolute and relative projection constants in terms of n, improving this estimate slightly. Estimates for X(X) are useful to construct extensions of operators into X (or from X) since X(X) = inf{X > 0: YY c Z, T E s
X) 3T E s
X),
TIy = T, IITII ~< ZlITII}. This follows by embedding X isometrically into l ~ ( l ) and using the extension property of l ~ (I), i.e., coordinatewise application of the Hahn-Banach theorem. One gets for any such embedding X(X) = X(X, l ~ ( l ) ) . Further,
X(X) = y ~ ( X ) "-- inf{ IIRII IISIl R E s SEs
1~(I)), (I), X), S R -- Idx }.
As an example of )~(X) > 1, consider
This 2-dimensional space has the regular hexagon as its unit ball. The orthogonal projection P from R 3 onto X2 is the minimal norm projection with X(X2) = IIP I[ = 4/3. To estimate projection constants, we use trace-duality. By v and 7rp we denote the nuclear and p-summing norms, respectively. For their basic properties, see [19]. For the following, also see [44, Section 12]. LEMMA 29. Let X and Y be finite-dimensional with X c Y. Then X(X, Y) -- sup{ ]tr(T" X v-> X)]" T E s
v ( T ) = 1, T ( X ) c__X } .
PROOF. Since for projections P ' Y ~ X onto X, Itr(T" X v-->X)] = I t r ( T P ' Y ~ Y)I ~< v ( T P ) <~ v(T)IIPII - I I P I I , only the inequality ~< is non-trivial. There is a projection P0" Y v-+ X _ Y onto X of minimal norm, IIP0 II - X(X, Y). The convex sets
9- {s E s
IISII ~ IIPoll}
and
72 "--
P Es
P -- Po +
~ i=1
*
, X•
x i @ X i , n E N, x i E
C
y* ,
Xi E X
}
A. Koldobsky and H. K6nig
930
are disjoint since 79 consists of projections. Hence 13 and 79 can be separated by a linear functional on E(Y). By trace-duality, there is T e E(Y) such that Re(tr(TS)) ~< liP011 ~ R e ( t r ( T P ) ) ,
S ~/3, P ~ 79.
This implies that IIeo II - tr(T Po) and
v(T) = sup{ Itr(ZS)l/llSIl" 0 r S ~ ~(Y) } - 1. Considering P = P0 4- x* | x for x* e X •
x 6 X yields that T(X) c X.
Let Xn c_ YN with dim Xn = n and dim YN = N < co. Using Lemma 29, it was shown in [66] that
~(Xn, YN) <~f (n, N)"= ~/-n(v/-n/N 4- v/(N - 1)(N - n ) / N ) .
(20)
In particular, )~(X2, Y3) ~ 4/3 (equality in the above example). Cases of equality in (20) are discussed in [66]. Spaces with polytopes as unit balls embed isometrically into 1N, N finite. For operators T on l N, v(T) -----:rl (T). Then Lemma 29 and an approximation argument imply COROLLARY 30. Let X be finite dimensional. Then •(X) = s u p { l t r ( T ' X ~-->X)]; X c_.l~, T e s
Zrl(T)-- 1, T(X) c_. X}.
Since by [65, Proposition 5] the/2-norm of the eigenvalues of T (on l ~ ) is bounded by yr2(T) ~< yrl (T) -- v(T) -- 1, and since the trace tr(T : X ~ X) is the sum of n eigenvalues where n = d i m X , H61der's inequality yields )~(X) ~< ~/-ff, the result of Kadec and Snobar. n the map Tlx is in fact the identity map and In the case of X = l p, )~(lp)Ygl (Idl~)
-- tr(Ill~) = n.
For p = 2, using the Grothendieck-Pietsch factorization theorem one can easily calculate the Zrl-norm of l~, yielding )~(l~) ~ 2V~--~/-n for K = R. A similar fact holds for l~; both estimates were proved first by Grtinbaum [39]. Thus l~ is very badly complemented as a subspace of lc~; of course it is norm 1 complemented in any larger Hilbert space. The Kadec-Snobar estimate was improved in [69]: THEOREM 31. Let Xn be n-dimensional. Then
)~(Xn) <~g(n) :--
(2 + (n - 1)~/n + 2)/(n + 1) (1 4- (n - 1)~/n 4- 1)/n
if K - R,
(21)
if K = C.
We remark that g(n) - f (n, N(n)) where N(n) :-- n(n + 1)/2 i f K -- R, and N(n) :-- n 2 if K = C. There exist spaces Xn with equality )~(Xn) = g(n) if and only if there exist
Aspects of the isometric theory of Banach spaces N ( n ) equiangularvectors of s, t = 1 . . . . . N ( n ) with n = 2, 3, 8. For K -- R, n the regular dodecahedron,
Note that
931
(x~) c K n, i.e., Ilxl12 = 1 and I(Xs,Xt)l = a f o r ~ independent sr I f K = R, this is true f o r n = 2 , 3 , 7 , 2 3 , if K = C, f o r -- 2, 3, the unit balls of these Xn are the regular hexagon and respectively.
asymptotically g(n) = ~/~
~1
4-
O(~)ifK
__
R a n d g ( n ) = q r f f - ~14 -
O( 88 if K - C. The estimate (21) is almost precise also if n is such that N ( n ) equiangular vectors do not exist, as shown by examples in [63,68]" PROPOSITION 32. (a) Let n -- q 4- 1, q a prime power and N = n 2 - n 4- 1. Then there exist complex n-dimensional spaces Xn c__l u with )~(Xn) -- f (n, N). Thus
0 <. g(n) - )v(Xn) <. 1/(2n3/2). (b) Let n = 4 m Jr- 2 m 4- 1 f o r m E N. Then there exist real n-dimensional subspaces Xn of l~a (1~) with a 1 -unconditional basis and (relative) projection constant
)~(Xn) >1 v/-n
-
1.
We now indicate the main steps of the PROOF OF THEOREM 31. (i) By approximation, we may assume that X is embedded isometrically into 1N for some finite N E N, i" Xn ~-+ l N. By Corollary 30, there is T E s Zrl(T) -- 1 with T ( X n ) C Xn and tr(T" Xn ~ Xn) = )~(Xn). The norm zrl -- v on s
N) is the trace-dual of the operator norm, i.e., zrl (T) -- y~sN__l IlTesll~, where es =
(0 . . . . . 1 . . . . . 0) are the standard unit vectors. In fact, # -- ( # s ) N 1 , #s "--IITesll~, is the (discrete) Pietsch probability measure on the extreme points (e~)N1 of the unit ball in the dual (1N) * - l N with N
IITxll
x
(22)
N
t:l
Since [[/Zlll -- 1, the formal identity map j " l N ~ l x (#) has norm 1. Choose an orthonormal basis f l . . . . . fn in j i (Xn) c l u (/z). Then using (22) n
)~(Xn)
=
N
~(Tfu,
tr(Tlx,)-
u=l N
s--1
N
<-
u=l N
~
- ~~ (30
s=l
~fu(s)f.(t)
s--1 t = l
u--1
f~(s)f.
)
u--1
(90
N
11
ZE.s..
fu(s)(Tfu)(S)
u=l
N I1( T
tl
~fu(s)Tfu s--1
n
fu)l~(u) - ~ l Z s
-"
Z .s..l(Xs.X.)[. s,t=l
(23)
A. Koldobsky and H. KOnig
932 where Xs
-~(fu(S))u= 1 E
__
1
n
K n. The 12-norm of the ( n 1/2
Xs is (up to q/if) the square func-
tion of the fu'S, ~/-ffllXsll2 = ~ u = l (fu(s) 2) . It is a non-trivial fact, proved by using the Lagrange multipliers in [69], that in the extremal case of spaces with maximal (relative) projection constant in l N, the square function is constant, i.e., I[Xs[12 is independent of s. Since f l . . . . . fn were chosen of norm 1 in 1N (#), we get that [[xs [[2 - - 1. Hence [(Xs,
Xt)[ ~ 1.
(ii) Let 0 < c~ < 1. Then there is a unique polynomial p of the fourth order such that
lul ~ p(u) - yo +
y2 u2 -
y4U 4
U E [ - 1 1]
(24)
which touches u at c~ (p(ot) - c~, p1 (or) = 1) and satisfies p(1) - 1. Here Y0, 9/2, y4 depend on ot but are positive. Estimating the right-hand side in (23), we find N
+ Y21(Xs,x,)l 2 - y4l(Xs,Xt)l 4)
n
s,t=l N .yo +
- y4 --)2 . s . ,
I( s. x,)l 4
s,t=l N
2
Here n Y~s,t=l #s #tl(Xs, xt)[ -- 1 in view of the orthogonality of the f j ' s was used. By Sidelnikov's inequality used already in Section 3, N
Z
s,t=l
~s~t[(Xs'Xt)14~fs
n--I
fs
](x, y)l 4 d m ( x ) d i n ( y ) - fl n--1
where/~ - 3/(n + 2) if K = R and/3 - 2/(n + 1) if K = C. Thus
)~(Xn) <~nyo + Y2 - Y4~. Since Y0, ?'2, y4 depend on or, the right-hand side depends on c~. It is minimal for c~ = 1/~/n + 2 if K = R and c~ = 1/~/n + 1 if K = C and gives the value g(n) stated in Theorem 31. The proof shows that equality in the estimates requires that [(Xs, xt)l = ot or 1 since otherwise a strict inequality holds in (24): the vectors should be equiangular (the case of I(xs, xt)l = 1 for s # t is a degenerate one); also one needs "many" such (Xs) to get a large double sum in (23). In R 2, the 3 points el, ( - ( 1 / 2 ) e l + (~/-3/2)e2) provide such equiangular vectors giving as X2 the hexagonal space, in R 3 the 6 diagonals of the regular icosahedron provide such points; the dual of the icosahedron, i.e., the dodecahedron, giving the unit ball of the extremal example X3. [3
Idea for the examples in Proposition 32.
In a equiangular vectors in K n allow us to construct an g(n) by taking as its dual unit ball the absolutely N(n) -- n 2. For the values of n in (a), n 2 - n +
converse to the previous proof, N(n) n-dimensional space Xn with )~(Xn) convex hull of the points (Xs). In C n, 1 (almost as many as n 2) equiangular
Aspects of the isometric theory of Banach spaces
933
vectors can be constructed by looking at certain exponentials using B2-sequences. An easier construction of "almost" equiangular vectors (yielding slightly smaller projection constants) is given by the n 2 vectors in C n, 1 ((2zri(slj+szj2))) exp --if-
X(SI,S2 ) -- --~
I
=l
E Cn,
where s l, s2 E {1 . . . . , n } and n is a prime number. Gaussian sums yield 1
](x(~,,s2),x(t,,t2))} =
$2 r t2,
O, $2 = t2, Sl 5~ tl,
cf. Proposition 18 and [68]. The real construction in (b) uses incidence matrix combinations for projective planes of order 2 m , cf. [63]. Thus there are spaces Xn with a 1-unconditional basis with )~(Xn)/V/-ff --+ 1 as n --+ oo. For spaces with a 1-symmetric basis this cannot happen, see [67]: THEOREM 3 3. There is 0 < c < 1 such that the projection constant of any n-dimensional space Xn with a 1-symmetric basis is bounded by
k(x,) <~ cv/-~. Recall that a basis - which we may and will identify with the standard unit vector basis
(e j) in K n - is 1-symmetric if the norm of a vector x does not change if the coordinates of x with respect to (e j) are permuted or multiplied coordinatewise by scalars of modulus 1. For example, the s p a c e s l np have a 1-symmetric basis. Let G denote the symmetry group generated by the permutations and multiplication by scalars of modulus 1; in the real case G is discrete with IGI = 2nn!. Let m denote the normalized Haar measure on G. PROOF OF THEOREM 33 (Idea). For symmetric spaces X n , the estimate proceeds as in (23) except that we may average over G leading to
)~(Xn) ~< n s u p
{N L ~
.s.t
I(g(xs),xt)[dm(g)
s,t--1
} ,
(25)
where the sup is extended over finite sequences (Xs) N c S n-1 and ~ s =Nl #s = 1. Khintchine's inequality states that 1
~/~ Ilxl12 ~
E
• i=1
Ilxl[2,
Xi ri Ll
x ~" (Xi)in=.l E R n.
934
A. Koldobsky and H. K6nig
For "most" x, the expectation in the center behaves like ~/~/zr Ilx I]2; the following variant of Khintchine's inequality is proved in [67]:
I
vllxll2 ~ (1 - V ~ )
xiri
E
i=1
[[xl[~.
(26)
L1
With a large constant c instead of (1 - V/2-/Tr), this also follows from a corollary of the Berry-Ess6en theorem. Averaging over the subgroup Go of sign changes of G, (26) yields for arbitrary x, y E S n- 1 (R)
xiYiri
f G l ( g x , y)[ dmo(g) = E ~
o
i--1
L1
~[[(xiYi)ll2"+-(1--~)ll(xiYi)l]c
.
(27)
Further averaging over permutations a is required in (25), i.e., Xi should be replaced by x~(i); however, as easily seen,
Average~ Ilx~(i)yi 112
1
45
If for the smaller/~-norm, the corresponding average would be bounded by d/~/-ff with d < 1, (25) and (27) would imply Theorem 33 with c = ~/2/zr + (1 - v/2/rc)d < 1. This is false, however, if x and y are of very different form. In the sum (25), however, "diagonal" terms involve twice the same vector x = Xs (s = t), and for these x E S n-1 d Averagea Ilx~(i)xi 112 ~ ---~, Vn
d - 49/50,
(28)
cf. [67]; the methods to prove (28) are similar to those employed by Kwapiefi and Schtitt [71 ] to average norms over permutations. If x -- Xs and y = Yt have a rather similar form, (28) s011 holds for (xa(i)Yi). Combining this estimate "close" to the diagonal with a (worse) general estimate "off" the diagonal enables us to prove Theorem 33 by using (25) and (27). It is important here that the constant (1 - ~/2/zr) is optimal in (26). In the complex case, the Rademacher variables ri in (26) have to be replaced by Steinhaus variables and x/~/Tr by x/~-/2. Letting c = limn{)~(Xn)/V/-d: Xn 1-symmetric}, examples in [65] show c ~> ( 2 2~/2~/2~~-)-1 if K - R and c ~> (2 - 4%-/2) -1 if K = C. Problem: Are the latter inequalities in fact equalities? Inequality (26) is used in [67] also to prove that )~(lnp)/~/-ff --> d for 1 ~< p ~< 2 and n --+ where d = ~/2/zr if K = R and d - ~/~-/2 for K - C, independently of 1 ~< p ~< 2. []
Aspects o f the isometric theory o f Banach spaces
935
References [1] I. Aharoni, B. Maurey and B.S. Mityagin, Uniform embeddings of metric spaces and of Banach spaces into Hilbert spaces, Israel J. Math. 52 (1985), 251-265. [2] A.L. A1-Husaini, Potential operators and equimeasurability, Pacific J. Math. 76 (1978), 1-7. [3] T. Andr, Contractive projections on Lp-spaces, Pacific J. Math. 17 (1966), 391-405. [4] K. Ball, Cube slicing in R n, Proc. Amer. Math. Soc. 97 (1986), 465-473. [5] K. Ball, Some remarks on the geometry of convex sets, Geometric Aspects of Functional Analysis, J. Lindenstrauss and V.D. Milman, eds, Lecture Notes in Math. 1317, Springer, Heidelberg (1988), 224-231. [6] E Barthe, M. Fradelizi and B. Maurey, A short solution to the Busemann-Petty problem, Positivity 3 (1999), 95-100. [7] Y. Benyamini and J. Lindenstrauss, Geometric Nonlinear Functional Analysis, Amer. Math. Soc., Providence, RI (2000). [8] E.D. Bolker, A class of convex bodies, Trans. Amer. Math. Soc. 145 (1969), 323-345. [9] J. Bourgain, On the Busemann-Petty problem for perturbations of the ball, Geom. Funct. Anal. 1 (1991), 1-13. [10] J. Bourgain and J. Lindenstrauss, Distribution of points on spheres and approximation by zonotopes, Israel J. Math. 64 (1988), 25-31. [11] J. Bourgain and J. Lindenstrauss, Approximating the ball by a Minkowski sum of segments with equal length, Discr. Comp. Geom. 9 (1993), 131-144. [12] J. Bourgain, J. Lindenstrauss and V.D. Milman, Approximation ofzonoids by zonotopes, Acta Math. 162 (1989), 73-141. [13] J. Bretagnolle, D. Dacunha-Castelle and J.L. Krivine, Lois stables et espaces Lp, Ann. Inst. H. Poincar6 Probab. Statist. 2 (1966) 231-259. [ 14] M. Burger, Finite sets ofpiecewise linear inequalities do not characterize zonoids, Arch. Math. (Basel) 70 (1998), 160-168. [ 15] D. Burkholder, Boundary value problems and sharp inequalities for martingale transforms, Ann. Probab. 12 (1984), 647-702. [ 16] H. Busemann and C.M. Petty, Problems on convex bodies, Math. Scand. 4 (1956), 88-94. [ 17] N.L. Carothers, R. Haydon and P.K. Lin, On the isometries of the Lorentz function spaces, Israel J. Math. 84 (1993), 265-287. [18] E Delbaen, H. Jarchow and A. Petczyfiski, Subspaces of Lp isometric to subspaces of lp, Positivity 2 (1998), 339-367. [19] J. Diestel, H. Jarchow and A. Pietsch, Operator ideals, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 437-496. [20] S.J. Dilworth and A. Koldobsky, The Fourier transform of order statistics with applications to Lorentz spaces, Israel J. Math. 92 (1995), 411-425. [21] L. Dor, Potentials and isometric embeddings in L|, Israel J. Math. 24 (1976), 260-268. [22] M. Eaton, On the projections of isotropic distributions, Ann. Statist. 9 (1981), 391-400. [23] T.S. Ferguson, A representation of the symmetric bivariate Cauchy distributions, Ann. Math. Statist. 33 (1962), 1256-1266. [24] T. Figiel, J. Lindenstrauss and V.D. Milman, The dimension of almost spherical sections of convex bodies, Acta Math. 129 (1977), 53-94. [25] R.J. Fleming and J.E. Jamison, Isometries on Banach spaces: a survey, Analysis, Geometry and Groups: a Riemann Legacy Volume, Hadronic Press, Palm Harbor (1993), 52-123. [26] M. Frrchet, Sur la d~finition axiomatique d'une classe d'espaces vectoriel distances applicables vectoriellement sur l'espaces de Hilbert, Ann. of Math. 36 (1935), 705-718. [27] R.J. Gardner, Intersection bodies and the Busemann-Petty problem, Trans. Amer. Math. Soc. 342 (1994), 435-445. [28] R.J. Gardner, A positive answer to the Busemann-Petty problem in three dimensions, Ann. of Math. (2) 140 (1994), 435-447. [29] R.J. Gardner, A. Koldobsky and Th. Schlumprecht, An analytic solution to the Busemann-Petty problem on sections ofconvex bodies, Ann. of Math. 149 (1999), 691-703.
936
A. Koldobsky and H. K6nig
[30] D.J.H. Garling and P. Wojtaszczyk, Some Bargmann spaces of analytic functions, Function Spaces (Edwardsville, IL, 1994), Lecture Notes in Pure and Appl. Math. 172, Dekker, New York (1995), 123-138.
[31] I.M. Gelfand and G.E. Shilov, Generalized Functions 1. Properties and Operations, Academic Press, New York (1964). [32] I.M. Gelfand and N.Ya. Vilenkin, Generalized Functions 4. Applications of Harmonic Analysis, Academic Press, New York (1964). [33] A. Giannopoulos, A note on a problem of H. Busemann and C.M. Petty concerning sections of symmetric convex bodies, Mathematika 37 (1990), 239-244. [34] J.E Goethals and J.J. Seidel, Cubature formulae, polytopes and spherical designs, Collection: The Geometric Vein, Springer (1981), 203-218. [35] E Goodey and W. Weil, Zonoids and generalizations, Handbook of Convex Geometry, North-Holland, Amsterdam (1993), 1297-1326. [36] Y. Gordon, Some inequalities for Gaussian processes and applications, Israel J. Math. 50 (1985), 265-289. [37] E.A. Gorin and A. Koldobsky, On potentials of measures in Banach spaces, Siberian Math. J. 28 (1987), 65-80. [38] A. Grothendieck, Resum~ de la theorie metrique des produits tensoriels topologique, Bol. Soc. Mat. de Sao Paulo 8 (1956), 1-79. [39] B. Grtinbaum, Projection constants, Trans. Amer. Math. Soc. 95 (1960), 451-465. [40] C.D. Hardin, Jr., Isometries on subspaces of LP, Indiana Univ. Math. J. 30 (1981), 449-465. [41] C.D. Hardin, Jr., On the spectral representation of symmetric stable processes, J. Multivariate Anal. 12 (1982), 385-401. [42] C.D. Hardin, Jr. and L.D. Pitt, Integral invariants of functions and L P isometries on groups, Pacific J. Math. 106 (1983), 293-306. [431 C. Herz, A class ofnegative definite functions, Proc. Amer. Math. Soc. 14 (1963), 670-676. [441 W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84. [45] W.B. Johnson and G. Schechtman, Embedding lpm into l~, Acta Math. 149 (1982), 71-85. [46] W.B. Johnson, B. Maurey, G. Schechtman and L. Tzafriri, Symmetric structures in Banach spaces, Mem. Amer. Math. Soc. 217 (1979). [47] E Jordan and J. von Neumann, On inner products linear metric spaces, Ann. of Math. 36 (1935), 719-723. [48] M.I. Kadets and M.G. Snobar, Certain functionals on the Minkowski compactum, Math. Notes 10 (1971), 694-696. [49] N.J. Kalton and B. Randrianantoanina, Surjective isometries on rearrangement-invariant spaces, Quart. J. Math. Oxford Ser. (2) 45 (1994), 301-327. [50] J.H.B. Kemperman, General moment problem, geometric approach, Ann. Math. Stat. 39 (1968), 93-115. [51] A. Koldobsky, The Schoenberg problem on positive-definite functions, Algebra & Analiz 3 (3) (1991), 78-85; Translation: St. Petersburg Math. J. 3 (1992), 563-570. [52] A. Koldobsky, Isometries of Lp(X; Lq) and equimeasurability, Indiana Univ. Math. J. 40 (1991), 677705. [53] A. Koldobsky, The Fourier transform technique for convolution equations in infinite-dimensional lqspaces, Math. Ann. 291 (1991), 403-407. [54] A. Koldobsky, Generalized L~vy representation of norms and isometric embeddings into L p-spaces, Ann. Inst. H. Poincare Probab. Statist. 28 (1992), 335-353. [55] A. Koldobsky, Common subspaces of L p-spaces, Proc. Amer. Math. Soc. 122 (1994), 207-212. [56] A. Koldobsky, Isometries of L p-spaces of solutions of homogeneous partial differential equations, Function Spaces (Edwardsville, IL, 1994), Lecture Notes in Pure and Appl. Math. 172, Dekker, New York (1995), 251-263. [57] A. Koldobsky, Positive definite functions, stable measures and isometries on Banach spaces, Interaction Between Functional Analysis, Harmonic Analysis, and Probability (Columbia, MO, 1994), Lecture Notes in Pure and Appl. Math. 175, Dekker, New York (1996), 275-290. [58] A. Koldobsky, Intersection bodies, positive definite distributions and the Busemann-Petty problem, Amer. J. Math. 120 (1998), 827-840. [59] A. Koldobsky, Intersection bodies in R 4, Adv. Math. 136 (1998), 1-14.
Aspects o f the isometric theory o f Banach spaces
937
[60] A. Koldobsky, Second derivative test for intersection bodies, Adv. Math. 136 (1998), 15-25. [61] A. Koldobsky, A generalization of the Busemann-Petty problem on sections of convex bodies, Israel J. Math. 110 (1999), 75-91. [62] A. Koldobsky and Y. Lonke, A short proof ofSchoenberg's conjecture on positive definite functions, Bull. London Math. Soc. 31 (1999), 693-699. [63] H. K6nig, Spaces with large projection constants, Israel J. Math. 50 (1985), 181-188. [64] H. K6nig, Isometric imbeddings of Euclidean spaces into finite dimensional lp-spaces, Banach Center Publ. 34 (1995), 79-87. [65] H. K6nig, Eigenvalues of operators and applications, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 941-974. [66] H. K6nig, D.R. Lewis and EK. Lin, Finite dimensional projection constants, Studia Math. 75 (1983), 341-358. [67] H. K6nig, C. Schtitt and N. Tomczak-Jaegermann, Projection constants of symmetric spaces and variants of the Khintchine inequality, J. Reine Angew. Math. 511 (1999), 1-42. [68] H. K6nig and N. Tomczak-Jaegermann, Bounds for projection constants and 1-summing norms, Trans. Amer. Math. Soc. 320 (1990), 799-823. [69] H. K6nig and N. Tomczak-Jaegermann, Norms of minimal projections, J. Funct. Anal. 119 (1994), 253280. [70] J.L. Krivine, Plongment des espaces normes dans les Lp pour p > 2, C. R. Acad. Sci. Paris 261 (1965), 4307-4310. [71] S. Kwapiefi and C. Schiitt, Some combinatorial and probabilistic inequalities and their applications to Banach space theory, Studia Math. 82 (1985), 91-106. [72] J. Lamperti, On the isometries of certain function spaces, Pacific J. Math. 8 (1958), 459-466. [73] D.G. Larman and C.A. Rogers, The existence of a centrally symmetric convex body with central sections that are unexpectedly small, Mathematika 22 (1975), 164-175. [74] E L6vy, Th~orie de l'Addition de Variable Al~atoires, Gauthier-Villars, Paris (1937). [75] M. Lewandowski, Shifted moment problem for Gaussian measures in some Orlicz spaces, Probab. Math. Statist. 10 (1989), 107-118. [76] J. Lindenstrauss, On the extension of operators with finite dimensional range, Illinois J. Math. 8 (1964), 488-499. [77] J. Lindenstrauss and L. Tzafriri, On the complemented subspaces problem, Israel J. Math. 9 (1971), 263269. [78] W. Linde, Moments of measures on Banach spaces, Math. Ann. 258 (1982), 277-287. [79] W. Linde, On Rudin's equimeasurability theorem for infinite-dimensional Hilbert spaces, Indiana Univ. Math. J. 35 (1986), 235-243. [80] W. Linde, Uniqueness theorems for measures in Lr and CO(H ), Math. Ann. 274 (1986), 617-626. [81] W. Linde, Shifted moments of Gaussian measures in Hilbert spaces, Note Mat. 6 (1986), 273-284. [82] W. Linde, Uniqueness theorems for Gaussian measures in lq, 1 ~< q < cx~, Math. Z. 197 (1988), 319-341. [83] A. Lisitsky, One more solution to Schoenberg's problem, unpublished manuscript. [84] Y. Lonke, On zonoids whose polars are zonoids, Israel J. Math. 102 (1997), 1-12. [85] V.A. Lopachev, L P spaces of solutions of homogeneous linear differential equations with constant coefficients, Operators and their Applications, Leningrad. Gos. Ped. Inst., Leningrad (1985), 52-56. [86] V.A. Lopachev and A.I. Plotkin, Isometric classification of L P-spaces of pluriharmonic functions, Operators and their Applications, Leningrad. Gos. Ped. Inst., Leningrad (1983), 62-67. [87] V.A. Lopachev and A.I. Plotkin, Isometric mappings in LP-spaces ofpolyanalytic functions, Functional Analysis, No. 26, Ulyanovsk. Gos. Ped. Inst., Ulyanovsk (1986), 122-129. [88] E. Lukacs, Characteristic Functions, Griffin, London (1970). [89] W. Lusky, Some consequences of Rudin's paper "Lp-isometries and equimeasurability", Indiana Univ. Math. J. 27 (1978), 859-866. [90] E. Lutwak, Intersection bodies and dual mixed volumes, Adv. Math. 71 (1988), 232-261. [91] Y. Lyubich, On the boundary spectrum of a contraction in Minkowski spaces, Siberian Math. J. 11 (1970), 271-279. [92] Y. Lyubich and O.A. Shatalova, Almost euclidean planes, Funkt. Analiz i Priloz. 32 (1998), 76-78.
A. Koldobsky and H. K6nig
938
[93] Y. Lyubich and L. Vaserstein, Isometric embeddings between classical Banach spaces, Geom. Dedic. 47 (1993), 327-362. [94] J. Matou~ek, Improved upper bounds for approximation by zonotopes, Acta Math. 177 (1996), 55-73. [95] L. Mattner, Completeness of location families, translated moments and uniqueness of charges, Probab. Theory Related Fields 92 (1992), 137-149. [96] V.D. Milman, A few observations on the connections between local theory and some other fields, Lect. Notes in Math. 1317 (1988), 283-289. [97] V.D. Milman and A. Pajor, Isotropic position and inertia ellipsoids and zonoids of the unit ball of a normed n,~tmenstonal space, Geometric Aspects of Functional Analysis, J. Lindenstrauss and V.D. Milman, eds, Lecture Notes in Math. 1376, Springer, Heidelberg (1989), 64-104. [98] J. Misiewicz, On norm dependent positive definite functions, Bull. Acad. Sci. Georgian SSR 130 (1988), 253-256. [99] J. Misiewicz, Positive definite functions on l ~ , Statist. Probab. Lett. 8 (1989), 255-260. [100] J. Misiewicz, Substable and pseudo-isotropic processes - connections with the geometry of subspaces of L~-spaces, Dissertationes Math. (Rozprawy Mat.) 358 (1996). [ 101 ] J. Misiewicz and Cz. Ryll-Nardzewski, Norm dependent positive definite functions, Lecture Notes in Math. 1391 (1987), 284-292. [102] C. MUller, Spherical Harmonics, Lect. Notes in Math. 17 (1966). [103] A. Neyman, Representation of Lp-norms and isometric embedding into Lp-spaces, Israel J. Math. 48 (1984), 129-138. [104] M. Papadimitrakis, On the Busemann-Petty problem about convex, centrally symmetric bodies in R n, Mathematika 39 (1992), 258-266. [105] A.I. Plotkin, Isometric operators in spaces of summable analytic and harmonic functions, Dokl. Akad. Nauk SSSR 185 (1969), 995-997. [106] A.I. Plotkin, Isometric operators on subspaces of L p, Dokl. Akad. Nauk SSSR 193 (1970), 537-539. [107] A.I. Plotldn, Continuation of L P-isometries, Zap. Naucn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), Vol. 22 (1971), 103-129. [ 108] A.I. Plotkin, Isometric operators in LP-spaces of analytic and harmonic functions, Investigations on Linear Operators and the Theory of Functions, III, Zap. Nau~n. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), Vol. 30 (1972), 130-145. [109] A.I. Plotkin, An algebra that is generated by translation operators and LP-norms, Functional Analysis, No. 6: Theory of Operators in Linear Spaces, Ulyanovsk. Gos. Ped. Inst., Ulyanovsk (1976), 112-121. [110] G. Polya, On the zeroes of an integral function represented by Fourier's integral, Messenger Math. 52 (1923), 185-188. [ 111 ] G. Polya and G. Szego, Aufgaben und Lehrsatze aus der Analysis, Springer-Verlag, Berlin (1964). [112] L. Rabinovich, Explicit isometric and almost isometric embeddings, M.Sc. Thesis, Weizmann Institute of Science (1997). [ 113] B. Randrianantoanina, 1-complemented subspaces of spaces with 1-unconditional bases, Canad. J. Math. 49 (1997), 1242-1264. [114] B. Randrianantoanina, Injective isometries in Orlicz spaces, Function Spaces (Edwardsville, IL, 1998), Contemp. Math. 232, Amer. Math. Soc., Providence, RI (1999), 269-287. [115] B. Reznick, Sums of even powers of real linear forms, Mem. Amer. Math. Soc. 463 (1993). [116] J. Rosinski, On uniqueness of the spectral representation of stable processes, J. Theoret. Probab. 7 (1994), 615-634. [ 117] W. Rudin, L p-isometries and equimeasurability, Indiana Univ. Math. J. 25 (1976), 215-228. [118] R. Schneider, Zonoids whose polars are zonoids, Proc. Amer. Math. Soc. 50 (1975), 365-368. [119] R. Schneider, Convex Bodies: the Brunn-Minkowski Theory, Cambridge University Press, Cambridge (1993). [120] R. Schneider and W. Weil, Zonoids and related topics, Convexity and its Applications, Birkhauser, Basel (1983), 296-317. [121] I.J. Schoenberg, Metric spaces and positive definite functions, Trans. Amer. Math. Soc. 44 (1938), 522536. [122] K. Stephenson, Certain integral equalities, which imply equimeasurability of functions, Canad. J. Math. 29 (1977), 824-844. .
Aspects o f the isometric theory o f Banach spaces
939
[123] M. Talagrand, Embedding subspaces of L 1 into l~, Proc. Amer. Math. Soc. 108 (1990), 363-369. [ 124] A.V. Vasin, Isometries between L P-spaces ofpluriharmonic and polyanalytic functions, Functional Analysis, No. 26, Ulyanovsk. Gos. Ped. Inst., Ulyanovsk (1986), 62-67. [125] G. Wagner, On a new method for constructing good point sets on spheres, Discr. Comp. Geom. 9 (1993), 11-129.
[126] W. Weil, Zonoide und verwandte Klassen konvexer KOrper, Monatsh. Math. 94 (1982), 73-84. [ 127] J.H. Wells and L.R. Williams, Embeddings and Extensions in Analysis, Springer-Verlag, New York (1975). [128] V.P. Zastavnyi, Positive definite functions depending on the norm, Russian J. Math. Phys. 1 (1993), 511522. [129] V.E Zastavnyi, Positive-definite functions that depend on a norm, Dokl. Ros. Akad. Nauk 325 (1992), 901-903. [130] M.G. Zaidenberg, A special representation of isometries of function spaces, Studies in the Theory of Functions of Several Variables, Vol. 174, Yaroslav. Gos. Univ., Yaroslavl (1980), 84-91. [131] G. Zhang, Centered bodies and dual mixed volumes, Trans. Amer. Math. Soc. 345 (1994), 777-801. [132] G. Zhang, Intersection bodies and Busemann-Petty inequalities in R 4, Ann. of Math. 140 (1994), 331346. [ 133] G. Zhang, A positive answer to the Busemann-Petty problem in four dimensions, Ann. of Math. 149 (1999), 535-543. [134] A.A. Zinger, A.V. Kakosyan and L.B. Klebanov, A characterization of distributions by mean values of statistics and certain probabilistic metrics, J. Soviet Math. 59 (1992), 914-920. [135] V.M. Zolotarev, One-Dimensional Stable Distributions, Amer. Math. Soc., Providence, RI (1986).
This Page Intentionally Left Blank
CHAPTER
22
Eigenvalues of Operators and Applications Hermann K6nig Department of Mathematics, University of Kiel, Kiel, Germany E-mail: hkoenig @math. uni-kiel, de
Contents 1. Compact operators, eigenvalues and s-numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Weyl-type inequalities and s-number ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Eigenvalues of p-summing and nuclear operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Application to integral operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
HANDBOOK OF THE GEOMETRY OF BANACH SPACES, VOL. 1 Edited by William B. Johnson and Joram Lindenstrauss 9 2001 Elsevier Science B.V. All rights reserved 941
943 948 959 969 973
This Page Intentionally Left Blank
Eigenvalues of operators and applications
943
1. Compact operators, eigenvalues and s-numbers Classical results relate the order of decay of the eigenvalues of integral operators in L pspaces to the regularity and integrability properties of the defining kernel: the smoother the kernel is, the faster the eigenvalues tend to zero. These operators are clearly compact. Riesz' theory of compact operators admits an abstract quantitative analogue, as far as the decay of the eigenvalues is concerned, for operator classes like p-summing and nuclear operators or operators with summable s-numbers. These latter results will be presented here. Applications to integral operators will be given which extend the well-known classical theorems. The main ingredients of this theory are Weyl's inequality relating eigenvalues and singular numbers, factorization methods for operator ideals - see also [7] - and interpolation theory. All classes of operators which we consider form operator ideals, i.e., they are stable under compositions with continuous operators. To apply the abstract results, an integral operator T in L p is factored over a standard map, like a Sobolev imbedding map, which is known to belong to the operator ideal considered; then T itself belongs to the ideal. Remarks on the historical development of the theory can be found in Pietsch's book [32]. All Banach spaces X, Y. . . . will be complex. We let Bx = {x E X I [Ixll ~< 1} and denote the linear continuous and compact operators, respectively, between X and Y by s Y) and/C(X, Y), respectively, with/2(X) := s X) and/C(X) :=/C(X, X). A map T Es is power-compact if there is n E N such that T n E/C(X). Let a (T) denote the spectrum and p(T) the resolvent set of T. The classical Riesz' spectral theory asserts for power-compact operators T E s (i) For all )~ E C\{0}, (T - )~I) is a Fredholm operator and has finite and equal ascent and descent, say l, with X -- ker((T - )~I) l) @ im((T - )~I)l). (ii) All non-zero spectral values )~ E a ( T ) are eigenvalues of finite multiplicity. They have no accumulation point except possibly zero. These constitute the properties of a Riesz-operator. We recall that the ascent [descent] of (T - )~I) is the smallest integer 1 E 1Nsuch that [im(T - XI) l -- im(T -/kI)/+l].
ker(T - )~I) l = ker(T - )~i)/+1
The (algebraic) multiplicity m(T, )~) of )~ is the dimension of the space of principal vectors ker((T - )~l)l). Principal vectors associated with different eigenvalues are linearly independent. For integers k E 1N
a ( T k) -- a(T) k, m( Tk' #) -
a(T*) -- a(T), E
re(T, )0,
0 =/=# ~ a(Tk),
#=)k, )~Ea(T)
re(T*, )~) = m(T, ~),
0 :/: )~ E a(T).
H. KSnig
944
See, e.g., [32] or [17]. If T is a Riesz-operator, we denote by ()vn(T))nEl~ the sequence of eigenvalues of T, ordered in the following way: they are non-increasing in absolute value, i.e., ];Vn(T)[/> I)~n+l (T)I and each eigenvalue is repeated as often as its multiplicity indicates. If there are no more than n non-zero eigenvalues in this sense, we let ~n+~ (T) = 0 for k E N. The order could be non-uniquely determined; we choose a fixed order of this form. The following localization result is important for our purposes. PROPOSITION 1 (Localization). Let T E s be power-compact, n E N and )~n(T) # O. Then there is an n-dimensional subspace Xn of X, invariant under T, such that T ix, has precisely )~1(T) . . . . . )~n(T) as its eigenvalues. PROOF. If ( # l . . . . . /,t I =: #) are the different eigenvalues in the sequence (~1 (T) . . . . . )~n(T)) and ]#ll ~>"" ~> ]#l], we have # = )~n(T) and l-1
Z
1
m(T, # j ) < n <~Z
j--1
m(T, #j).
j--1 1-1
Let k "-- n - Y ~ j = I m(T, p t j ) and Yj "= ker(T - # j I ) m j , m j - - ascent of # j for j -1 . . . . . I. Then T(Yj) c Yj and a ( T i y j ) -- {#j} with m ( T i y s, # j ) = m(T, #j). By the Jordan decomposition theorem, the "last" space Yl admits a basis of principal vectors yl . . . . . yp, p = m(T, #) such that Tiy t : Yl --+ Yl has a matrix representation with respect to (yp . . . . . yl) which is a block diagonal sum of matrices of the form
# 1 O) 1 . 0
# l-1
Put Y / " - - S p a n [ y l . . . . . Yk], Xn : - - ~ j - 1 YJ (~ Y/" Then Y/ and Xn are invariant under T, Xn is n-dimensional and TIx,, has as its eigenvalues precisely the sequence (Zl (T) . . . . . )~n(T)). D1 Two operators T 6 s S E Li(Y) are related provided that there are maps P L;(X, Y) and Q E s Y) such that T = Q P and S = P Q. Pietsch's principle of related operators [32, 3.3.3] is very useful in the context of operators on X factorizing through Y. PROPOSITION 2 (Related operators). Let T E s and S E s and T be power-compact. Then S is power-compact and cr(T)\{0} -- a(S)\{0},
be related operators
m(T, )~) = m(S, )~) for all 0 # )~ E a ( T ) .
Hence we may write )~n(T) = )~n(S) for all n E N.
945
Eigenvalues of operators and applications
PROOF. Choose P, Q as above with T -- Q P, S = P Q. If T k E 12(X) is compact, so is S k+l = P T k Q E s For 0 ~ ~. E p(T), )~ E p(S) and conversely, since
(S - ) ~ I ) - 1 = ) - 1 ( P ( T - )~I)-1 Q _ I). For 0 :fi ,k E a ( T ) and 1 E N, (S - )~I) 1Px = P ( T - )~I)lx. Hence x w-~ Px defines a map q9 : ker(T - )~I) 1 --+ ker(S - )~I) 1. Similarly,
1 y~--~ Q / ~ ~ ( ~ ) ( - S / ~ )
j-I
j=l defines a map ~ : ker(S - )~I) 1 --+ ker(T - ),i)t, because
(T - )~I) 1 Q/L Z
'(,) j
(-S/)~)J-1 Y - Q/)~ Z
'(,) j
(-S/)~)J-l (s - )~I) l y.
j--1
j=l
One checks that ~ is the inverse of q). Therefore m(T, )~) = m(S, )~). Let T E K~(H, K) be a compact operator between (complex) Hilbert spaces H and K and T t denote the Hilbert space adjoint. The singular numbers sn(T) of T are the (nonnegative) eigenvalues of (the positive map) [TI := (T'T) 1/2 E )U(H), sn(T) := )~n(ITI) = )~n(TIT) 1/2 for n E N. By polar decomposition,there is a partial isometry U E s K) such that T = U ]TI, iTI = UIT. This polar decomposition and the spectral theorem for the compact selfadjoint operator ITI E K~(H) yields that there are orthonormal sequences (ONS) (Xn) C H, (Yn) C K of eigenvectors of T ' T and T T ~, respectively, such that for all
xEH Tx -- Z s n ( T ) ( x ,
(1)
xn)Yn.
n
The concept of s-numbers introduced by Pietsch in [28] generalizes the notion of singular numbers from the Hilbert space to the Banach space setting. We are mainly concerned with three particular examples of them related in some way to approximation processes. Let T E s Y) be an operator between Banach spaces and n E N, and
an(T) "-- i n f { l l T - Znll I Z~ E s
Y), rank(Tn) < n},
c . ( T ) "-- inf{ IlZlx~ III x~ ~ x has codimension < n in X}, xn(T) "-- sup{an(TA) l A E E(/2, X), IIAII ~< 1}. Here an(T) are the approximation numbers, cn(T) the Gelfand numbers and xn(T) the Weyl numbers. We will see below that these sequences coincide with the singular numbers sn(T) if T E ~ ( H ) . Thus we may write sn(T) for any of these sequences. The following
946
H. K6nig
properties hold for them: IITII = s~ (T)/> s2(T) ~>... ~> 0,
(2)
Sm+n-1 (S + T) <~sm(S) q- sn(T),
s n ( Q T P ) <~ IlOllsn(T)llell,
(3)
Rank T < n =~ Sn (T) = 0 and Sn (II~) - 1.
Here n , m E N, S , T ~ /2(X,Y), P 6 /2(X0, X), Q 6 s Y0). Any sequence Sn :/2 --+ •+ satisfying (2) and (3) is an s-number sequence by definition. The sequence of entropy numbers en (T) of T,
en(T) := inf,
q
~ 0 I :::ly1..... yq6y, q<~2n-1 " Z ( n x ) C U ( { y i } -~- ~ B y ) i=1
]
,
satisfies (2) but not (3). For a survey on the relation between these parameters see the article [10]. Taking Xn = kerTn, one gets cn(T) ~ an(T). By (2), also xn(T) ~ an(T) is true. If X = H is a Hilbert space and S ~ s Y), cn(S) = an(S) since, given Hn ~ H with codim Hn < n, choose Sn := S Pn where Pn = I - Qn and Qn is the orthogonal projection onto Hn to get rank Pn < n, an(S) <<. IIS - Snll - IISQnll -- IISInn II ~< cn(S). It follows that xn(T) <<.cn(T) <~an(T).
(4)
These three s-number sequences and the entropy numbers are multiplicative in the sense that Sm+n-1 (ST) <<.sm(T) sn(S),
T ~ E ( X , Y), S E s
Z).
(5)
This is not difficult, we verify (5) in the (most complicated) case of the Weyl numbers xn(T): Given T and S and A E s X) with IIAII = 1, let m, n 6 N and e > 0. Choose successively Um ~ s Y) and Vn ~ s X) with rank U m < m and rank Vn < n such that
IITA- Umll ~ (1 + e)am(TA), Ils(za -Um)- Wnll (a + e ) a n ( S ( T A -
Urn)).
Since rank (SUm + Vn) < m + n -- 1, a m + n - l ( S T A ) <<. I I S T A - S U m - Vnll ~ (1 4 - e ) a n ( S ( T A - Um)) ~< (1 + e) IIT A - U m II Xn (S) ~< (1 -+- e) 2 a m ( T A ) xn(S) <~ (1 + ~)2 xm(T) xn(S).
Eigenvalues of operators and applications
947
Hence Xm+n-1 (ST) <<,xm(T)xn(S) holds. We now prove the coincidence of the sequences in Hilbert spaces which was mentioned already. (In fact, all s-number sequences coincide there, cf. [28].)
PROPOSITION 3. Let H, K be Hilbert spaces and T ~ /C(H, K). Then
an(T) -- cn(T) -- xn(T) -- )~n (ITI). PROOF. Clearly, for T defined on H, xn(T) = an(T) which is also = cn(T) by (4). In this proof, we reserve the symbol Sn for the singular numbers, Sn (T) = )~n([T[). By the spectral decomposition of T,
T - Z sj(T)(.,xj)yj,
(Xn),(Yn) ONS,
y see (1). We choose
Tn'--Zsj(T)(',xj)y j
j,
rank Tn < n
to find that
an(T) <<,l i T - Znll --sn(T). To prove the converse, let Tn ~ s K) with m := rank Tn < n. There are sequences (z~) ~ H, (Zk) ~ K for k - 1 . . . . . m such that m
T. -
( ' , z k* ) z k 9
k--1 Let (O/j)jm 1 be a non-trivial solution of the (m • n) linear system n
Xj, Zk )Olj = 0 ,
k = 1, . . . , m ,
j=l /7
normalized by Y-~j=I IO/Jl2 Thus
/7
l. Then xo " - Y ~ j = I ~ j x j satisfies Ilx011- 1 and T n x o - O.
l i T - Tnll ~ I I ( T - Tn)xOII - IITxoII-
IotjlZsj(T) 2
sn(T),
j=l
and hence an(T) ~ sn(T).
D
948
H. KOnig
2. Weyl-type inequalities and s-number ideals
The starting point of all asymptotic estimates of eigenvalues of operators in this paper is THEOREM 4 (Weyl's inequalities in Hilbert spaces). Let T e 1C(H) be a compact operator in a Hilbert space H. Let n ~ N and 0 < p < oo. Then
I 1 I)~j(T)I ~< I2I s j ( T ) , j=l
(6)
j=l
n
n
Izj (T)I p ~<~ sj(r) p. j=l
(7)
j=l
PROOF. (a) We may assume that ~,n (T) -f: 0. By Proposition 1, there is an n-dimensional subspace Hn C_ H with T(Hn) C_ Hn, Tn := TBHn e 12(Hn) such that )~j(T) = )~j(Tn) for j = 1 . . . . . n. Let Tn = U ITnl be the polar decomposition of the finite rank map Tn. Then [det Tnl = det [Tn[ and n n i2 I YI I)~j(T)I = l--I I)~j(T~)I- Idet T~I = det IZ~lsj(Z~), j=l
j=l
j=l
since the determinant is the product of the eigenvalues. If in:Hn --+ H is the inclusion map and Pn : H --+ Hn the orthogonal projection, Tn = Pn T in. Proposition 3 implies that the s-number properties hold for the singular numbers so that sj(Tn) <<, IlPn els j ( T ) lOinII = sj (T). This proves the multiplicative inequality (6). In his original proof, Weyl [34] identifies I-Ij=l ])~j(T)] with the first eigenvalue of the alternating tensor product A n (T) on A n (H) and uses that VIj=I sj (T) can be identified with the operator norm of A n (T) with respect to the natural Hilbert space norm on A n (H). (b) The additive inequality (7) follows from (6) and the following, purely analytic lemma. El LEMMA 5. Let 0 < p < oo, n e N, al ~> ... ~> an >1 0 and bl >/... >/bn >~0 be such that k
k
llbJ
j--1
Then p
p
bj ~ Z aj " j=l j--I
fork-1
. . . . . n.
949
Eigenvalues o f operators a n d applications
PROOF. Let
Olj
--
p ln aj and/~j
0/1 ~ ' ' " ~Cgn ~--OO, k
-
p lnbj. Then
-
and
fll ) ' ' " ~ fin ) - - 0 0
k
Zflj<~~jotj, j--1
j - - 1 . . . . . n.
j--1
Define fn, gn" R ~ IR+ by n
f,(t)
([~j - t)+,
"-
g,(t)
"--
Z(OIj
j=l
--
t)+,
tER.
j=l
Then fn <. gn on R: If fll <~ t, fn (t) -- 0 <. gn (t) is trivial. If fll > t, let k "= max{j 6 N lflj > t}. Then also k
k
fn(t) "-- Z ( f l j
- t ) <<.Z ( o t j
j=l
- t ) <~ gn(t).
j=l
Integration by part gives
el3 -- fR (/3 - t)+e t dt. Hence
~bf j=l
e/3j --
s
s fn(t)e t dt ~
n gn(t)e t d t -
j=l
~e
n ~; -- y ' ~ a ; .
j=l
j=l
Theorem 4 is optimal since for positive operators )~n(T) = sn(T). We recall that an operator T 6/C(H) is a Hilbert-Schmidt operator if for some orthonormal basis (ONB) (en) c_ H of H, ~.n ]]Tenll 2 < oo. This is independent of the special ONB" if (fm) c_ H is another ONB, by Parseval's equality
'~ llVe, ll~ n
' ~ I
I<e,, V'/m>l ~ -- ' ~ llV'/mll ~
n,m
If T has the spectral decomposition (1),
~_~ IlTxnll 2 - ~ rl
sn(T) 2 --. hs(T) 2,
rl
and hence T is a Hilbert-Schmidt operator iff Y~n Sn (T) 2.
m
H. KOnig
950
We also need the (p, q)-(absolutely) summing operators on Banach spaces: Let 1 ~< p ~< q < cx~. T e E(X, Y) is (p, q)-summing, T e Hp,q(X, Y), if there is c > 0 such that for any finite sequence (Un) c_ X,
(Z
IlTunll q
)lq
(8)
<<.C 8p(Un),
11
where
8p(Un) -- sup
{(r
I(x*,
Un)l p
}
I I[x*llx, = 1 .
r/
The smallest constant c ~> 0 in (8) is the (p, q)-summing norm IJp,q(T). In the most important case p = q, we write Hp = l-Ip,p and 7rp(T) = 7rp,p(T). For the properties of this operator ideal of p-summing operators (Hp, Zrp) see [16] and [7] in this handbook. Defining A :lp, --+ X by en ~ Un, an application of the Hahn-Banach theorem shows that IIA : lp, ~
XII -
ep(Un).
(9)
We can now characterize the Hilbert-Schmidt operators. PROPOSITION 6. Let H be a Hilbert space and T ~ 1C(X). Then the following are equiv-
alent: (1) T is a Hilbert-Schmidt operator. (2) T is 2-absolutely summing. (3) If (I2, Ix) is a measure space such that H = L2(S2, Ix), there exists k ~ L2(S2 x I2, Ix x Ix) such that T is the integral operator defined by k, T f (s) -- fs? k(s, t) f (t) dix(t), If (1)-(3) hold, hs(T) = ~2(T) = IlkllL2(S~• summable with I&.(T)I 2
f~H. and the eigenvalues of T are square-
~ hs(T) -- rc2(T).
PROOF (Sketch). (1) =~ (2). Let (Um) C H be a finite sequence with 82(Um) decomposition (1). Then by Bessel's inequality
(10)
--
1 and T have the spectral
E I[TumII2 - E E Sn(z)21(um'xn)12 ~ E sn(Z)2 m m n n i.e., T e H e ( H ) with zre(T) ~< hs(T).
Eigenvalues of operators and applications
951
(2) =~ (1). For an ONB (en) G H, 82(en) = 1. Hence
)1/2 hs(T) --
Z
llTenll2
~<7r2(T).
n
(2) =r (3). If T has the spectral decomposition (1), the orthonormal systems (Xn), (Yn) _ H = L 2 (S-2, #) consist of L 2-functions and
k(s, t) "- Z
sn(T)yn(s)xn(t),
(s, t) e $2 x $2,
n
converges in L2(f2 x S-2,/.t • This is the kernel of T as an integral operator in H with IlkllL2(f2xS2) -- ( ~ n sn(T)2)l/2" (3) =~ (1). If (en) is an ONB of H, one uses Bessel's inequality to check
~](k, ~
IlZe~ll 2 F/
0 em)L2(S-2•
S2)[2 ~
Ilkll2L 2 ( a Q
•
9
n,/T/
The estimate for the eigenvalues of T follows from Theorem 4. In fact, for all 1 ~< p < oo, the p-summing operators coincide with the Hilbert-Schmidt operators on Hilbert spaces, though not isometrically if p r 2, cf. Pelczynski [27]. The eigenvalue estimate (9) is easily generalized to 2-summing operators on arbitrary Banach spaces. PROPOSITION 7. Let T E/72(X). Then T 2 is compact and
) 1/2 ~<7r2(T).
~-~ IXn(T)l 2 n
PROOF. By the Grothendieck-Pietsch factorization theorem, see [7] or [16], there is a probability space (12,/z), an isometric imbedding I ' X --+ L~(I-2, ~ ) , the natural 2summing inclusion map i ' L l ( S - 2 , #) --+ L2(12, #) and an "extension" T ' L 2 ( 1 2 , #) --+ X such that T -- T i I and I I T I I - zr2(T), zr2(i) -- 1. This implies that T 2 is Hilbert-Schmidt and compact. Let H = L2(S2, #), P "= i I ~ s H) and Q "- T ~ s X). Then T -- Q P and S "- P Q E s is a Hilbert-Schmidt operator since
hs(S)
- 7r2(S) ~ zr2(P)IIQII ~ zr2(i)lllll IITII- 7r2(T).
By the principle of related operators, Proposition 2, S and T have the same non-zero eigenvalues. Hence, using Proposition 6, IZn(T)12- ~ /7
n
I)~n(S)l2 ~ hs(S) 2 <~7r2(T) 2.
H. KiJnig
952
It is our aim to extend Weyl's inequality to Banach spaces, replacing the singular numbers by suitable s-numbers. For this purpose, we need another fact about the 2-summing norms due to Garling and Gordon [9]. The present proof goes back to Kwapiefi. LEMMA 8. Let X be an n-dimensional Banach space. Then
7r2 (Ix) = q/-n. PROOF. "~>" Since the identity map Ix has an n-fold eigenvalue 1, this follows immediately from Proposition 7. "~<" We claim that zr2(Ix) = sup{zr2(A) I[IA :12 --+ XII = 1 }. Clearly zr2(A) ~< [IAllrr2(Ix) = zr2(Is). Conversely, if (un) c__X is given with e2(Un) = 1, define A" 12 --+ X as in (9) with p -- 2. Then Y~n Ilun II2 -- ~ n IIAen II2 ~< zr2(A)e2(en) --
7r2(A). Now let IIA :/2 --+ X][ = 1. We have to prove that 7t'2(A) ~< v/ft. By factoring A over 12/ker A = 1~n with m ~ n, we may assume that [IA'l~ --+ X ll - 1. But then by the ideal property of the 2-summing norm and Proposition 6, 7r2(A) ~< IIAIIrr2(@) - h s ( @ ) - ~
<~~/-n.
The Weyl numbers xk(T) of 2-summing maps decay like O = (1/~/k). LEMMA 9. Let p >/2 and T ~ //p,2(X, Y) be a (p, 2)-summing operator. Then
supxk(T) k 1/p <~7rp,2(T). k
(11)
PROOF. The proof relies on an idea of D.R. Lewis. Since
xk(T) -- sup{ak(TA) I IIA'I2 ~ Xll -- 1} and ~p,2(TA) <~7t'p,2(T), we may assume that X = H is a Hilbert space. Let e > 0. We claim that there is an orthonormal sequence (fk) C_ H such that
(12)
xk(T) <<,(1 + e)IITAII.
For k = 1, choose fl 6 H, IIf~ II = 1 with Xl (T) = IITII ~ (1 + e) IITfl II. If k - 1 elements fl . . . . . fk-1 have been chosen, let Hk := Span(fl . . . . . fk-1) C H. Since dimHk < k, x k ( T ) = ck(Z) <~ IITIH~II. Thus there is f~ ~ H ~ with I l f ~ l l - 1 and xk(T) <~ (1 + e)IlZfkll. This proves (12). Since ee(fk) ~< 1 by Bessel's inequality,
(
~xk(T) p k
(
~ (1 -+-6) Z ][TfkllP k
~ (1 -Jr-8):rt'p,z(T),
if T 6/-/p,2(12, Y). For X g= H, we only have the weaker estimate in (11) since the maps A in the beginning of the proof may depend on k. D
953
Eigenvalues of operators and applications
The previous results may be used to extend Weyl's inequality to Banach spaces. Given a sequence x = (Xj)j6 N, we denote by 2 = ()cj)j6N the "doubled" sequence 2 = (x l, x l, x2, x2, x3, x3 . . . . ). The following result is due to Pietsch [29], extending a previous result for approximation numbers in [ 15]. THEOREM 10 (Weyl's inequalities in Banach spaces). Let X be a (complex) Banach space and T ~ E ( X ) be power-compact. Let c -- x / ~ , n E N, 0 < p < e~ and Cp - c 21/p. Then
(13)
I)~j(T)I ~ I--I (c2j(T)), j=l
j=l
1/p I)~j(T)I p
<<.Cp
j=l
(14)
~_xj(T) p j=l
PROOF. (a) Let T E E(X) be power-compact and n 6 N. By Proposition 1, there is an n-dimensional space Xn c_ X consisting of principal vectors such that T(Xn) c_ Xn and Tn := rlxn ~ s has precisely )~l(r) . . . . . 1.n(T) as its eigenvalues. By Lemma 8, zr2(Ixn) -- fin. Hence by the Grothendieck-Pietsch factorization theorem [22], there are maps A ~ E(Xn, H), B ~ s Xn) with B A = Ixn and Jr2(A) = x/~, IIBII = 1. We may assume that the Hilbert space H is n-dimensional (by restricting), H -- l~, so that B = A -1. Lemma 9 yields the Weyl number estimate xk(A) <. , f n / k for k = 1 . . . . . n. Define S, " - A T. A -1 6 s & has the same eigenvalues ()~j(T))~= 1 as Tn. Using Weyl's inequality in H (Theorem 4), the identity sj(Sn) = xj(Sn) (Proposition 3) and the multiplicativity of the Weyl numbers (5), we find for even n = 2m
IXj(T)I j=l
/)
IXj(Sn)l ~
j=l
r? s j ( S n ) - I--In xj(Sn) j=l
j=l
m
m
= H X2k-1 (~n)X2k(Sn) ~ H X2k-1 (Sn) 2 k=l
k=l
m
H(xk(Tn)xk(A)
I~I j--1
m
k=l
k=l
I-Ix (r.) 2
IIA-1II) 2
k-1
By Stirling's formula H~:m=l(n / k) <~c n with xk(Tn) --xk(iTn) <~xk(T). Thus
m
c -
x/~.
If
i" Xn --~ X is the natural injection,
/7
I)~/(T)I ~ V I (c2j(T)). j--1
The argument for odd n ~ N is similar. This proves (13).
H. K6nig
954 (b) By (13) and Lemma 5
fi
n I~'J (T)Ip <~~-~(c2j(T)) p = 2cP Zg/ xJ (T)p
j=l
j=l
j=l
Thus (14) holds with Cp = 21~pc.
D
PROBLEM. It is unknown whether there is c > 0 such that for all Riesz operators T E s and all n E N n
n
[I Izj(r)l H(cxj(r)) j=l
(15)
j=l
holds, i.e., whether the doubled Weyl number sequence could be replaced by the Weyl number sequence itself. Clearly c in (15) would have to be > 1, as simple examples like T -- (~~ 912 --+ 12 show. The problem is equivalent to the question whether the constants Cp in (14) can be chosen to be bounded independent of p as p --+ 0. The best known estimate for Cp as p --+ 0 is Cp <~2e/v/-fi, cf. [17, 2.a.6]. By Theorem 10, operators with pth power summable Weyl numbers will have absolutely pth power summable eigenvalues provided they are power-compact. To prove that this is true, we use a proposition comparing Gelfand and Weyl numbers and giving a weak converse of the general inequality Xn (T) <~Cn(T). The result is also due to Pietsch [29]. PROPOSITION 11. Let T E s
f cj(T) i
Y), n E N and c = 2e. Then
<~ (cn) n/2 FI .icj(T).
j--1
j=l
PROOF. (a) Let e > 0. Inductively, we construct vectors xi E X , y j E Y* with Ilxill ~ 1,
IlYjll ~< 1,
yj(Txi) = 0
for j < i and cj(T) <~ (1 + e)lyj(Txj)l. If they have been found for i, j < k, let
. - {x
s l y J (Tx) -- 0 for all j -- 1 . . . . . k - 1 }.
By definition of the Gelfand numbers, there is Xk E Xk, Ilxk II : 1 with
ck(T) ~ IITIx, II ~ (1 -t-~)llTx~ll. Choose y~ E Y* with
Ily~ll =
1 and
IITxkll
= y~(Txk).
Eigenvalues of operators and applications
(b) Now define A E s
X) and B E s l~) by ei tively. By H61der's inequality, [IA II ~< J-if. Also
955
e-->Xi and y e-->(Y~(Y))~=I, respec-
Hence, using again L e m m a 9, Xk (B) <~ ~fff-/k for k - 1 . . . . . n. The operator Sn "-- B T A E /2(1~) has a lower triangular matrix representation (yj (Txi))j,i=l. Thus by Weyl's inequality (Theorem 4)
fi [yj(Txj)[ -
[detSn[- f i
j=l
IXj(Sn)[ <~ fi xj(Sn).
j=l
j=l
Using the multiplicativity, we find similarly as in the previous proof
(1 -+- s) -n
iI
cj(T) ~
j--1
fi
n
xj(Sn) ~ (cn) n/2 I-I 2 j ( T ) . j--1 j=l
[]
We next introduce the Lorentz sequence spaces lp,q and prove Weyl's inequality for the Weyl number ideals based on lp,q. Given a null sequence x = ( X n ) n E N E CO, the decreasing rearrangement x* of x is defined inductively by x *1 -- Ilxll~ and
n-1 * " - max Xn Ilnl=n
E Ixjl j=l E* jEIn --
X j,
n>l.
Thus x~ ~> x~ ~> ..- ~> 0 and the sets {Ixjl} and {x j} are the same, elements occurring with the same multiplicity. Let 0 < p < oo, 0 < q < oo. The Lorentz sequence space lp,q is defined by
Ip,q "-- {x E co IIIxllp,q <
IIx IIp,q "--
~},
(EnENX*nqnq/p--1)l/q
0
SUPn Xn* nl/P
q--oo
For p -- q, lp,q is the usual space lp with I1" IIp,p -- I1" lip. For q -- oo, lp,e~ is also called weak lp. The spaces are ordered lexicographically: Let 0 < p l , p2 < 00, 0 < ql, q2 ~< 00. Then
lpl,q I ~ lp2,q 2
if Pl < P2 or if (Pl -- P2 and ql < q2).
H. KOnig
956
The spaces lp,q are quasi-Banach spaces. In the case of 1 < p < cx~, 1 ~< q ~< p, I1" IIp,q is actually a norm and in the case of 1 < p < oo, 1 ~< q ~< cx~, there is a norm I1" I1"p,q equivalent to the quasinorm I1" Ilp,q which is given by Ilxllp,q "-- IIx**llp,q where x** =
(Xn**)n~N, Xn** - nl ~j=ln
(xj + yj)* ~ j=l
xj + j=l
I1" IIp,q and
the equivalence of
* I1" IIp,q
xj.* The triangle inequality for
(
follows from
yj, j=l
II 9IIp,q from Hardy's inequality [12]
)"q(p' Ex: qnq/'-I) nEN
1/q (16)
nEN
where the constant pt is the best possible (if p - q). The analytic L e m m a 5 admits a variant for Lorentz spaces. LEMMA 12. Let a -- (an), b = (bn) c_ R+ be non-increasing sequences with
f
i
bj <~ I I
j=l
aj
f o r all n ~ N.
j=l
Let O < p < oo, O < q <<.oo and
1 p~q} Cp =
el/p
P < q
Then a ~ lp,q, implies b E lp,q with Ilbllp,q ~ cpllallp,q. PROOF. The case of p >~ q follows directly from L e m m a 5, applied to the non-increasing sequences (aj j - u ) , (bj j - u ) where ot = 1/q - 1 / p ~ O. If p < q, choose r > 1 such that s "-- p r / q > 1. Using L e m m a 5 with index q / r (instead of p) and Hardy's inequality (16), we find
llbll,,,q
-
II (bq/r)
IIr/qs,r~
I1( bjq/r)
q,r.)
/n
j=l
n s,r
uj j=l
r/q
/n
n s,r
<<. (s')r/qll (a q/r) IIr/q -- (s')S/Pllallp,q. s,r For r --+ oo, i.e., s --+ ~ , (st) s / p ~ e 1/p.
[--]
The definition of the Weyl number ideals Sp,q is now straight-forward. They generalize the Schatten class ideals in Hilbert spaces. Again let 0 < p < oo, 0 < q ~< oo and X, Y be
Eigenvalues of operators and applications
957
Banach spaces. The Weyl number ideals Sp,q are defined by
Sp,q(X, Y)"-- {Z E s
Y) l o'p,q(T) "-- II
I1 , q
Then (Sp,q, ~yp,q) is a quasinormed operator ideal, i.e., r are quasinorms on the component vector spaces Sp,q (X, Y) and the Sp,q-property is invariant under compositions with linear operators on both sides, as for compact or p-summing operators. For X = Y, we abbreviate Sp,q(X) -- Sp,q(X, X) and Sp = Sp,p. THEOREM 13 (Weyl's inequality for Lorentz-s-number ideals). Let 0 < p < cx~, 0 < q <. and T E Sp,q(X). Then T is power-compact and the eigenvalues of T belong to lp,q with
II (r))II p,q dpap,q(r), where dp -- 21/p~r
if p >/q and dp -- (2e) 1/p+l/2 if p < q.
PROOF. (a) Let T E Sp,q (X). We first prove that T is power-compact. Enlarging p when p < q, we may assume that T E Sp(X) since lp+s c_ lp,q for all q. Let m E N with m ~> p/2. The multiplicativity of the Weyl number implies that T m E Sp/m (X) C_ 82 (X). Since n" <<,n!e n, Proposition 11 implies for S := T m E S2(X)
[-I (j-1/2cj(S)) <~fi (dAj(S)), j=l
d :--- ~/r2e.
j=l
Since (xj(S)) E 12, also (d2j(S)) E 12. Thus by Lemma 5, also ( j - 1 / 2 c j ( S ) ) E 12, i.e., ~ j E N c j ( S ) 2 / j < (Y3. This implies that l i m j ~ c c c j ( S ) - 0 which yields that S - T m is compact. (b) Formula (13) of Theorem 10 and Lemma 12 imply that the eigenvalue sequence ()~n(T)) belongs to lp,q and
II(x.(T))ll ,q <<. I with Cp as in Lemma 12 and c -- ,~/~.
V]
Theorem 13 again goes back to [29] and [15]. PROBLEM. Let 0 < p < 2, 0 < q ~< cx~ and 1/r = 1/p - 1/2. Suppose that the Weyl numbers of T E L~(X, Y) belong to the Lorentz sequence space lp,q. Does it follow that the approximation numbers of T lie in the sequence space Ir,q 9. For concrete operators like integral operators, the Weyl numbers can be estimated from above either by the approximation numbers and these in turn by approximation theory
H. K6nig
958
results provided the defining kernel is regular or by using Lemma 9 and proving a (p, 2)summing property given integrability conditions of the kernel. Both can be combined by applying the multiplicativity of the Weyl numbers. Thus Theorem 13 may be used to find asymptotic upper estimates for the eigenvalues of various integral operators in Lp-spaces. A few examples are considered in Chapter 4. In various concrete cases, it is also possible to find upper estimates for the entropy numbers, see Carl and Stephani [3, Chapter 5]. The basic estimate of eigenvalues by entropy numbers, due to Carl and Triebel [4], may then be applied. It states: THEOREM 14. Let T ~ E(X) be a Riesz operator and k, n ~ 1% Then
1~,1(T) . . .~,n(T)] 1In <~2(~:-l)/2nelc(T). In particular, I~n(T)l ~< x/2 en(T). PROOF. Let k, n 6 1~ and ,kn(T) =/:O. By Proposition 1, there is an n-dimensional subspace Xn C_C_X with T(Xn) c_ Xn such that Tn "-- Tlxn E E(Xn) has precisely/~,1 ( T ) . . . . . ~,n(T) as its eigenvalues. Fix 6 > e~(T). Then there exist yl . . . . . YN E X, N "-- 2 k-1 such that N Tn(Bxn) C___U({Yi} + ~Bx) N Xn. i=1
(17)
We identify the complex n-dimensional space Xn with ]I~2n to use volume and determinants. The operator Tn, as a map in R 2n, has determinant equal to 12.1(T)...)~n(T)]2; this is because Tn" ~2n ~ ]I~2n admits a super-diagonal (2 • 2) block matrix representation Jordan form, with blocks on the diagonal
(
Re)~i(T) Im)~i (T)
-Im~ki(T) ) Re)~i (T)
'
having determinant 12.i(T)I 2. Denoting the volume in ]~2n by vol2n, (17) implies 12,1(T)...)~n(T)12vol2n(Bx,,) -- vol2n(Tn(Bxn)) <~2k-182nvol2n(BX,,). Hence [~.l (T).--)~n(T)I 1/n <~ 2(k-1)/2n•. The second inequality follows for k - n since I)~n(T)I ~< I)~l(T)...)~n(T)l ~/n. D The number N - 2 k-1 of balls in the definition of ek(T) is chosen to make this characteristic multiplicative; the previous proof also works for
{
N
eN(T) "--inf e/> 013y , .....YNeY" T ( B x ) C U ( { y i } + eBy) i=1
I
Eigenvalues of operators and applications
959
giving
[)~j(T)...)~,(T)I 1/~ <. N1/2nSN(T). Introducing the quantity gn(T) := infuel~ N1/2neN(T), we thus have
I ~ ( T ) . . . ~ ( r ) l ~/~ ~ gn(T),
(18)
see Carl and Stephani [3]. Recall that by the spectral radius formula for a Riesz operator T Es
12.,(T)l-lim
IITml
lim s,(Tm) 1/m
m----~ o o
m----> o o
This formula admits a generalization for higher eigenvalues and s-numbers; also (18) allows a variant with an equality:
PROPOSITION 15. Let T e s
be a Riesz-operator. Sj be an s-number sequence and
n ~ N. Then I,kn(T)]-- lim sn(Tm) 1/m
and
m --> o o
1 / " - lim gn(Tm) 1/m m---~ o o
For the proof and related results we refer to Makai and Zemanek [25], Carl and Stephani [3] and [ 18].
3. Eigenvalues of p-summing and nuclear operators The eigenvalues of p-summing operators are pth power absolutely summable provided that p ~> 2. This is the main result of this chapter, it extends the simple case of p = 2 given in Proposition 7. Moreover, nuclear operators on a Banach space X have absolutely summable eigenvalues iff the space X is isomorphic to a Hilbert space. An immediate consequence of theorem 13 (for q = oc) and Lemma 9 is COROLLARY 16. Let p ~ 2 and T E //p,2(X) be (p, 2)-summing. Then T is power-
compact and its eigenvalues decay of order O(n -I/p) with sup I)~n(T)l n 1/p <<.2earp,2(T). nE1N
As mentioned, for p-summing maps T ~ ITp(X) c_/-/p,2(X) with p ~> 2, this can be improved to (~.n(T)) E lp. To prove this, we need a tensor stability result for p-summing
H. KOnig
960
operators due to Holub [ 14]. The tensor product X | Y can be identified with a subspace of E(X*, Y),
Z -- ~
Xi @ Yi E X @ Y v-+ x* ~
i=1
(X*, xi)Yi
E
s
Y).
i=1
Here xi E X, Yi E Y, x* E X*. The e-tensor product on X | Y is the restriction of the operator norm on s Y) to X @ Y, m
e(z) - s u p { Z ( X * , xi) (y*, Yi)
IIx*llx*- Ily*llY* - 1}.
i=1
For the normed space we write X | Y and denote its completion by X ~e Y. Given S E s X1), T 6 s Y1), there is a well-defined operator S +~ T E s (X +~ Y, X1 +~ Y) induced by (S @ T)(z) -- zim=l Sxi | Tyi. Holub's stability result reads: PROPOSITION 17. Let 1 <~p < cx~. The operator ideal Hp of p-summing operators is stable under forming e-tensorproducts. More precisely, if S E Hp (X, X l) and T E 17p (Y, Y1), then S +e T E 17p (X 5~ Y, X1 +e Y1) with
7rp(S +e T) <~:rcp(S)zrp(T). PROOF. By the Grothendieck-Pietsch characterization theorem for p-summing maps, there are Borel probability measures #, v on the dual unit balls K -- Bx. and L -- Br. in their weak*-topologies such that for all x 6 X and y E Y
[]Sxl[ ~< top(S)
](x*,x)l p d#(x*)
,
IITyll ~< zrp(r)
I(Y*, Y)I p dr(y*)
.
Let z E X | Y again have the form z -- ~im=l xi | Yi, xi E X, Yi E Y. Then (S @ T)(z) = ~~im=l Sxi @ Tyi and for arbitrary a* E Bx T, b* E Br,~ m
~ ( a * , Sxi) (b*, Tyi) i=1
<~ S
(b*, Tyi )xi i=l
1/p
<<.rrp(S)
(b*, ryi) (x*,xi) i=1
d/z(x*)
Eigenvalues of operators and applications
961
1/p (x*, Xi) Yi
T
d#(x*)
i=1 m
Z ( x * , xi) (y*, Yi)
p
) lip dr(y*) d/z (x*)
i=1
This means that
e((S | T)z) ~ Zrp(S)Trp(T) ( fK • I(x* | y*, z)l p d(/z x v)(x*, y*))
lip
This inequality can be extended to the completion X +e Y. Since (# x v) is a Borel probability measure on K x L, this means that S @e T is p-summing and Zcp(S ~ T) <~
7rp(S)Trp(T).
D
The eigenvalues theorem for p-summing operators was first proved in [15]. THEOREM 18. Let 1 <. p < oo, q "--max(p, 2) and T ~ 17p(X) be p-absolutely summing. Then T is power-compact and its eigenvalues are q th power absolutely summable with
) 1/q Z
I)~n(T)l q
<~Trp(T).
ncN
PROOF. We present a proof due to Pietsch [32, 3.7.2] using the "tensor product trick". If 1 <~p <<.2, 17p(X) c/72(X), and the result follows from Proposition 7. Now let 2 < p < cx), q = p. Let T c 17p(X). Then T ~ 17p,2(X) and by Corollary 16, T is power compact with sup I)~n(T) In 1/p <~2e~p,2(T) ~< 2ezrp(T). nEN
This implies for any r > p
)l/r ~
I&~(T)I r
~ CprZrp(T),
(19)
nEN
where Cpr " - 2e(p/(r - p))1/r. Let dpr denote the smallest possible constant Cpr such that (19) holds for all p-summing operators in arbitrary Banach spaces. Thus dpr ~ Cpr. We claim that dpr -- 1. Again take T ~ 17p (X). If x, y E X are eigenvectors (principal vectors) associated to the eigenvalues )~n (T),)~m (T), then x | y is an eigenvector (principal vector) of T ~ T on X ~ X. Hence T ~ T has all values Xn(T)Xm(T) among its eigenvalues.
H. K6nig
962
Since T @E T is p-summing by Proposition 17, we find
)2/r Z
r
I;~n(T)lr
)l/r
nEN
nEN
dpr yrp(T@eT) ~ dprrCp(r) 2. d2r <~dpr,
dpr
rgp(T) for
This implies t h a t i.e., that ~< 1. This proves that II()~n(T))llr <<. all r > p. Taking the limit for r --+ p, we get that ()~n(T)) E Ip and
The result is the best possible: If p ~> 2, any/p-sequence defines a p-summing diagonal operator on l p or lee. If 1 ~< p < 2, there are even nuclear operators with only squaresummable eigenvalues, see the proof of Theorem 22 below. The tensor stability in Proposition 17 may be extended to compositions of pi-summing maps. A corresponding argument as above then yields
m
COROLLARY 19. Let 2 <<.Pi < (x) for i -- 1,.. 9, m and 1/p " - - ~ i = 1 1/pi. Assume that T E 17pl o . . . o I-lpm (X), i.e., T is a composition of pi-summing maps, over Banach spaces possibly different from X. Then T is power-compact with pth power summable eigenvalues and
)lip ~_, [Xn(Tll p
<~ Yrpl 0 . . . 0
Yrpn (T).
nEN
The latter quasinorm is the infimum of the product of the pi-summing norms of composition factors of T. PROPOSITION 20 ([17, 2.b.ll]). Let 1 <. v < u <. oo and T E s Let I :Iv --+ lu denote the formal identity map and 1/p := 1/v - 1/u. Then I T E 12(lu) is compact and (~.n(IT)) E lp,ee. If v <~ 2 <. u, one even has ()~n(IT)) E lp. PROOF (Idea). By a result of Pitt, any operator T E s lv) is compact, if u > v. Thus IT Es is compact. First assume that 1 ~< v < u ~< 2. By a result of Bennett [1] and Carl [2], the identity map I ' I v --+ lu is (p, 2)-summing if 1/p - 1/v - I / u , with 7rp,2(/) ~< .~/-2. Thus, using Corollary 16,
sup lXn(IT)l nl/p <<.2erCp,2(IT) <<.~/-22elITll. rl
Now suppose that 1 ~< v < u -- 2. Again I T E/-/p,2(12), but in this case the eigenvalues of I T are even pth power summable: if I T = ~_.n~i~Sn(IT)(',xn}Yn is the spectral decom-
Eigenvalues of operators and applications
963
position (1) of I T, e2 (Xn) ~< 1 by Bessel's inequality and thus
1/p O'p(IT) --
<~rrp,2(lT) ~
s.(IT) p
~/211TII.
11
By Weyl's inequality in Hilbert spaces, T h e o r e m 4, we conclude
)l/p Z
])~17(IT)iP
<~~p(IT) <~x/2 IITll.
17
The situation 2 ~< v < u ~< ec is dual to 1 ~< v < u ~< 2. This leaves the case of 1 ~< v ~< 2 u ~< cx~. In this case T ~ s lv) factors through 12, cf. [21], T = SR, R ~ s S E s IlSll IIRII ~< 2 IITII. Let I1 :lv --+ 12, 12 :/2 --+ lu be the natural inclusions. Then I T = I S R has the same eigenvalues as (I1S)(RI2). Let 1/pl = 1 / 2 - I/v, 1/p2 = 1 / u 1/2. By the previous case and its dual, Crp, (I1 S) ~ ~
IISII,
Crp~(R/2) ~< ,/2 IIRII.
Then 1/p = 1/pl + 1/p2. Weyl's inequality and the multiplicativity of the singular numbers implies
)l/p Z
<~ Crp(I1S RI2) <~c apl (I1 S) O'p2(RI2)
IX17(IT)I p
17
~< 2cl] Sll IIR II ~< 4c IIT II. The statement actually holds with c = 1. Since sequences in lp define bounded maps from lu to lv if v < u and 1/p = 1/v 1/u, by H61der's inequality, the previous result for v ~< 2 ~< u is actually optimal. It is an open question whether the lp,oc estimate for ()~17(IT)) can be improved to lp also in the remaining cases. LEMMA 21. Let 1 <. p < ec, T E I?.(X, lp) such that T* E 17p(lp,, X*). Then T is psumming with rrp(T) <~7rp(T*). PROOF. Take (Un) c_ X with ep(Un) = 1 and define A:lp, --+ X by en --~ Un and let S := TA E s Ip). By (9), IlAll = Ep(Un) = 1. We find using ]lA*ll = 1 and ep(ej) = 1 in lp,,
Ilrunll p
-
IlSe,
-
I<Se,, e >l"
J 1/p =
Z l ( e n , S*ej)l p n <~ 7rp(S*)ep(ej) <~Jrp(T*).
-
[IS*ejll p j
H. KOnig
964
As an application of the eigenvalue results for p-summing operators, we now determine the asymptotic behaviour of the eigenvalues of nuclear operators. Recall that T 6 s Y) is nuclear, T E N'(X, Y), if there are sequences (X'n) c X*, (Yn) ~ Y with
T - - Z x*(')yn
and
~
nEN
IIx*ll IlYnll < c~.
(20)
nEN
In this case, T has a factorization T :X
Rx "=
R
D
~ lee
> ll
S
> Y,
(x*(x)/[Ix*11)n er~'
S(On)n~N - ~
(21) D(~n)neN-
(llxn* IIIlYnIlffn)n~r~,
OnYn/llYnll
tl
with Ilell IIDII IISII ~ En~N IlXn*II IlYnII. The infimum of all sums on the right side in (20) is the nuclear norm of T, v(T). Any nuclear operator is compact and 2-summing with yr2(T) ~< v(T).
THEOREM 22. Let T E ./V'(X) be nuclear. Then the eigenvalues o f T are square summable, II0~n(T))ll2 ~< v(T). In general this is best possible. If X is an r-convex and r~-concave Banach lattice, where 1 < r <, 2, the eigenvalues of T are absolutely pth power summable for l i p -- 1 - (1/r - 1/2). In particular, for X = Lq(I-2, IX), the eigenvalues o f T belong to lp with 1/p = 1 - I1/q - 1/21 and
II (Z )II The first part goes back to Grothendieck [ 11 ]. For the notion of r-convex and rt-concave lattices see [16]; the result in these spaces is due to Pisier [33]. In his paper, Pisier proves that these spaces are complex interpolation spaces between a Hilbert lattice H and a general Banach lattice X1, X = (X1, H)o, where 0 = 2 / r ~. Any space X = Lq(U2, IX) is a min(q, q~)-convex and a max(q, q ') -concave Banach lattice. It is an open problem whether the eigenvalues of nuclear operators on spaces X of type r with r > 1 are better than 2-summable, i.e., belong to lp for some p < 2. This would follow if, e.g., all T E s X) would be in/-/(s,2)(ll, X) for some fixed s = s(X) < oo. For a discussion of the importance of type and cotype in the local theory of Banach spaces see Maurey [26]. Following [ 15], we will only give the proof in this simpler case X = Lq (S2, IX). PROOF. The first statement follows from Proposition 7 and 7r2 ~ Vl. Now assume that X = Lq(Y2, IX). By duality, we may assume that q ~> 2. By localization it is possible to reduce the proof to the case of X = lq. Thus let T E ./V'(lq) where q ~> 2. Let e > 0. Assume, T has a representation (20) with Z n E N IIx*ll Ily~ II ~< (1 + e)v(T). Let f i n ' - v/llx* II Ily~ II and define D1 "lee --+ 12 and D2"12 --+ ll both by mapping (~n)nEN ~-+ (Crn~n).
Eigenvaluesof operatorsandapplications
965
Let R, D, S be as in (21). Then D - D2D1 and T -- S D2D1R. Here D1 is 2-summing with 7rz(D1) -- lID111 - IIo ll2. Also D~ -- D1. Thus SD2 E/~(/2, lq) has a dual 2-summing and thus q-summing map,
7rq((SD2)*)<~7r2((SD2)*)-
7r2(D1S*) ~< ~2(D~)IIS*II- 11~112IISII.
By Lemma 21, (SD2) is q-summing with 7cq(SD2) ~ 11~11211511.Hence
7rq o 7r2(T) ~
7rq(SD2)Yr2(D1R) <~ IISII Ilcrll211RII
~< Ilcr II2 ~< (1 -t- e)v(T).
We now apply Corollary 19 to conclude that
(~n
I)~n( T) lp
)lip
7rq o ~2(T) ~ (1 + e)v(T),
where 1 / p - 1/2 + 1 / q - 1 - ( 1 / 2 1/q). Concerning the optimality of the first result, given any/z-sequence (Xn) 6/2(Z), there exists f E C[0, 2zr] such that the Fourier coefficients f ( n ) satisfy If(n)l ~> I)~nl, n E Z, cf. [19]. The convolution with f is a nuclear operator in C[0, 2zr] with v ( f , ) <, Ilfll~, having f ( n ) as its eigenvalues. The result in Lq(S2, lZ) is optimal, too, cf. [17, 2.b.15]. D In particular, nuclear operators in Hilbert spaces have absolutely summable eigenvalues. This fact characterizes Hilbert spaces as was shown in [15]" THEOREM 23. Let X be a (complex) Banach space. Then the following are equivalent: (1) Any nuclear operator on X has absolutely summable eigenvalues. (2) X is isomorphic to a Hilbert space. We only have to show (1) =~ (2). To do so, we use the Complemented Subspaces theorem of Lindenstrauss and Tzafriri [23], cf. also [22], which characterizes Hilbert spaces by uniform complementation of subspaces:
COMPLEMENTED SUBSPACES THEOREM. Let X be a Banach space. Then the following are equivalent: (a) Any closed subspace of X is complemented in X. (b) There is c > 0 such that for all finite-dimensional subspaces Y c X, X(Y, X) ~< c. (c) X is isomorphic to a Hilbert space H. Here )~(Y, X) is the relative projection constant of Y in X, )~(Y, X) "-- inf{ IIPII I P E L ( X ) , p2 _ p is a projection onto Y}. PROOF OF THEOREM 23 (Sketch). (1) =~ (2). Assuming (1), we will verify (b) above and thus (c) holds, i.e., X ~ H.
H. K6nig
966
(i) If T E E(X) is an operator of finite rank, T 6 .T(X), the finite nuclear norm vo(T) is defined similarly as v(T), except that in (20) only representations of finite length (n = 1 . . . . . N) are allowed; thus v(T) <<.vo(T). Let Y _ X be a finite-dimensional subspace of X and j " Y ~ X be the imbedding map. We first claim that
)~(Y,X) = sup{ltr(S)l I S E E(Y), v o ( j S ) - 1}.
(22)
For the operator j S E E(Y, X) with dim Y < cx~, the integral, nuclear and the finite nuclear norm actually coincide, cf. Pietsch [30, 6.8.3] or [17, 4.b.4]. Hence
vo(jS) =sup{ltr(LjS)l l llL" X -+ YII
~< 1}.
(23)
Consider the "extension norm" of R 6 E(Y) Ilelle -- inf{ Ilell I e ~ s
Y), /~lY -- e}.
With R = Lj ~ Z;(Y), (23) implies
v o ( j S ) - sup{ Itr(RS)lle E s
~ 1}.
Ilell
Thus vo(j.) and I1" lie are norms on the finite-dimensional space E(Y) which are in traceduality. As a consequence, we have dually Ilelle -sup{Itr(RS)l l S ~ s For R -
vo(jS) <. 1}.
Iy, we find X(Y, X) --IIIylle = sup{ltr(S)l I S E E(Y), vo(jS) ~< 1}.
(ii) We now show that if there is c > 0 such that for all T 6 U(X) I)~n(T)l ~ cvo(T),
(24)
/7
then X is isomorphic to a Hilbert space. Take Y _ X finite-dimensional and let i ' Y X be the inclusion map. Let S ~ s By definition of vo(jS), given e > 0, there is a representation N
jS - Z n=l
N
y* (')Xn,
y* E Y*, Xn E X, Z
IlYn~II Ilxn II ~ (1 + e)vo(jS).
n=l
~ with IIxn9 I I - IlYn II. Then Extend Yn* E Y* to x n E X 9 , Xn~: Iv = Yn,
N
n=l
Eigenvalues of operators and applications
967
with vo(S) ~ (1 + e) vo(j S). Hence by (22)
)~(Y, X) --sup{
~x~(s)
[ vo(S) ~< 1, S l Y - j S } .
k
Since the eigenvalues of S in Y are among the eigenvalues of S in X, assumption (24) implies that X(Y, X) ~ c where c is independent of Y ___X. Thus by the Complemented Subspaces Theorem, X is isomorphic to a Hilbert space. (iii) Finally, we verify, that assumption (1) of Theorem 23 implies that (24) holds. Let
c(X) "- inf, c > 01 ~
IX/7(T)I ~< cvo(T) for all T E ,T'(X) }.
/7
We want to show that c(X) < ~ . If c(Z) < cx~ for some subspace Z __ X of finite codimension, Z "" H1 by (ii) and hence also X ~ H2. Thus if c(X) -- oo, the same is true for any subspace Z of finite codimension, c(Z) - oo. Suppose c(X) - cxz. Choose T1 6 U(X) with Y~n [)~n(T1)[ ~/22v0(T1). Looking at finite representations of T1 almost attaining vo(T1), one finds X1 ___X, dim X1 < oo with
TI(X) ~ X1,
~ I~n(T~)l~ 22v0(T1), /7
where T1 E/~(Xl) is the restriction and astriction of/'1. Choose Z1 ___X, codimZ1 < cx~ such that Ilx + zll ~> 1/2llxll for all x 6 X1, z ~ Z1. Since c(Z1) < oo, there is X2 ~ Z1, dimX2 < oo and T2 6 U(Z1) with
T2(Z1) c X2, Z IZ~(T2)I~ 24v0(T2). /7
In general, choose Zk ___ X, codimZk < oo such that Ilx + zll ~> 1/2llxll for all x E X1 @ . . . | Xk and z E Zk and find Xk+l C_ Zk, dimXk+l < c~ and Tk+l ~ U(Zk) with
I~.(Tk+l)l
Tk+l (Zk) ~ Xk+l,
> 2 2k+2
vo(Tk+l).
/7
By normalization, assume that vo(Tk) -- 1 for all k. Define
kEIN
kEN
kEN
kEN
for Xk E Xk. Then T is nuclear and hence has a nuclear extension T to all of X. But the eigenvalues of T contain those of 2-kTk for all k 6 1N. Thus ~ n ])~n(T)] -- oo, contradicting the assumption (1) in Theorem 23. D
H. K6nig
968
The slightly weaker assumption that nuclear operators X have eigenvalues which are of order [)~n( T ) I - O(n -1) leads to the notion of weak Hilbert spaces, see [10]. PROBLEMS.
(1) Let 0 < p < cx~, 0 < q ~< co. Characterize those Banach spaces X such that all nuclear operators on X have absolutely pth power summable eigenvalues (more generally, that the eigenvalues belong to lp,q). (2) Let 0 < p < c~ and Cp(X, Y) denote the set of all operators T :X --+ Y such that there is c > 0 so that for all operators S: Y ~ X of finite rank
)l/p I)~n( SZ) lp
~ cllSII.
(25)
ncN Let ep(T) "-- inf{c > 0l (25) holds}. Is (C~p,ep) a quasi-Banach ideal? (3) With the same notion as in (2), suppose that A and/3 are operator ideals with A _ Cp and/3 ___Ep. What can be said about A + 13? We end this section with a few comments on traces and determinants for operators in Banach spaces 9 Nuclear operators on Banach spaces X, where X has the approximation property, T E A/'(X), admit a well-defined (matrix) trace given by trm (T) -- ~n~N x*(yn) if T has the form (20) (with X - Y). If X does not have the approximation property, (matrix) traces may not be well-defined in the sense that they depend on the representation (20) of the operator. If some operator ideal A _ A/', like S1, has the property that all operators T E A(X) are Riesz operators with absolutely summable eigenvalues for all Banach spaces X, one may define a (spectral) trace trs "A ~ C by letting trs(T) -- Y~n~r~)~n(T). Trace formulas then claim that trm (T) = trs (T) for all T e ,A(X). For X = H and ,A -- .N', this is Lidskii's formula [20], for A = S1 (with respect to the approximation numbers) see [ 17]. Under the condition on A given above, the trace formula is true in general, as shown recently by White [35]. The main problem is to show that the function Em )~m(')" .A(X) ~ C is linear. In the proof, often determinants of operators of infinite rank are used. If the eigenvalues of T E A(X) are absolutely summable, one introduces
D(~,) "--det(I + ~,T)"-- H (1 + X)~j(T)) jeN
and uses that trs(T) -- D'(O)/D(O). The decay of the eigenvalues ~j(T) is reflected in the growth properties of D and vice-versa. Classically, growth properties of determinants have been used to estimate the asymptotic behaviour of eigenvalues, e.g., by Hille and Tamarkin [ 13]. Even before, determinants were introduced for infinite matrices of the form A - I + T where T - (tij), Y~i supj Itijl < CX),by Hill (1877) and von Koch (1901) as the limit of determinants of finite sections of A. Von Koch also showed that
1
~
~
detA-l+E~Z...Zdet~ nEl~
kl=l
{
tklkl
"""
tklkn )
"
kn---1
tknk 1
...
tknkn
Eigenvalues of operators and applications
969
This approach was transferred by Fredholm (1903) to continuous kernels K defined on the unit square, det(I + T/()
-1+~~
l f0, f01 ...
det
K(tl tl)
""
K(t~,tl)
...
K(tl,tn)) 9
dt| .. 9dtn.
K(tn, tn)
Historical remarks concerning traces and determinants can be found in [32].
4. Applicationto integraloperators We apply Theorems 13 and 18 for Weyl number ideals and p-summing operators to estimate the order of decay of the eigenvalues of certain types of integral operators where the kernel satisfies summability and/or regularity assumptions. There is, of course, a vast literature on the asymptotic distribution of the eigenvalues of integral operators 9 The results usually rely on the estimation of singular numbers in Hilbert spaces 9For a recent study, see Edmunds and Evans [8]. For the distribution of the eigenvalues of random Gaussian matrices see the article of Davidson and Szarek [6]. We present results where the Banach space methods of the previous chapters are essential. We start with the Hille-Tamarkin operators 9 PROPOSITION 24. Let (S2, lZ) be a a-finite measure space, 1 < p < ec and k : S2 x S-2 -+ C be a measurable kernel such that
1/p ]]kllp,p'
9~
]k(s, t)]
p!
d/z(t)
d/z(s)
< (X).
Then Tk f (x) 9 fs? k(x, y) f (y) dlz(y)
(26)
defines a power-compact operator Tk : Lp(S2, lZ) --+ Lp(S-2, lZ) with (Xn(Tk)) E lq, where q = max(p, 2) and II)~n(Tk)llq ~ Ilkllp,p,. PROOF. The fact that Ilk!lp,p, < ~ means for a-finite measure spaces that f :$2 --+ X := Lp,(S-2, lZ) given by f ( s ) ( t ) := k(s, t) belongs to the space of Bochner integrable functions Lp(S-2, lZ; X). Lemma 25 below shows that Tk E 1-Ip(Lp(S-2,/z)). Apply Theorem 18. [] LEMMA 25. Let (s
#) be a measure space, 1 <~ p < oo and f ~ Lp(#; X). Then Tf :X* --+ L p ( ~ , #) given by x* ~ X* ~ x* o f E L p ( ~ , #) is p-summing with 7rp(rf ) <~ Ilfllg~(n,,;x).
970
H. KOnig
PROOF. Let E = {co E S2 I f(co) r O} and (x*) n
i = 1
c X* Then
--
"
1/p IlTfx*ll p
-- ~
i=1
Xi
[[f(w) ll
IIf(co) IIp d# (co))
i=1
<. IlfllLp(n,r,;x) " 6p(X?).
Hence Tf is p-summing with ~ p ( Z f ) ~ IlfllZp(n,~;S). A corollary is the well-known Hausdorff-Young inequality. COROLLARY 26. Let 2 <. p < oo and f E Lp,(O, 27r). Then f E Ip(Z) with Ilfllz~ ~ IlfllL/. PROOF. The convolution operator with the (2zr)-periodic extension f of f has kernel k(s, t) = f ( s - t) with Ilkllp,p, - IlfllLp,. The eigenvalues of f , are just the Fourier coefficients f .
D
For weakly singular integral operators, the following is true. PROPOSITION 27. Let 12 c ]l~N be a bounded domain, k" X22\ A --+ C be a measurable kernel given by k(s, t) - l(s, t)/ls - tl N-u, with I space (a) (b) (c)
A "-- { ( x , x ) l x E I-2} and
0 < ot <, N,
E Lee(S22). Then Tk as in (26) with respect to Lebesgue measure is compact in any Lr (I2), 1 <. r <. oo with eigenvalues satisfying (Xn(Tn)) E 1U/~,oo if ~ < N / 2 , ()~n(Tn)) E 12 if ot > N / 2 , IXn(Zn)l <. c(ln(n + 1)/n) 1/2 if ~ - N / 2 .
PROOF (Idea). If ot > N / 2 , Tk ~ H2(Loc(12))./2(Loc(X2)) factors over the identity map I : Loc(I2) --+ Lp,1 (S2) where p := N / u > 2 and Lp,1 (S2) is the Lorentz function space. One has to check that I is (p, 2)-summing to be able to use Corollary 16. For ot = N / 2 , one estimates the Weyl numbers of I : L~(S2) ~ L2,1 (S'2) by a limiting process of p "a 2 in the previous situation, showing that Xn(I) ~ c(ln(n + 1)/n) 1/2. The results are optimal. For details, see [17, 3.a. 11 ]. El We now turn to integral operators satisfying regularity assumptions which are formulated in terms of Sobolev-Besov spaces. For kernels in 2 variable sets, we need to introduce vector-valued Besov spaces. Let S-2 c R N be open, r E N0, 1 ~< p < c~, and X be a Banach space. Using the standard multiindex notation for derivatives, we introduce the vector-valued Sobolev-spaces Wp ([2; X ) " - { f
E
Lp(ff2; X) ID~
E Lp([2; X) for all o~ E Ny with
I~1 ~ r}
Eigenvalues of operators and applications
971
p with n o r m I]f]lp,r.X "- (Y~lc~l~
O.)l(f, t)p "--
+ h E n for all h E R N with IIhll~ < t},
sup IIf(" + h) - f(')llLp(n(t);x); Ilhll~
col (f, .) is the modulus of continuity of f in Lp. Now let additionally 1 ~< q ~< oc and ;~ > 0. Put r " - [)~] and 0 " - ,k - r. If )~ r N, 0 > 0, the vector-valued Besov-spaces are defined by
Bp,q(n; X)"--
I
f E w p ( n ; X)
i
Iflp,q,,~;x "- sup Ic~l=r
(fo'(~
7
<
with norm Ilfllp,q,~;x : : Ilfllp,r;x + Iflp,q~;x. Thus the distributional derivatives of f of order I~1 - r satisfy some integrated H61der condition. For )~ E 1% a slightly different definition using the second modulus of continuity has to be used, cf. [ 17]. In our application, X will be another Besov space X -- B~, v ( n ) ; then
f EBp,q can be identified with a kernel of two sets of variables, k(s, t) = f ( s ) ( t ) ; s, t E n . In the scalar case, there are well-known interpolation theorems relating Sobolev- and Besov-spaces. They also hold in the vector-valued case. We have in terms of the real interpolation method
x), w (n; X))o,q
- 8 ,,q (n., x),
(27)
isomorphically, if 0 < )~ < m and 0 -- )~/m, provided that 0 n is not too irregular; cf. [ 17, 3.b.7]. The decay of the eigenvalues of integral operators defined by Besov-type kernels k E Bp,q (B~u,v) can be analyzed using the eigenvalue theorem for the Weyl number ideals St,v. THEOREM 28. Let 1 <, p , q , u , v < oo, )~,cr > O w i t h ) ~ + ~ > N ( 1 / p + i / u - 1) and ,.(-2 C R u be sufficiently regular. Define t by 1/t "-- O~ + cr)/N + 1 / m a x ( 2 , u'). Let k E B p,q )~ ( n ; Bu~v ( n ) ). Then the eigenvalues of the integral operator Tk defined by (26) belong
to the Lorentz sequence space lt,q with
The constant c depends only on the indices and n . The space lt,q is the best possible, i.e., smallest among the Lorentz sequence spaces.
H. KOnig
972
REMARK. The integral operator T~ is well-defined, e.g., in any space Ls (I2) where s satisfies )~ > N ( 1 / p - 1/s), cr > N ( 1 / s - 1/ut), as follows from the Sobolev-Besov imbedding theorems, or in the spaces Bp,q (I2) or Bu~, (s It is a consequence of the principle of related operators that the eigenvalue sequence is independent of the choice of these spaces 9 It is interesting to note that the space It,q does not depend on the indices p and v. v
9
PROOF (Ideas). The proof relies on the estimate of the Weyl numbers of Tk in suitable function spaces and is fairly technical. We only illustrate the basic techniques in a simple case of indices, namely if u I = p. Any f ~ B p,q (S2", X) defines an operator Tf" X* ~ Bp,q (S-2) by mapping x* ~ X * o f . Here Tk "~ -- Tf " Bu~,v(S2) * --+ Bp,q ~ (S2) . If )~ + cr > N ( 1 / p + 1 / u inclusion map I" Bp,q ( n ) ~
1), as assumed, the
Bu,~ v ( n ) *
exists, 9 actually, I = I~ o 11 where I i ' B p ,~q (f2)--+ Ls(f2) and /2" BuOy(n)--+ Ls,(f2) where s is determined as in the above remark. Hence T~'Bp,q (~2) --+ Bp,q z (f2) factors as T~ -- T~ I. A more complicated analogue of L e m m a 21 derived by interpolation using (27) shows that T~ is actually p-summing if p -- q. In this case, )~ ) ~ Clrt-1/1 rcp(Tl~) "~ <~ cln - 1 / l Ilklls~,q(8~.~), Xn(fl~" Bu~,* --+ Bp,q
where 1 = max(p, 2). For u f -- p, we may take s = p in I1, 12 and the Weyl number of these imbeddings may be estimated by the approximation numbers and these by classical theorems of approximation theory
~ ~ x n ( l l " Bp,p --+ Lp) ~ an(I1 " Bp,p
L p ) <~ c z n -~./N ,
x n ( l ~ ' L p --+ Bp,*p,) <. an(I2" Bp,,p,--+ Lp,) <. c2n -~r/N. The multiplicativity of the Weyl-numbers then implies that Xn (I) = O(rt -(~+cr)/N) and
xn(T~) -- xn(T~I) ~ c3n-((z+~)/N+l/t)llklls~,p(B~,v). This means that k ~ T~ induces a continuous linear map ~ (n; 99 9Bp,p
Bu~ v (n))~s,~(x*) , ,
(28)
where X - BuOy(I-2) (or X - - Lu(I2)) and 1/t -- O~ + c r ) / N + 1/1. To improve the second index ec to q in (28) and treat the case of p # q, one may interpolate (28) for different orders of differentiability )~. The Weyl number ideals interpolate according to
(s,,,~(x * ) , s,:,~(x*))o,q_C
St,q ( X * )
(29)
Eigenvalues o f operators and applications
973
where 1/t = (1 - O ) / t l -k-O/t2. Using (27) and (29) as well as (28) for two different orders )~l, )~2, the reiteration theorem of the real interpolation method yields that k ~ Tk induces an operator
(n., ,,+, (n))
+)
Hence the eigenvalues belong to It,q by Theorem 13. For details see [17, 3.d.5]. In more complicated cases of indices (p, u), classical approximation theory and a method of Maiorov [24] can be used to discretize the problem and reduce it to the estimation of the Weyl numbers of identity maps l pm ---->Ium . The order of these may be determined, see, e.g., the arguments used in the proof of Proposition 20. A different method has been used by Pietsch [31, I]. Extending the basis constructions of Ciesielski and Figiel [5] in Besovspaces to the vector-valued case, he derived isomorphisms between the Besov-spaces and weighted Ip(lq)-type sequence spaces; between these Weyl number estimates then can be derived. D Pietsch succeeded to extend Theorem 28 to all of R N, if additional weights of the type (1 -Jr-Ixl2) c~/2 are introduced. Though this increased the number of parameters by 2 (or and/~), he was able in [31, II] to determine the optimal Lorentz space except for some singular index combinations. One may also estimate the entropy numbers of integral operators by finding the entropy asymptotics of Sobolev-Besov imbeddings and then use the Carl-Triebel Theorem 14 to derive eigenvalue estimates, see Carl and Stephani [3, Chapter 5]. Applying Theorem 28 to convolution operators, one finds tOc~(X) :
COROLLARY 29. Let 1 <<,u, q < cx~, )~ > 0 and f E B u,q((-Tr, ~ 7r)U). Define t by 1/t : )~/N + 1/max(2, u'). Then f E lt,q(Zn). Here the Besov-space should be taken with respect to a periodic modulus o f smoothness.
Examples show that Corollary 29 and thus Theorem 28 cannot be improved in the Lorentz sequence space setting. Further references on the topic of this paper can be found in [3,17,32]. The latter book also contains a historical survey.
References [1] G. Bennett, Inclusion mappings between lP-spaces, J. Funct. Anal. 13 (1973), 20-27. [2] B. Carl, Absolut (p, 1)-summierende identische Operatoren von lu nach lv, Math. Nachr. 63 (1974), 353360. [3] B. Carl and I. Stephani, Entropy, Compactness and the Approximation of Operators, Cambridge Univ. Press (1990). [4] B. Carl and H. Triebel, Inequalities between eigenvalues, entropy numbers and related quantities of compact operators in Banach spaces, Math. Ann. 251 (1980), 129-133. [5] Z. Ciesielski and T. Figiel, Spline bases in classical function spaces on compact C~-manifolds, Stud. Math. 76 (1983), 1-58 and 95-136. [6] K.R. Davidson and S.J. Szarek, Local operator theory, random matrices and Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 317-366.
974
H. KOnig
[7] J. Diestel, H. Jarchow and A. Pietsch, Operator ideals, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 437-496. [81 D.E. Edmunds and W.D. Evans, Spectral Theory and Differential Operators, Clarendon Press (1987). [9] D.J.H. Garling and Y. Gordon, Relations between some constants associated with finite dimensional Banach spaces, Israel J. Math. 9 (1971), 346-361. [10] E.D. Gluskin, Y. Gordon and A. Pajor, Entropy, approximation numbers and other parameters, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [11] A. Grothendieck, Produits tensoriels topologiques et dspaces nucl~aires, Mem. Amer. Math. Soc. 16 (1955). [12] G.H. Hardy, J.E. Littlewood and G. Polya, Inequalities, Cambridge Univ. Press (1934). [13] E. Hille and J. Tamarkin, On the characteristic values of linear integral equations, Acta Math. 57 (1931), 1-76. [14] J.R. Holub, Tensor product mappings, II, Proc. Amer. Math. Soc. 42 (1974), 437-441. [15] W.B. Johnson, H. K6nig, B. Maurey and J.R. Retherford, Eigenvalues of p-summing and Ip-type operators in Banach spaces, J. Funct. Anal. 32 (1979), 353-380. [16] W.B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, Handbook of the Geometry of Banach Spaces, Vol. 1, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001), 1-84. [17] H. K6nig, Eigenvalue Distribution of Compact Operators, Birkh~iuser (1986). [18] H. K6nig, A formula for the eigenvalues of a compact operator, Stud. Math. 65 (1979), 141-146. [19] K. de Leeuw, Y. Katznelson and J.E Kahane, Sur les coefficients de Fourier des fonctions continues, C. R. Acad. Sci. Paris 285 (1977), 1001-1003. [20] V.B. Lidskii, Nonselfadjoint operators having a trace, Dokl. Akad. Nauk SSSR 125 (1959), 485-487. [21] J. Lindenstrauss and A. Pdczyfiski, Absolutely summing operators in 12p-spaces and their applications, Studia Math. 29 (1968), 275-326. [22] J. Lindenstrauss, Characterization of Hilbert space, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [23] J. Lindenstrauss and L. Tzafriri, On the complemented subspaces problem, Israel J. Math. 9 (1971), 263269. [24] V.E. Maiorov, Discretization of the problem of diameter, Usp. Mat. Nauk 30 (1975), 179-180. [25] E. Makai and J. Zemanek, Geometrical means ofeigenvalues, J. Oper. Theory 7 (1982), 173-178. [26] B. Maurey, Local theory: history and impact, Handbook of the Geometry of Banach Spaces, Vol. 2, W.B. Johnson and J. Lindenstrauss, eds, Elsevier, Amsterdam (2001) (to be published). [27] A. Pdczyfiski, A characterization of Hilbert-Schmidt operators, Studia Math. 28 (1966/67), 355-360. [28] A. Pietsch, s-numbers of operators in Banach spaces, Studia Math. 51 (1974), 201-223. [29] A. Pietsch, Weyl numbers and eigenvalues of operators in Banach spaces, Math. Ann. 247 (1980), 149-168. [30] A. Pietsch, Operator Ideals, North Holland (1980). [31] A. Pietsch, Eigenvalues of integral operators, I, II, Math. Ann. 247 (1980), 169-178 and 262 (1983), 343376. [32] A. Pietsch, Eigenvalues and s-numbers, Cambridge Univ. Press (1987). [331 G. Pisier, Some applications of the complex interpolation method to Banach lattices, J. d'Anal. Math. 35 (1979), 264-281. [341 H. Weyl, Inequalities between two kinds of eigenvalues of a linear transformation, Proc. Nat. Acad. Sci. USA 35 (1949), 408-411. [35] M.C. White, Analytic multivalued functions and spectral trace, Math. Ann. 304 (1996), 665-683.
Author Index Roman numbers refer to pages on which the author (or hidher work) is mentioned. Italic numbers refer to reference pages. Numbers between brackets are the reference numbers. No distinction is made between first and co-author(s)
Abramovich, Y.A. 87, 88, 90-94, 99-105, 108, 110. Amemiya, I. 89, 118 [30];440,490 [ 2 ] 111, 116,117 111: 117 121; 117 131; 117 [41; Amir, D. 744, 745, 770, 772 [9]; 772 [lo]; 793, 820. 117 151: 117 161;117 171: 117 181; 117 PI; 822,830 151; 830 [61 118 rioi; 118 1111; 118 1121; 118 ~131;118 ri41; Andersen, N.T. 341.360 131 118 1151; 118 ~161:118 ~171;118 1181;118 1191: Anderson, J. 332-335, 360 [4]; 360 [ 5 ] :360 [6]: 118 1201; 118 ~211;535,558 rii 360 [7] Aharoni, I. 829, 830 111: 830 121; 906, 935 [11 And& T. 110, 118 1311: 130, 147, 156 1101; 255, Aizenman, M. 354,360 [ 1 1 265 131: 904,935 131 Angelos, J.R. 320,360 [8] Akilov, G.P. 106, 107, 120 1741 Ansari, S. 545,547,558 121 Al-Husaini, A.L. 905, 935 121 Ansel. J.P. 385,390 [I] Aldous, D.J. 136, 156 111; 237,265 [I]; 515, 526, Antipa, A. 257,265 (41 527 [I] Apostol, C. 330, 332,360 [9]: 360 1101; 360 [I11 Alencar, R. 812,830 131 Arai, H. 704, 704 121 Alesker. S. 714, 731-735, 772 [I]; 772 [2]; 772 [ 3 ) ; Arenson, E.L. 87. 118 1151 772 [4); 772 15) Argyros, S. 139, 154,157 [ I l l ; 157 [121: 822,830 [71 Alexander, H. 674, 675, 704 [ I ] Arias-de-Reyna, J. 716, 772 [ I I ] Alexandrov, A.D. 727, 731, 772 [6]; 772 [7] Amold, L. 343.360 [12] Alexopoulos. J. 515, 528 [2]; 528 [31 Aron, R.M. 555, 558 [ 3 ] ;676, 704 131; 812, 830 131 Alfsen, E.M. 310, 313 [I]; 611-614, 620, 621. 627, Aronszajn, N.535.558 141 665 [11; 665 121; 665 131 Arveson, W.B. 339, 340, 360 [13]; 360 1141: 535, Aliprantis. C.D. 21-24.83 [I]: 87-90.92-97.99-105, 558 [S]; 558 [6] 108, 110, 111, 113, 1 1 6 . 1 1 7 [ 4 ] : 1 1 7 [ 5 ~ ; 1 1 7 ~ 6 ] ; Ash, M. 215,230 [I]; 230 121 117 171; 117 p i ; 117 PI; 118 1101: 118 [ i l l ; Asimov, L. 615.6 18.620, 62 1, 626, 627, 665 141 118 ~121;118 1131:118 [MI; 118 1221: 118 [231; Asmar. N.H. 249,265 151; 265 [6] 118 1241: 118 ~251;118 1261: 118 ~271;118 1281: Asplund. E. 663.665 151; 792,798,805,828,830 181; 118 ~291;535,558 111 830 [91 Allekhverdiev, D.E. 453, 490 [ I ] Aubin, J.P. 433 [I]; 798. 830 [lo] Allen, G.D. 527, 528 [4] Azagra, D. 409,422,433 121; 433 131: 799, 830 [ I I] Allexandrov, G. 304,313 121 AZOK,E. 328,360 risl Alon, N.358. 360 121; 770, 772 181 Alonso. J. 793,830 [4] Bachelier, L. 369,390 [2] Alspach, D.E. 59,83 [22l; 133, 135. 147. 151, 154, Bachelis, G.F. 204,230 [31; 884. 896 [2l 156,156 121; 156 131; 156 [41; 156 151; 156 [61; Bachir, M. 411, 416,418, 433 [4l; 433 151 156 [71; 156 181; 279,304,313 131; 313 141; 581, Baemstein, A. 259,265 171: 444,490 131 595 111; 839, 862,868,868 [I]: 875.896 [I] Bai, Z.D. 344, 348, 353, 358,360 1161: 360 1171; Altshuler, Z. 134, 156 [9]: 51 1, 527, 528 [ 5 ] ;528 [6] 360 [IS]; 360 1191: 361 [20]: 366 11871 Amann, H. 250,265 121 Bakhtin, LA. 88. 118 1321
915
976
Author Index
Bakry, D. 350,361 [21] Ball, K.M. 163, 165, 171, 177, 183, 185, 187, 192, 193 [I]: 193 [2]: 193 [3]: 193 [4]: 193 [5]; 716, 718, 722,724,725,728,772, 772 [l I]; 772 [12]: 773 [13]: 773 [14]; 773 [15]; 773 [16]; 773 [17]: 773 [18]; 901,918,935 [4]: 935 [5] Banach, S. 273,313 [5]; 444,490 [4]; 524,528 [7] Banaszczyk, W. 767, 773 [19] Bang, T. 183,193 [6] Bafiuelos, R. 259,265 [8]; 265 [9]; 265 [lo] Bapat, R.B. 98, 113,118 [33] Barany, I. 175, 193 [7] Barles, G. 426,431,433 [6] Barthe, F. 164, 171, 173, 178, 193 [8]: I93 [9]; 718, 773 [20]; 921, 935 [6] Bastero, J. 519, 528 [81: 528 [9]: 769, 773 [21] Bauer, H. 62 1,665 [6] Beauzarny, B. 6,7,52,57,83 [2]; 444,474,479,480, 490 [51; 490 [61; 491 171; 549,550,555,558 [31: 558 [71; 558 [81; 558 [9]: 792, 804,830 [I21 Becker, R. 613,665 [7] Beckner, W. 481,491 [S] Bellow, A. 260, 265 11 11 Ben Arous, G. 345,346,361 [22] Benedek, A. 107, 118 1341 Benitez. C. 793. 830 141; 830 [131 Bennett, C. 505, 515,528 [lo] Bennett, G. 147, I57 1131: 230 141; 342,361 [23]; 463, 491 [9]; 491 [lo]: 748, 773 [23]: 866, 867,868 [2]; 962, 973 [ 11 Benveniste, E.J. 331, 359, 361 [24] Benyarnini, Y. 35, 36, 38,42, 48, 83 [3]: 279, 313 [6]; 342, 354,361 [25]: 630,634,665 [8]; 766, 773 [221: 822, 829,830 [14]; 830 [151; 906, 935 [71 Bercovici, H. 341, 361 [26] Berg, I.D. 323-325, 328, 329, 331, 361 [27]; 361 [28]: 361 [29] Bergh, J. 74, 76-78, 80,83 [4]; 505, 528 [ 111; 577, 595 [2]: 692, 704 [4] Berkson, E. 237,249,250,265 [51; 265 [12]: 265 [13] Berman, A. 98, 113,118 [35] Berman, K. 336,361 [30]: 859,868 [3] Bernstein, A.R. 535, 558 [lo] Bernstein, S. 470,491 [ 111 BernuCs, J. 769, 773 [21] Bessaga, C. 603,648,665 [9]; 665 [lo]: 792,799, 830 [161: 830 [I71 Bhatia, R. 327,329, 347,361 [31]: 361 [32]: 361 [331: 361 [341; 361 [351: 361 [361 Biane, P. 360, 361 [37] Billard, P. 569,595 [3] Bishop, E. 608, 610, 640, 641, 665 [ll]; 665 [12]: 665 [ 131 Bjork, T. 374, 390 [3]
Black, F. 369, 371, 390 [4] Blasco, 0. 250, 265 [14]; 265 [15]: 894, 896 [3] Blower, G. 264,265 [16] Boas, R.P. 156,157 [14] Bobkov, S.G. 350,358,361 [38] Bochner, S. 265 [17]; 480,491 [12] BoEkariov, S.V. 569,575,595 [4]; 595 [5]: 595 [6] Bohnenblust, F. 140,157 [IS]: 465,491 [13] Bolker, E.D. 525,528 [12]; 768, 773 [24]: 911, 921, 924, 935 [8] Bornbieri, E. 207,230 [S]: 555,558 [9] Bornze, I.M. 112,119 [43] Bonic, R. 413,433 [7]; 799, 830 [18] Bonsall, F.F. 608, 665 [14] Border, K.C. 87, 100,118 [22] Borell, C. 346,361 [391: 717, 773 [251 Borovikov, V. 614,665 [15] Bonvein, J.M. 396, 399,418,433 [Sl; 433 [9]: 664, 665 [161; 798, 820,830 [I91 Bossard, B. 793, 805, 828,830 [20]; 830 [21]: 830 [22] Bourgain, J. 139, 147, 150, 154, 157 [16]: 157 [17]: 157 [181: 157 [191; 157 [20]; 157 [211: 175, I93 [12]: 201,204, 206, 209, 213, 214,219-221, 223, 226, 229, 230,230 [6]: 230 [7]; 230 [8]: 230 [9]; 230 [lo]: 230 [ I l l : 231 [12]; 231 [13]; 231 [141; 237,245,247,250,265 1181: 265 [19]; 265 [201: 265 [21]: 265 [22]: 336, 337, 358, 359, 361 [40]; 361 [41]: 361 [42]: 361 [43]: 361 [MI; 445,465,468,482,491 [14]; 491 [IS]: 491 [16]; 491 [17]; 576, 580, 581, 590, 592, 595 [7]: 595 [8]; 595 [91; 595 [ 101; 595 [ I 11; 633, 634, 636, 658, 665 [171; 665 [18]; 665 [19]: 665 [20]; 675, 686, 687,699,703,704, 704 [5]; 704 [6]; 704 [7]: 705 [81; 705 [9]: 705 [lo]: 705 [ll]: 718,724,736, 742,755,759,763, 766772, 773 [26]; 773 [27]: 773 [28]: 773 [29]; 773 [30]; 773 [31]: 773 [32]: 773 [33]; 773 [34]; 773 [35]; 773 [36]: 773 [37]; 773 [38]; 773 [39]: 773 [40]: 805,830 [23]: 844, 845, 853, 854, 859, 860, 862-867,868 [4]: 868 [5]; 868 [61; 868 [71: 868 [8]; 868 [9]: 868 [lo]; 868 [ I l l ; 868 [12]; 868 [13]: 884, 885, 891, 892, 895, 896,896 [41; 896 [5];897 [61: 897 [71: 897 [81: 918,925,927,935 [9]; 935 [lo]; 935 [ll]: 935 [12] Bourgin, R.D. 260,265 [23]: 633,634, 665 [21]; 665 [221: 793,830 [241 Boutet de Monvel, A. 349,361 [45] Boyd, D.W. 514,528 [13] Brascarnp, H.J. 164, 193 [lo]; 718, 773 [41] Bratteli, 0. 620, 665 [23] Braverman, M.Sh. 523,528 [14]: 528 [15] Brenier, Y. 173, 193 [Ill: 713, 773 [42]
Author Index Bretagnolle, J. 140, 157 1221; 524, 527, 528 1161; 528 1171; 855,868 1141; 906, 935 1131 Bronk. B.V. 344, 361 1461 Brown, D.J. 113, 116. 118 1231; 118 [24] Brown, L.G. 323, 331, 361 1471; 361 1481 Brown, S.W. 341,362 [49]; 543,558 1111 Brudnyi, Yu.A. 74, 83 [S] BU, S.Q. 139, 157 1231;263,265 1241 Bukhvalov, A.V. 87, 89, 96, 106, 108, 118 [37]; 118 1381; 119 1391; 119 1401; 119 ~411; 263, 265 1251; 265 1261; 265 1271: 638, 665 [24] Burago, Y.D. 710. 712, 726, 774 1431 Burger, M. 935 [14] Burger, R. 112, 119 1421; 119 1431 Burkholder, D.L. 10, 83 1231; 126. 128, 157 [24]; 157 [251; 231, 238, 241-246. 248-251. 253, 254, 256-259,262,266 1281;266 1291;266 1301; 266 ~311;266 1321; 266 1331; 266 1341;266 1351: 266 1361: 266 1371: 266 [381; 266 1391; 266 ~401; 266 1411;266 ~421;266 1431: 266 ~441;471.482, 491 1181; 491 1191;491 1201; 491 [ ~ I I491 ; ~221; 523,528 1181;590,595 ~121; 895.897 191; 901, 906, 935 [15] Burkinshaw, 0. 21-24.83 [ I ] ; 87-90, 92-97, 99-105, 108, 110, 111, 113, 117r41; i i 7 r s i ; 1i7[61; 117171; 117[8]; 117[9); 118(10); 118[11]; 118 1121; 118 1131; 118 1231; 118 1251; 118 [261; 118 1271; 118 [2s1; 118 1291; 535,558 [ I ] Busemann, H. 177, 193 [13]; 918. 935 1161 Cabello Sanchez, F. 805, 830 [25] Caffarelli, L.A. 714, 774 1441 Calderh, A.P. 245,266 [45]; 266 1461 Calkin, J.W. 439, 491 1231 Canturija, Z.A. 579,596 [I81 Caradus, S.R. 558 1121 CarathCodory, C. 712. 774 [451 Carl, B. 175, 193 [141; 452,463,491 [24]; 491 [25]; 854, 868 [IS]; 958, 959,962. 973. 973 121: 973 [ 3 ] ; 973 141 Carleson, L. 223, 231 [IS] Carothers, N.L. 135, 147, 156 [7]; 304,313 131; 504, SOS,SI 1,523, s2s-s27,528 1191; 528 1201; 528 [211: 528 1221; 528 [231; 528 1241; 528 1251; 528 ~261; 528 1271; 528 1281;528 1291; 528 1301; 528 [ ~ I I 528 : (321; 901.935 (171 Cartier, P. 610, 665 1251 Cartwright, D.I. 92, 119 [44] Casazza, P.G. 12. 14.58, 60,83 1241: 133, 136, 140, 157 1261; 157 1271;157 1281; 276,279,285, 286, 291-294,297-299,302, 304, 309.31 1, 313, 313 171; 313 rsi: 313 191; 313 1101; 31-7 1111; 313 1121; 313 1131; 313 114); 313 [151; 313 [16];
977
313 1171;313 risi; 511, 527,528 161; 529 ~331: 529 r341; 812,821,830 1261; 830 ~ 2 7 1 Caselles, V. 110, 119 1451; 119 1461 Cepedello, M. 813, 828, 830 1281; 830 [29]; 831 1301; 831 1311 Chang, K-C. 433 [lo] Chatterji, S.D. 260, 266 1471; 266 I481 Chen. S.T. 515, 529 1351; 529 1361; 529 I371 Chen, Z.L. 97, 119 [47l Cheridito, P. 383, 390 151 Chevet, S. 354,362 [50]; 461,491 1261 Cho, C. 310,313 1191 Choi, C. 259, 266 1491; 266 1501 Choi, M.D. 101, 119 [48]: 323,362 [SI] Choquet, G. 60.5, 608.61 I , 613, 614, 629, 665 [26]; 665 1271; 665 [2X]; 666 1291; 666 1301; 666 1311; 666 1321; 666 [33]: 793, 831 [32] Choulli, M. 433 [ I 11 Choulli, T. 389, 390 161; 390 [7] Christensen, E. 340, 362 1521 Christensen, J.P.R. 624, 666 1341 Ciesielski, Z. 575, 583, 585-587, 595 [13]; 595 1141; 595 1151: 595 [16]: 595 1171; 973, 973 [S] Clarke, F. 412,423,425, 433 [12]; 433 1131: 434 1141; 434 [IS]; 434 1161 Clarkson, J.A. 128, 157 1291; 484,491 1271 Clement, P. 245, 266 [Sl] Cobos, F. 266 [52] Cole, B.J. 676, 682, 704 131; 705 1121 Comes, B. 215,231 1161 Contreras, M.D. 642, 666 1351 Cooke, R. 214,231 1171 Corson, H.H. 624, 629. 662, 666 [32]: 666 1361; 666 1371 Corvellec, J-N. 425,434 [I71 Coulhon, T. 250, 266 [531 Cowen, C.C. 320,360 181; 471,491 1281: 491 1291 Crandall, M.G. 420, 421,426. 431,434 [IS]; 434 [19]; 434 [20] Creekmore, J. 505, 529 1381 Cwikel, M. 505, 529 [39] Dacunha-Castelle. D. 140, 157 [22]: 157 [30]; 455, 491 [301; 524, 527,528 [16]; 528 1171; 855, 868 1141; 906, 935 1131 Dadarlat, M. 332,362 1531 Dalang, R.C. 374,390 [Sl Dancer, E.N. 112, 119 1491 Danilevich, A.A. 263, 265 [27]; 638, 665 1241 Dar, S. 714, 724. 731, 732, 763, 772 [ S ] ; 774 [46]; 774 [47]; 774 1481 Daubechies, I. 566, 578,596 [ 191 David, G. 250,266 [541; 591,596 (201; 702, 705 1131 Davidson, K.R. 320, 323-325. 328,331-333, 339-341.361 1291: 362 1541; 362 1551: 362 1561:
978
Author Index
362 [571; 362 [58]; 362 [59]; 362 [60]; 362 [61]: 362 [62]: 362 [63]; 362 [64]: 362 [65]; 362 [66]; 362 [67]: 859,868 1161: 969, 973 [6] Davie, A.M. 283,313 [20]: 313 [21]: 673, 705 [I41 Davis, B.J. 258, 266 [43]: 523, 528 [18] Davis, C. 327-329, 339, 347, 360 [15]: 361 [34]: 361 [35]; 362 [681 Davis, W.J. 96, 119 [SO]; 265 [21]; 266 [55]: 443, 491 [311: 646,666 [381: 743, 766, 774 [491: 792, 793, 821,831 1331: 831 [341: 831 [351; 850, 868 [ 171 Day, M.M. 792,798, 821,831 [36]; 831 [37] de Branges, L. 10.5, 118 [36]; 541,558 [13]; 603, 666 [39] de Figueiredo, D.G. 434 [21] de Leeuw, K. 184,193 [IS]: 608,610, 665 [12]: 965, 974 [19] de Pagter, B. 110, 11 I , 121 [106]: 245.266 [51] de Valk, V. 515,529 [50] Debs, G. 793,831 [38] Deddens, J.A. 339,362 [69] Defant, A. 466,467,484,488,490,491 [32] Defant, M. 590, 596 [21] Degiovanni, M. 425,434 [I71 Deift, P. 349, 362 [70] Delbaen, F. 374, 377-386, 389, 390 [91; 390 [lo]; 390 [ l l l ; 390 [12]: 390 [13]; 390 [14]: 390 [151; 390 [161; 465,491 [16]: 658,665 [18]: 901,905, 91 1-915, 935 1181 Dellacherie, C. 374-376, 390 [17] Deville, R. 33, 34, 57, 83 [6]: 83 [25]: 406, 408,409, 413,415,418422,433 [21: 433 [31; 433 [111: 434 [22]: 434 [23]: 434 [24]; 434 [25]; 434 [26]: 434 [27]: 434 [28]; 434 [29]; 434 [30]; 434 [31]; 476, 491 [33]: 644, 659, 666 [40]: 666 [41]; 792, 793, 795, 798, 799, 805, 812-814, 821, 831 [39]; 831 [401: 831 [41]: 831 [42]: 831 [43]; 831 [MI; 831 [45]: 831 [46]: 831 [47] DeVore, R.A. 575,596 [22] Diestel, J. 13, 18, 35, 36, 3 8 4 0 , 47, 55, 60, 65, 66,72, 83 (71; 83 [8]; 83 [9]; 106, 107, 119 [51]: 259, 260, 263, 266 [56]: 445,455, 458,459,464, 466, 471, 474,479,480,483,484,491 [17]; 491 [34]: 491 [35]; 491 [361: 491 [37]: 518,529 [401; 558 [14]; 675,681,690, 705 [15]; 792,793, 806, 831 1481; 831 [49]; 831 [SO]; 867, 868 1181; 879, 881, 882, 887,897 [lo]: 929, 935 [19]: 943,950, 951, 974 [7] Dilworth, S.J. 504, 511, 518, 519, 523-527, 528 [22]: 528 1231; 528 [24]; 528 [25]; 528 [26]: 528 [27]; 528 1281: 528 [29]: 529 [41]: 529 [421; 529 1431: 529 [441: 529 [45]: 529 [46]: 529 [47]; 529 [48]; 529 [49]; 769. 770, 774 [50]; 774 [55]: 906, 935 I201
Dineen, S. 676, 705 [16]; 812,830 [3] Dixon, A.C. 451,491 [38] Dixon, P.G. 286, 310,313 [22] Dmitrovskii, V.A. 347, 362 [71] Dobrowolski, T. 799, 830 [I 11 Dodds, P.G. 94, 119 [52] DolCans-Dade, C. 389,390 [18] Domenig, T. 471,492 [39] Doob, J.L. 260,267 [57] Dor, L.E. 131, 147, 155, 157 [13]; 157 [31]; 157 [32]; 230 [4]: 255,256,267 [58]; 565,596 [23]: 748, 768, 773 [23]; 774 [51]: 862,866, 867,868 [2]: 868 [19]; 906, 935 [21] Dore, G. 250, 267 [59]; 267 [60]: 267 [61]; 267 1621 Douglas, R.G. 323, 331, 361 [47]: 361 [48] Doust, I. 256,267 [63] Dowling, P.N. 263,267 [64]: 267 [65] Downarowicz, T. 617, 666 [43] Driouich, A. 267 1661 DmovSek, R. 102, 119 [53] Duan, Y. 515,529 [36] Dubinsky, E. 473,492 [40] Duffie, D. 374,390 1191 Dugundji, J. 603,666 [44] Dunford, N. 106, 107, 119 [54]: 119 [55]: 439,444, 492 [41]: 492 [42] Dupire, B. 371,390 [20] Duren, P.L. 675,694, 705 [17] Durier, R. 650,655,666 [45] Durrett, R. 5 , 6 , 8 3 [lo] Dvoretzky, A. 136, 157 [33]: 458, 475,492 [43]: 492 [44]: 720,735, 774 [52]; 774 [53]; 774 [54] Dykema, K.J. 345,357, 358,362 [72]: 366 [I781 Dynkin, E.B. 616,666 [42] Eaton, M. 910,935 [22] Ebenstein, S.E. 204,230 [3]: 884, 896 [2] EdelStein, I.S. 149, 157 [341 Edelstein, M. 664, 666 [46] Edgar, G.A. 260, 263, 267 [65]: 267 [67]: 267 [68]: 267 [69]: 631, 633, 636, 639, 665 [22]: 666 [47]: 666 [48]; 666 [49] Edmunds, D.E. 969,974 [8] Edwards, D.A. 607, 615, 620, 666 [SO]: 666 [51]: 666 [52] Edwards, R.E. 538,558 [I51 Effros, E.G. 310,313 [I]: 621,622,626,666 [53]: 666 [54] Egghe, L. 260,267 [70] Einstein, A. 369, 390 [21]; 390 [22]; 390 [23] Ekeland, I. 395,409,434 [32]: 434 [33]: 798, 830 [lo]: 831 [51] El Haddad, E.M. 418,420,421,434 [27]: 434 [28]
Author Index El Karoui, N. 382,390 1241 El-Gebeily, M.A. 156, 157 13.51 El-Mennaoui, 0 . 2 6 7 1661 Ellis, A.J. 615, 618, 620, 621. 626, 627, 665 [4] Emery, M. 378, 379,390 [25] Enflo. P. 98, 119 [56]: 129, 133, 134, 156 [2]: 157 [36]; 157 1371: 235, 237, 267 [71]; 273, 279, 280,283,285, 313 141: 313 1231;446,451,479, 489,492 [45): 492 [46]; 543, 545,547, 549,550. 555, 558 [2]: 558 [3]: 558 [9]: 558 [16]: 558 [17]: 558 ~181; 558 1191; 766,772, 774 ~561;774 [571: 804, 805, 829,831 1.521: 831 [53]:831 [54] Erdiis, P. 359,362 [73]; 444,492 1471 Evans, W.D. 969, 974 [S] Exel, R. 326, 362 [74] Fabian, M. 413, 418,434 1341; 434 [35]:434 [361: 434 [37]; 792, 793, 798, 812, 814, 820-822, 828, 831 ~411;831 831 1561: 831 [ s ~ I831 : [581; 831 [591; 832 [601: 832 [61]; 832 [62]; 832 1631: 832 1641: 832 [651 Fakhoury, H. 445,492 1481 Falconer, K. 220,231 1181 Fang, G. 425, 434 [38]: 434 [391; 434 [401; 434 [41l: 435 [42] Farmaki, V. 822, 830 [7] Feder, M. 304,313 [24] Fefferman, C. 220, 223,231 [19]: 231 [20]:267 1721 Feldman, J. 535,558 [S]; 61.5, 666 (551 Fell, J.M. 610, 665 [25] Feller, W. 606, 617. 666 [56]: 804, 832 1661 Fenchel, W. 727. 774 [58] Ferguson, T.S. 906, 935 (231 Fernandez, D.L. 250,267 [73] Fernique, X. 347, 350, 362 1751; 362 [761 Figiel, T. 96, 119 [SO]; 119 1571: 208, 231 (211; 249, 267 1741;267 [ ~ s I276,283,289,309,313 : [~sI; 313 ~261; 313 1271; 314 (281: 443,475,491 1311; 492 1491; 511,529 [Sl]: 583,585-587. 591-594, 595 [IS]: 595 1161: 595 [17]: 596 [24]: 596 [25]: 715, 716, 726, 735, 731, 747, 748, 751, 768, 770, 772, 773 mi: 774 [s91; 774 1601; 774 [611;792. 832 [671: 853, 854, 858, 863, 868 [20]: 869 [21]; 869 1221: 869 [23]: 916,925,935 [24]; 973,973 151 Fillmore, P.A. 323, 331, 339, 361 1471: 361 [48]; 362 [69] Finet, C . 434 [26]: 805,832 [68] Fleming, R.J. 901, 935 [25] Flinn, P. 279, 313 [6]; 314 1291; 51 I , 528 (301 Floret, K. 466,467, 484, 488, 490, 491 [32] Foiag, C. 330, 332, 360 191: 360 [lo] Follmer, H. 387, 388, 390 [26]; 390 1271 Fonf, V.P. 34, 35.83 1261: 61 1,637,641,644, 646-650, 653-659,661465,666 [40]: 666 1571:
[w:
979
666 [%I: 666 1591: 666 [ 6 0 ] ;666 [61]: 666 [62]: 667 [631: 667 [641: 667 1651: 667 [66]; 667 [671: 667 [681: 667 [691; 667 [70]: 667 [71]; 668 1961: 793,813,831 [42]: 831 1431; 832 1691 Force. G. 147, 151, 154, 157 [38] Forrester, P.J. 345, 362 [77] Fourie, J. 484,491 [341 Fradelizi, M. 921, 935 [6] Frampton, J. 413,433 [7]; 799,830 [I81 Franklin, Ph. 575, 596 [26] Frankowska. H. 433 [ I ] Frkchet, M. 906,935 [26] Fremlin, D.H. 89,92,94, 119 1521; 119 [ S S ] : 119 [591 Friis, P. 325, 362 1781; 363 [79] Frobenius, G. 98, 119 [60] Frontisi, J. 828, 832 [70] Fry, R. 813,828,832 [71] Fuhr, R. 626,627,667 [72] Furedi, Z. 175, 193 [7] Gamelin, T.W. 60, 84 [27]; 675-677, 682, 704 131: 705 1121: 705 [18]: 705 [19]; 705 [20]; 879,895, 897 [ I l l Gamlen, J.L.B. 129, 157 [39] Gantmacher, V.R. 443, 492 [SO] Garcia del Amo, A. 527, 529 1521 Garcia, C.L. 285.313 [lo] Garcia-Cuerva, J. 505, 529 [53] Gardner. R.J. 177, 193 [16]: 193 [17]: 193 [18]: 918, 919,935 [27]; 935 [28]: 935 [29] Garling, D.J.H. 137, 139, 157 [40]; 246, 263, 264, 266 1551: 267 [76]: 267 [77]: 267 [78]: 463, 492 [511: 516, 519, 527,529 [54]; 529 [%I: 681, 705 [211: 793, 831 [33]:90.5, 936 1301: 952, 974 [9] Garnett, J.B. 674, 675, 705 1221 Garsia, A. 209, 231 1221 Gaudet, R.J. 129, 157 [39] Gaudin, M. 34.5, 364 11221 Geiss, S. 267 [79] Gelfand. I.M. 902, 908, 919, 920, 936 1311; 936 [32] Geman, S. 344,353,363 [80] Georgiev. P. 828, 832 [72] Gevorkian, G.G. 565, 573,596 1271; 596 [28] Geyler, V.A. 87, 89, 90, 118 [16]: 119 [40] Ghoussoub, N. 34.83 [25]:97, 119 [61]; 119 [62]; 119 [ a ] ; 263,267 1801: 267 p i ] : 267 1821: 402-40s. 425,434 ~411;435 1421;435 ~431; 435 [MI:435 r4sl; 435 ~461:435 1471: 435 [481; 435 [49]: 435 [SO]: 445, 483, 492 [53]:635, 637-639, 667 1731: 667 1741: 667 [75]; 795,798, 799, 821, 831 [34]: 831 [44]: 832 [73] Giannopoulos, A.A. 47.84 1281: 164, 169, 177, 180, 193 ~191;193 ~201:193 1211;193 ~221;342,358,
980
Author Index
363 [811; 719, 722, 725, 726, 729, 737, 754, 755, 766, 774 [62]; 774 [631: 774 [641; 774 [651; 774 (661; 774 [67]; 774 [68]: 774 [69]; 774 [70]; 844, 859, 867,869 [241; 918, 936 [331 Giga, M. 250,267 [84] Giga, Y. 250, 267 1831; 267 1841 Gillespie, T.A. 237, 249, 250,265 [5]; 265 [12]; 265 [ 131 Ginibre. J. 219, 231 [23] Girardi, M. 443, 492 1521; 518, 529 [44] Girko, V.L. 344,363 [82] Glasner, E. 619,620,667 [761 Gleit, A. 658,667 [77] Glimm, J. 542,558 1201 Gluskin, E.D. 175, 193 [23]; 454, 492 [54]; 743, 765, 766,774 r711; 775 ~721;775 ~731;775 [741; 854, 869 1251; 946,968, 974 [lo] Godefroy, G. 33, 34, 57, 83 [6]; 84 [29]; 97, 119 1641; 154, 156,157 [I 11; 157 1411; 235,268 ~851; 285, 295,310,314 1301: 314 1311;406,408,409,415, 434 1241; 434 [25]; 435 [51]; 476,491 [33]; 644, 645, 666 [41]; 667 [78]; 792, 793, 798, 799, 805, 812-814. 821, 822, 828, 829, 830 [22]; 831 [38]; 831 L451; 831 [461; 832 [601; 832 1731; 832 [741; 832 [751; 832 [761; 832 1771; 832 [781; 832 1791; 832 1801: 832 1811;832 r821; 832 1831; 832 [841 Godement, R. 616,667 1791 Goethals, J.P. 916, 936 [341 Gohberg, I.C. 439,492 [55] Gonzalez, M. 812,832 [SSl Gonzalo, R. 413,434 [30]; 812, 813, 831 [47]; 832 [85] Goodey, P. 91 1, 936 [35] Goodman, V. 147, 157 [13]; 230 [4]: 342,361 [231; 463,491 [lo]; 748, 773 [23]; 866, 867,868 121 Gordon, Y. 278,314 [32]; 342, 352, 354,361 [25]: 363 [83]: 363 [84]; 363 [85]: 454,463,466, 492 [511; 492 [54]; 492 [56]; 735,740,749,750. 759, 766,769, 773 [221: 775 [751; 775 [761; 775 1771: 775 [781; 775 [79]; 858, 869 [26]; 896, 897 [121; 925, 936 [361: 946, 952, 968, 974 [9]; 974 [lo] Gorelik, E. 829, 832 [86] Gorin, E.A. 905, 936 [371 Gotze, F. 350, 358,361 [38] Goullet de Rugy, M. 622, 667 [SO] Gowers, W.T. 18, 84 [30]; 136, 157 [42]; 222, 231 [241; 277, 304,310,314 [33]; 770, 775 [Sol; 775 [81]; 812, 813, 832 [87]; 832 [88] Graham, C.C. 874,897 [I31 Granas, A. 603, 666 [44] Grandits, P. 389, 390 [28] Granville, A. 207, 230 151 Greenleaf, F.P. 876, 897 [14]
Gripenberg, G. 573,596 [29] Grobler, J.J. 110, 111, 119 [65]; 119 [66] Gromov, M. 347,363 [86]; 714,717,744,745, 775 [82]; 775 (831; 775 [84]; 775 [85]; 775 [86] Gronbaek, N. 286,314 [34] Grothendieck, A. 273, 281,282, 288, 289, 309, 314 [35]; 440,444,449452,457,459,466,467, 483,486,489,490,492 [57]; 492 [581: 492 [591; 492 [60]: 492 1611: 492 1621; 91 1,936 [38]; 964, 974 [ I l l Griinbaum, B. 930,936 [39] Grzpilewicz, R. 515,529 [56] GuBdon, 0.718, 749, 775 [781: 775 [871: 775 [881 Guerre-Delabribre. S. 134-136, 139, 158 [43]; 158 ~441;519,529 1571; 813,833 ~ 8 9 1 Guionnet, A. 345, 346,361 [22]; 363 [87] Gundy, R.F. 128, 157 [25]: 242, 257,258, 266 [42]; 266 [43]: 266 [44]; 268 [86]; 523, 528 [18]; 880, 897 [ 151 Gurarii, P.I. 581,596 [30] Gurarii, V.I. 581,596 [30]; 626,667 1811 GutiCrrez, J.A. 237, 268 [87]: 676, 705 [23] Gutman, A.E. 87, 106. 108,119 [41] Haagerup, U. 358,363 [MI; 363 [89] Habala, P. 6, 7, 20, 34, 36,41,42. 63, 83 [Ill: 83 [12]: 792, 832 [62]; 833 [90] Hadwiger, H. 733, 775 [89] Hadwin, D. 329,363 [90] Hajek, P. 6,7, 20, 34, 36,41,42, 63,83 [ I l l ; 83 [12]; 434 [36]; 659,666 [40]; 667 [85]: 792, 799, 813, 821, 828,831 [311; 831 [42]; 831 [43]; 832 [61]; ; 833 19I I; 833 ~ 9 2 1 832 [ a ] ; 833 ~901 Halberstam, H. 200, 231 [25] Halmos, P.R. 106, 119 [67]; 535,558 1211 Halperin, I. 505, 526,529 [58] Halpern, H. 336,361 [30]; 859,868 [31 Hardin, C.D., Jr. 901, 902, 905, 936 [40]: 936 1411: 936 [42] Hardy, G.H. 529 1591; 956, 974 1121 Hare, D. 799,831 [45] Harmand, P. 310,314 [36]; 829,833 [931 Harper, L.H. 745, 775 [901 Harrison, J.M. 374, 376, 377,391 [29]; 391 [30] Haydon, R.G. 140, 158 [45]; 408,409,435 [52]; 435 [53l; 435 [541; 526,528 1321; 611,615,644, 667 [821; 667 1831; 667 [841; 799, 813, 833 [94]; 833 [95]; 901, 935 [I71 Heath-Brown, D.R. 229, 231 [26] Heinrich, S. 305-308, 314 [37]; 443,444,455, 492 [63]; 492 [641; 492 [65]; 793,829,833 [96] Hensgen, W. 250,263,268 [88]; 268 [89] Hensley, D. 175, 176, 193 [24]; 193 [25]; 724, 775 [91]
Author Index Hernandez, EL. 518, 527.529 [52]:529 [60]: 529 [el]; 530 1621; 530 [631; 530 [64]; 530 1651: 530 [661 Herrero, D.A. 329-333, 359,362 [621: 363 [91]; 363 [921; 363 [931: 363 [94]; 363 [951 HervC, M. 608, 667 [86] Herz, C. 906. 936 1431 Hewitt, E. 616, 667 [87] Hilbert, D. 439,441,492 [66] Hille, E. 442,492 [67]: 968, 974 [13] Hirsberg, B. 626,667 [88] Hitczenko, P. 257.268 1901 Hoeffding, W. 5 19.530 1671 Hoffman, A.J. 328,363 [961 Hoffman. K. 675,678, 705 1241: 875,878,887, 897 1161 Hoffman-Jmgensen. J. 473,492 [681: 769, 775 1921 Holbrook, J. 327, 361 [36] Holickg, P. 798. 833 [97] Holub, J.R. 960, 974 1141 Hormander, L. 223,231 1271: 731, 775 [931 Horn, A. 447,492 [69] HSU,YP :. 526,529 [45] Huang, C.-F. 374,390 [19] Hudzik, H. 515, 527.529 [37]; 529 [561: 530 [68] Huff, R.E. 633,667 (891; 667 1901 Huijsmans, C.B. 96, 119 1681 Hunt, R.A. 505, 530 [69] Hustad, 0.626.627, 667 [91]; 667 [92] Hutton, C.V. 441. 493 [70] Ioffe, A.D. 418,435 [551: 798,833 1981 Ionescu Tulcea, A. 260,268 1911 Ionescu Tulcea. C. 260, 268 [91] Ishii, H. 420,421,431,434 1181 Ivanov, M. 419,434 [29] Iwaniec, T. 259,268 [921: 268 [931 Jacka, S.D. 385,391 1311 Jacobson, C. 214,231 [28] Jacod. J. 376, 391 [32] Jahandideh, M.T. 102, I19 [69] James, R.C. 235,268 [94]; 275,314 [38]; 477,479, 493 ~711;493 1721; 581,596 p i ] ; 643.667 1931; 792, 804,833 [99]: 833 [loo] Jameson, G.J.O. 87,119 [701; 458,459,493 [73] Jamison, J.E. 901, 935 [25] Janovsky, L.P. 92, 118 [I71 Jaramillo, J.A. 413,434 [301; 676, 705 [23];812, 813. 831 [47]: 832 [SS] Jarchow, H. 47,55, 60,65,66,72,83 [9]; 286, 309. 313 [ill; 455, 458, 459,464, 466, 471, 473, 474, 480, 491 1351; 493 [74]; 493 1751: 493 [76]: 518,
98 1
529 ~401;675,681, 690, 70s [is]: 867,868 [is]; 879, 881, 882, 887. 897 [lo]; 901, 905, 911-915, 929, 935 [IS]: 935 [191; 943,950,951. 974 [71 Jarosz, K. 702,703, 705 1251: 705 1261 Jayne, J.E. 631,668 [94] lensen, R. 420,435 [561 JimCnez Sevilla, M. 664, 665 1161 Jimenez, M. 799,833 [ l o l l Johansson, K. 346,363 [97] John, F. 463,493 1771: 718, 775 [941 John, K. 489,493 [781: 805,833 11021 Johnson, W.B. 88-91.96.97, 105, 108,119 [SO]: 119 [62]: 119 [63]: 119 [71]; 119 [72]: 125, 126, 129. 131, 134-136, 140-143, 145-149, 154, 156, 156 [81: 157 [13]: 158 [46]: 158 [47]: 158 [48]: 158 1491: 158 [sol; 158 [sii; 158 [521; 158 [ s ~ I ; 158 [54]: 158 [ S S ] ; 158 [56]: 158 [57]; 164, 190, 193 1261: 230 141:257,262,268 1951;268 ~961; 273-280,285,288-296,298-300,305,306,309, 310,312,313 [lo]; 313 1121;313 1191; 313 r271: 314 [281; 314 [39]: 314 [40]: 314 [41]; 314 [421; 314 1431: 314 1441; 314 [ ~ s 314 I ; [461; 314 [471; 314 [481: 314 r491; 314 [sol: 314 p i ] ; 336,338. 363 [98]; 443, 445, 459,483, 491 [31]; 492 [52]: 492 [53l: 493 [79]: 493 [SO];511, 521-523, 529 [511: 530 [701; 530 [71]; 530 [72]; 563. 573, 581, 588, 596 [321; 596 1331; 630, 634, 636, 640, 641, 655, 668 1951;744,748,755, 769,770, 773 [23l: 775 [95]: 775 [96]: 775 [97]: 784, 792, 793,805,821,829,831 1351:832 1671: 833 [1031: 833 [1041; 833 [iosi: 833 r1061: 833 11071; 833 [IOS]; 839-845,850,851,853-855, 858-860, 863, 864, 866, 867,868 121: 868 1201; 869 [21]: 869 [221: 869 [27]: 869 1281; 869 [29]: 869 [30l: 869 1311: 869 [32]; 869 [33]: 869 [34]; 906, 925, 929. 936 1441: 936 1451: 936 [46]; 950, 951, 953, 957,961,964,965, 974 [151; 974 1161 Jones, L. 148, 158 [47]; 842, 843,869 [27] Jones, P. 305,314 1.521 Jordan, P. 906, 936 [47] JournC, J.L. 250,266 [54]; 591,596 1201 Junge, M. 590,596 [21l: 724, 775 1981;775 [99] Kadets, M.I. 142, 147, I58 [%I; 463, 493 [81]: 510, 530 [731: 580, 596 [35]: 61 I . 647, 667 [68]; 668 1961: 792, 793, 798,833 [109]: 833 [ I 101: 833 [I 111; 833 [ I 121; 883,897 [IS]; 929,936 [48] Kadison, R.V. 333, 334,363 [99]: 859, 869 1351 Kaftal, V. 336, 361 1301: 859, 868 (31 Kahan, W.M. 339,362 [68] Kahane, J.P. 184. 193 [15]: 363 (IOO]: 644, 645, 668 [971; 873, 897 [17]; 965, 974 [I91 Kakosyan, A.V. 905, 939 11341 Kakutani, S . 443.465.493 1821; 493 t831; 493 1841
982
Author Index
Kalenda, 0. 822, 833 [ 1131 Kalton, N.J. 74, 81, 84 [31]; 95, 120 [73]; 143, 156, 157 [41]; 158 [59]; 158 [60]; 278, 279, 291-295, 297, 301, 302, 310, 311, 313 [13]; 313 [14]; 313 1151; 313 (161; 314 1301;314 ~531; 314 rs41; 314 [SS]; 403, 435 [57]; 505, 511, 518, 527, 530 [74]: 530 [75]; 530 (761; 530 1771; 569, 596 [341; 697, 705 [271; 769, 775 [77]; 792,793, 821, 829, 830 [261; 832 [77]; 832 1781; 832 1791; 832 [80]; 832 [81]; 833 (1141; 843, 854, 868 171; 869 1361; 886,897 [191; 901, 936 [491 Kaminska, A. 5 15,527,529 [37]; 530 [681: 530 [78]: 530 [791: 530 [80]; 530 [81] Kamont, A. 565, 596 1271 Kanter, M. 177,193 [27]: 524,530 [82] Kantorovich, L.V. 87, 91, 92, 106, 107, 120 [74]; 120 1751;120 [761 Karatzas, I. 374,391 1331 Kashin, B.S. 336, 338, 342, 358, 363 [loll; 363 11021: 363 [ I O ~ I575,596 : ~361;7 4 9 . 7 6 ~ 7 7 5riooi Kato, T. 445,493 1851 Katz, N. 222, 231 1291: 231 [301 Katznelson, Y. 184, 193 [15]; 569, 596 [37]; 965, 974 [ 191 Kaufman, R. 793, 828,830 [22]: 833 11151 Kazarian, K.S. 505,529 [53] Kazhdan, J.L. 621,622,666 [54] Keller, 0.-H. 603,668 1981 Kelly, B.P. 249, 265 [6] Kemperman, J.H.B. 914,936 [50] Kendall, D.G. 613,614,668 [99] Kenderov, P. 435 [58] Kesten, H. 358,363 [I041 Ketonen, T. 859, 869 (371 Khovanskii, A.G. 732, 776 [lo21 Kislyakov, S.V. 60, 64, 84 [27]: 84 [32]; 468, 493 [86]; 675, 679,690, 691, 695, 697,699-701, 705 [28]; 705 [291; 705 [30]; 705 [31]; 705 [32]; 879,882,895. 897 [ i i i ; 897 1201: 897 1211 Kitover, A.K. 87, 100.118 [151; 120 1771 Klain, D. 733, 775 [I011 Klebanov, L.B. 905, 939 [I341 Klee, V.L. 602,603,629,650,652,661-663, 666 [32]; 668 [IOO]: 668 [loll: 668 [102]: 668 [1031; 668 [1041; 668 [IOS]; 668 11061; 668 11071; 793,798,833 [116] Klemes, I. 875, 897 [22] Knaust, H. 821,833 [117] Knothe, H. 713, 776 [lo31 Koldobsky, A. 71, 84 [33]; 177. 193 1181; 193 [28]; 524, 529 [46]; 853,869 1381; 905, 906, 909-911, 918, 919, 921-923, 935 (201; 935 [29]; 936 [371: 936 [SII;936 ~521;936 1531; 936 rs41; 936 ~551;
936 [561; 936 [57]; 936 [%I; 936 [59]; 937 [60]; 937 1611; 937 [62] Kolmogoroff, A.N. 370,391 [34] Komorowski, R. 279,314 [56]; 315 [57] Konig, H. 71,84 [331; 268 [97]; 268 [98]; 452, 463, 469,470,493 (881: 493 [89]; 493 [90]; 724,770, 776 11041; 776 [iosi; 853,869 ~381;915-917, 930-934, 937 [63]; 937 [64]: 937 [65]; 937 [66]; 937 [67]; 937 [68]: 937 [69]; 944, 953, 954, 957, 959,961,962,964966,968,970,971,973, 974 [15]; 974 [17]; 974 [18] Koosis, P. 327. 361 1341 Korotkov, V.B. 87, 106, 108, 119 [41]; I20 [78] Kothe, G. 457,493 [91]; 570,596 [381 Krasnoselsky. M.A. 87, 88, 118 [32]; 120 [79]; 120 [80] Krawczyk, L. 389,390 [6]: 390 171; 390 [281 Krein, M.G. 88, 99, 120 1811; 120 [82]; 439,492 [551 Krein, S.G. 74, 78, 81, 83 [13]: 87, 120 [83] Krengel, U. 87, 120 [84] Kreps, D.M. 374,376,391 1291: 391 [351 Kriecherbauer, T. 349,362 [70] Krieger, H.J. 110, 120 [85] Krishnaiah, P.R. 353,366 [187] Krivelevich, M. 346,363 [I051 Krivine, J.L. 136, 138, 139, 158 [61]; 158 [62]; 284, 315 [%I; 455, 491 [30]; 515, 524, 528 [16]; 530 [831; 812,833 [1181; 906,935 [13]; 937 [701 Kruglyak, N.Ya. 74, 83 [S] Kutateladze, S.S. 87, 106, 108, 119 [41] Kunen, K. 268 [99] Kurzweil, J. 798, 813, 833 [119]: 833 [120] Kusraev, A.G. 87, 106, 108,119 [41] Kutzarova, D. 304, 313 [2]; 821, 834 [122] Kwapien, S. 140, 158 [63]; 158 [64]; 254, 257, 268 [iooi; 275,315 1591: 338,347,363 [io61; 363 [1071; 465,466,473,481,493 1921; 493 [931; 493 1941; 527, 530 [84]; 530 [85]; 770, 776 [106]; 820,834 11211; 855, 856, 858, 869 [23]; 869 [39]; 875, 879, 880, 886, 887, 889, 897 [23]; 934, 937 1711 Kyriazis, G. 579,596 1391 Laba, I. 222,231 [29] Lacey, H.E. 465,493 (951; 862,869 [40] Lamb, C.W. 260,268 [loll Lamberton, D. 250,266 [53]; 374,391 [36] Lamperti, J. 905,937 [72] Lance, E.C. 339,364 [I081 Lancien, G. 792,793, 804, 805, 821, 829,832 [78]; 832 1791; 834 [123]; 834 (1241 Landes, T. 515,530 [861 Lapeyre, B. 374,391 [36] Larman. D.G. 177, 193 [29]; 735. 776 [107]; 918, 937 [73]
Author Index Larson, D.R. 341,364 [I091 Latala, R. 460,493 [96]; 718, 776 [ 1081 Lazar, A.J. 614, 622, 624-626,646,658,667 1881; 668[108];668~1091;668~110];668[111];
668[112];6681113] Leach, E.B. 798. 834 [125] Lebourg, G. 798, 831 [51] Lebowitz, J.L. 354,360 [ I ] Ledoux, M. 52.84 [34]; 338, 350, 361 1211; 364 [I 101; 364 [ I 1 I]; 364 [ I 121; 472, 493 [97]; 523,530 [87l; 740, 756, 776 [109l; 848, 849, 869 1411 Leduc, M. 407,409,435 [59]; 812,834 11261 Ledyaev, Y. 412,423,433 [13]; 434 [141; 434 [151; 434 [ 161 Lee, J.M. 238,268 [102]; 268 11031 Lehto, 0. 259, 268 [I041 Lennard, C.J. 526,528 1271; 528 [281; 529 [471 Leranoz, C. 569.596 [34] Leung, D.H. 504, 505,530 [XX]; 530 [891; 530 [901; 530 [91l; 530 [92]: 658,668 [ I 141 Levental, S. 384,391 1371 Levy, M. 51 I , 530 [93]; 813,833 1891 LCvy, P. 739,744, 776 [ I 101; 906, 907. 937 1741 Lewandowski, M. 905,937 [75] Lewis, D.R. 129, 158 [65]; 278. 279, 313 [6]; 314 [321; 466,492 1561; 493 1981; 726, 752, 776 [ I 1 I ] ; 840,858,869 1261; 869 [42l; 869 [43l; 930, 937 [66] Li, D. 156, 157 [41]; 829,832 [XO]; 832 [82] Lidskii, V.B. 451,493 1991; 968, 974 [201 Lieb, E.H. 164, 173, 193 [lo]; 193 1301; 718, 773 [411 Lifshits, E.A. 87, 120 [801 Lima, A.310,314 [36]; 626, 668 [I 151: 668 [ I 161 Lin, B.L. 133, 157 1271; 51 I. 527.528 161; 529 [33]; 529 [34] Lin, H. 325,364 [ 1131 Lin, P.K. 471, 493 [loo]; 526, 528 [32];901, 930, 935 1171; 937 1661 Linde, W. 475,484,493 [loll; 493 11021; 905, 937 [781; 937 [791; 937 [Sol; 937 [XI]; 937 1821 Lindeman, A. 259,265 [S] Lindenstrauss, J. 7, 10-14, 18, 21-27, 30, 33-36. 38, 42.48.50, 51, 78-80, 83 [ 3 ] ;83 [14]; 83 [15]; 83 [26]; 84 1351: 87-91, 105, 108, 119 [72]; 120 [86l; 120 [871; 125, 126, 129. 132-136, 140-143, 145, 146, 154. 156, 158 [48]; 158 [49]; 158 [661; 158 1671; 158 [68]; 158 1691; 158 [701; 159 [71]; 159 1721; 164, 175, 190, 193 [12]; 193 [261; 208,231 1211; 251, 257,258,262, 263, 267 [81l; 268 [951; 268 [l05l: 268 [106]; 268 [107]; 273,275,278,281,282,284,289,295,299-301, 305-307, 309, 31 I , 312, 314 [46]; 314 [47]; 315 [60];315 1611; 315 [62]; 315 [63]: 315 [64]; 315 1651; 315 1661; 315 [671; 403,435 [451; 440,
983
443,459,465,466,475,492 [49]; 493 [XO]; 493 11031; 494 [104]; 494 [105]; 505, 511, 514, 515,518, 530 1941; 531 1951; 531 [96]; 531 [97]; 531 1981; 531 [99]; 563. 573, 574, 577, 580, 581, 588,596 [32]; 596 [40]; 596 [41]; 596 1421; 596 [43]; 602,618,624-626, 630,634-636, 638-641,645,646,651,653-655,658,661-663, 665 [XI; 666 [37]; 666 [38]; 667 [69]; 667 [73]; 668[951;668[113];668[117];668[118]; 668 [1191; 668 11201; 668 [121]; 668 [122]; 668 [1231; 681, 705 1331: 715,716,735-737.742, 744, 747, 748, 766, 768-770, 773 [30];773 [31]; 773 [32];773 [37]; 773 1381; 773 [39]; 774 [61]; 775 1951; 776 [ 1 121; 776 [ 1 141; 776 [ 1 151; 784, 792, 793, 798, 805, 821, 822, 829, 830 [I]; 830 [2]; 830 161; 830 [14]; 831 1341; 831 [54]; 832 [69]; 833 11041; 833 11061; 833 [107]; 834 [1271; 834 [1281; 834 [1291; 834 11301; 839-842, 844, 845, 850, 855, 858, 860, 865, 866, 868 181; 869 1281; 906,916,925,9277929,935 171; 935 [lo]; 935 1111; 935 [12]; 935 1241; 936 [a]: 937 [76l; 937 [771; 950,951,953, 963-965, 974 1161; 974 [21]; 974 1221; 974 [23] Linhart, J. 769, 776 [ I 131 Lions, P.L. 420,421,426,431. 434 [IX]; 434 [19]; 434 [20] Lisitsky, A. 909, 937 [83] Littlewood, J.E. 956, 974 [I21 Litvak, A. 736,737, 746, 767,769, 773 [191; 776 11161; 776 [117]; 776 [ I l S ] Litvinov, G.L. 536, 558 [22] Llavona, J.L. 676, 705 1231 Lofstrom, J. 74, 76-78, 80, 83 [4]; 505, 528 [ I I]; 577, 595 [2]; 692, 704 [4] Lomonosov, V.I. 102. 105, 120 [XX]; 120 [89]; 536, 538, 543, 545,558 [22]; 558 [23];558 [24]; 558 1251: 558 [26]; 641, 668 [I241 Lonke, Y. 910, 91 I , 937 [62]; 937 1841 Loomis, L.H. 668 [I251 Lopachev, V.A. 905, 910,937 [SSl; 937 [86]; 937 1871 Lopez, J.M. 886,897 [24] Lorentz, G.G. 505, 524, 526,531 [loo]; 531 [ I O I ] ; 531 11021; 575,596 1221 Lorentz, R.A. 578,596 [44] Loring. T.A. 325, 326,362 [74]; 364 [I141; 364 [I 151 Lotz, H.P. 92, 119 [441; 455, 456,484,494 [106]; 531 [I031 Lovaglia, A.R. 792,834 [ 131] Lozanovsky, G.Ya.87,89, 106, 107, 119 1391; 120 [90]; 120 [9l] Lubotzky, A. 338,359,364 11161 Lucchetti. R. 435 [58] Luecking, D. 471,494 [I071 Lukacs, E. 937 [XX]
984
Author Index
Lusky, W. 273, 302, 303,315 [68]; 315 [691; 315 [701; 581,596 [45]; 596 1461; 596 1471; 626, 668 11261; 668 11271; 905,937 1891 Lust-Piquard, F. 192, 193 1311 Lutwak, E. 918,937 [90] Luxemburg, W.A.J. 87.96, 119 [68]: 120 [92]; 120 1931; 120 1941 Lyubich, Y. 914, 916, 917, 937 [911; 937 1921; 938 1931 Maaden, A. 435 1601 MacCluer, B.D. 471,491 1291 Magidor, M. 444,492 [47] Magill, M. 115, 120 [95] Maiorov, V.E. 973, 974 1241 Makai, E. 959, 974 [25] Makarov, B.M. 87, 106, 108,119 [41] Maleev, R. 814, 834 [132] Maligranda, L. 515,531 11041 Mandelbrot, B.B. 371,391 I381 Mani, P. 735,768, 776 [1071; 776 11191 Mankiewicz, P. 47, 84 [361; 277, 304, 315 [71]; 315 1721; 315 1731; 358, 364 [ I 171; 633, 668 11281; 766, 776 [120]; 776 [121]; 793, 829,833 1961 Mareenko, V.A. 343,344,353,357,364 [118] Marcinkiewicz, J. 250,268 [IOS]; 522,531 [lo51 Marcus, M.B. 523,531 11061 Margulis, G.A. 359,364 [I 191; 364 [I201 Martin, D.A. 144, 159 1731 Martin, G . 259,268 [93] Mascioni, V. 277,300, 306,315 1741; 315 I751 Mastylo, M. 527,530 1681 Matheson, A. 875, 896 [ 11 MatouSek, J. 798, 834 11331; 925,928, 938 I941 MatouSkovi, E. 798,834 [1331; 834 [I341 Mattner, L. 905, 938 [95] Mauldin, R.D. 458,494 [I081 Maurey, B. 11, 51,53,84 1371; 84 1381; 125, 133, 134, 136, 139, 140, 148, 158 1451; 158 [SO]; 158 [62]; 159 1741: 159 [75]; 159 1761; 159 1771; 237, 242, 251, 263,267 [SO]; 267 1811: 267 1821; 268 l1091; 277, 284, 304,310,314 1331: 315 1761; 402405, 435 [45]; 435 [46]; 435 1471; 435 1481; 435 1491: 459,473,474,477,494 [ I O ~ 494 I ; 11 101: 494 [ I l l ] ; 494 11121; 511, 515, 523, 530 [70]; 530 [83]; 635,637-639.667 [73]: 667 1741: 667 1751; 745,770,771. 776 11221; 776 11231; 776 [124]: 776 11251; 799,812,821,829,832 1731; 833 [118]: 834 [135]; 834 11361; 841,842,845, 855, 857, 858, 867,869 [291; 869 1441; 869 1451; 869 1461; 883, 884, 897 1251; 906, 921, 935 [I]; 935 161; 936 1461: 953, 957,961,964,965, 974 [IS]; 974 1261 Maynard, H.B. 634, 668 11291
McCann, R.J. 173, 193 [32]; 713, 776 [I261 McCarthy, Ch.A. 568,596 1481 McConnell, T.R. 250,264,268 [I 101; 268 11 1I]; 268 [I 121 McGehee, O.C. 576,577,596 1491; 874,897 I131 McGuigan, R. 658,667 [77] McIntosh, A. 327, 329,347, 361 [35] McLaughlin, K.T.-R. 349,362 1701 McMullen, P. 732, 733, 776 [127]; 776 [I281 Medzhitov, A. 515,531 [I071 Mehta, M.L. 342, 344,345,364 [121]; 364 [I221 Memin, J. 382,391 [39] Mendelson, S. 663,669 11301 Merton, R.C. 369,371,391 1401 Meyer, M. 179,194 1331; 724, 728, 749, 759, 775 1781; 775 [79]; 776 [IOS]; 776 [129]; 777 [130]: 777 [I311 Meyer, P.A. 374-376, 389,390 [17]: 390 [18]: 610, 614, 616,665 1251; 666 [331; 669 11311 Meyer, Y. 566,577,578,596 [SO] Meyer-Nieberg, P. 87,89,92,96, 120 1961 MCzard, M. 354,364 11231 Miao, B. 348,360 [I71 Michael, E. 596 [Sl] Michaels, A.J. 103, 120 1971 Milman, D.P. 792,834 [I371 Milman, V.D. 47, 48, 51, 53, 83 [16]; 84 1281; 169, 175, 193 [121; 193 [211; 208,231 [21]; 231 [31]; 342, 347, 3.58, 363 [811: 363 [86]: 364 [124]; 471, 475,492 [49]; 494 [113]; 523,531 [108]; 710, 713-719,722-724,726,729,731,732,735-750, 752,754,755,758,759,762,766-772, 772 151; 772 181; 772 191: 772 [lo]; 773 1331: 773 [34]; 773 1361; 773 1371; 773 [38]; 773 1391; 773 1401; 774 [491: 774 r611: 774 1641: 774 [651; 774 [661: 774 [67]: 774 [68]: 775 [85]: 775 [86]; 776 [I 141; 776 [117]; 776 [1181; 777 [132]; 777 [133]; 777 [1341: 777 [1351; 777 [136]; 777 [137]; 777 [138]; 777 [139]: 777 11401; 777 11411; 777 11421: 777 11431; 777 11441; 777 11451; 777 11461; 777 11471; 777 11481; 777 11491: 777 [ISO]; 777 [1511: 777 [1521; 777 l1531; 777 11541; 792, 821, 834 11381; 844, 845, 848, 850, 852, 859,865-867, 868 181; 868 [17]; 869 [24]: 869 1471; 884, 892,897 [S]; 897 1261; 914-916, 924,925,927,935 [121: 935 1241: 938 1961: 938 1971 Milne, H. 676, 705 1341 Minc, H. 98, 120 1981 Mirsky, L. 329,364 11251 Misiewicz, J. 906,909, 91 1, 938 [98]; 938 [99]; 938 [IOO]; 938 [loll Mityagin, B.S. 459,494 [ I 141; 906, 935 [I]
Author Index Molto, A. 792, 834 11391 Monat, P. 389,390 [lo] Monniaux, S. 250,268 11 131 Montesinos, V. 832 1621 Montgomery, H.L. 198, 224-226. 231 [32]; 345, 364 11261: 555,558 [9] Montgomery-Smith, S.J. 74, 81.84 1311; 249, 259, 265 [6]: 265 171: 523, 527, 529 [48]; 531 [lo91 Moreno, J.P. 664, 665 (161: 799, 833 [loll m on is, P. 633,667 ~891; 667 1901 Morton, A. 374,390 [XI Morzocchi, M. 425,434 [17] Muckenhoupt, B. 579,596 [52] Muhly, P.S. 237. 249, 250, 265 [121; 265 1131 Mujica, J. 677. 705 [35] Miiller, C. 926, 938 [ 1021 Muller, P.F.X. 130, 131, 159 [78]: 159 1791: 159 [XO] Murray, F.J. 129, 159 [Xl]: 484.494 11 151 Musiela, M. 374, 391 1411 Musielak, J. 515,531 [110]
985
159 [84]: 159 [85]: 255,256,267 [SX]: 276,279, 313 [4]; 313 [17]; 565, 581, 595 [l]: 596 [23]; 612, 645,669 [134]; 813, 820, 821, 833 [117]:
834 [1431: 834 [144]: 839, 862, 868, 868 [l] Odlyzko, A.M. 345,362 [77]: 364 11271 Oikhberg, T. 275. 285,314 [48] Oleszkiewicz, K. 460, 493 [96] Olevskii, A.M. 213,231 [34]; 251,269 11141; 269 r i i s i Olin, R.F. 341,364 [128] Olsen, G.H. 618,668 [I221 O’Neil, R. 505,511, 531 [1171: 531 [118] Ordower, M. 340, 362 1631 Orihuela, J. 792, 834 [ 1391 Orlicz, W. 458,473,494 [ I 181 0rno. P. 92, 121 [ 10.51 Otto, F. 350,364 11291
Pajor, A. 175, 179,193 [14]; 194 [33]; 454,492 1541: 718,722-724,728,750,756,759,762,767,769, 773 1191; 776 [ios]; 776 [i171; 777 [1301; Nachbin, L. 88,120 [99] 777 [i311: 777 [ M I ; 777 11471; 777 114x1; Nagasawa, M. 702, 705 [36] 777 [1551; 777 11561; 854,868 11.51;924, 938 [97]: Naimark, M.A. 607,669 [132] 946,968, 974 [lo] Nakano, H. 87,89, 120 [loo] Paley, R.E.A.C. 250,269 [ 1161 Namioka, I. 87, 88, 120 11011; 798, 834 [140] Palmon, 0 . 7 6 7 , 778 [157] Naor, A. 845, 869 1481 Panzone, R. 107, 118 1341 Narayan, S.K. 320,360 [XI Paouris, G. 765, 778 [lSXl Natanson, I.P. 507,531 [ 1111 Papadimitrakis, M. 164, 193 [22]; 725, 774 [69]: 918, Nathanson, M. 207,231 [33] 938 [ 1041 N Z K O V , F. 1x4. 192.194 1341 Papadopoulou, S. 660,669 11351 Negrepontis, S. 139, 157 1121; 659, 669 [133] Papini, L. 650,655,666 [45] Nemirovski, A.M. 812. 813, 834 [141] Parisi, G. 354,364 [123] Newman, C.M. 147, 157 1131; 230 141; 342,361 1231: Parrott, S.K. 339, 364 11301 463,491 [lo]: 748, 773 [23]:866,867,868 [2] Pastur, L.A. 343, 344, 349, 353, 357,364 [ I 181: Neyman, A. 906,938 [I031 Ng. K.F. 87, 122 [143] 364 11311;364 11321; 364 11331 Nica. A. 345. 357, 358,366 117x1 Paulsen, V.I. 320, 362 1641: 364 11341 Niculescu, C. 480,494 [ 1171 Pay& R. 642,666 1351 Nielsen, N.J. 146, 159 [82]: 268 [98]: 277, 279, 286, Pearcy, C.M. 322,364 r13s1;536,539,558 1271; 304,315 1711;315 1771; 315 17x1;315 1791;s i x , 559 12x1 531 [112] Peck. N.T. 278, 301, 302,314 1541; 403,435 1571 Nikishin, E.M. 516,531 [113] Peetre, J. 481, 494 [ I 191 Nikol’skii, S.M. 575,597 [53] Peirats, V. 518, 529 1601 Nordgren, E.A. 101,119 [48] Pelant, J. 832 [62] Nordlander, G. 793, 834 11421 Pelczynski, A. 96,119 [SO]; 125, 129, 132, 135, 142, Novikov, I. 523,531 11 141; 595,597 1541; 597 1551 146. 147. 156, 158 I.581: 158 [66]; 158 [67]; Novikov, S.Ya. 527,531 [ I 151: 531 [ I 161 159 1x61; 251, 255, 256,268 [IOS]: 269 [117]; Nussbaum, R.D. 99, 112,120 [102]: 120 [103]: 269 11181; 269 [119]; 274,277,290, 300, 305. 121 r i m ] 315 [XO]; 315 [XI]; 315 [82]; 440, 443,445, 458, 459,462466,473,479,491 1311; 492 [40]; Oberlin, D. 685, 705 [37] 494 [1041: 494 [114]: 494 [120]; 494 [121]; Odell, E. 59.83 1221; 133, 139, 141, 143-145, 147, 494 ~1221: 494 [mi;494 11241; 494 11251; 510, 149, 150, 154,156 121: 158 1511; 158 1521; 159 ~831; 530 ~731:577, 5 8 ~ 5 8 2596 , psi; 596 1411;
986
Author Index
596 [511; 597 [561; 597 [57]; 603, 648, 665 [9]; 665 [lo]; 675, 688, 703, 704, 705 [39]; 792, 830 [171; 858,869 [23]; 875, 876, 878-880, 883, 886, 887, 889, 894, 896 [3]; 897 [IS]; 897 [19]; 897 [231; 897 [27]; 897 [28]; 901,905,911-915, 935 [181; 951,963,974 [211; 974 [271 Pena, A. 769, 773 [21] Peressini, A.L. 87,121 [lo71 Perissin&, I. 722, 774 [70] Perron, 0.98.121 [lo81 Persson, A. 462,494 [I261 Petrushev, P. 579,596 [391; 597 [581 Pettis, B.J. 444,492 [41]; 792, 834 [145] Petty, C.M. 164, 177, 193 [13]; 194 [351; 725, 778 [159]; 918, 935 [16] Petunin, Yu.1. 74, 78, 81, 83 [13]; 87, 120 [83] Pezzotta, A. 662, 667 [70] Pfaffenberger, W.E. 558 [12] Phelps, R.R. 34, 35,83 [26]; 435 [61]; 610, 611, 614, 626,627,634,640, 641, 645,646, 659,664, 665 [131; 667 [72]; 668 [121]; 669 [136]; 669 [1371; 669 [138]; 669 [139]; 669 [140]; 793, 798, 832 [69]; 834 [140] Phillips, R. 338, 359, 364 [116] Piasecki, M. 264,269 [120]; 269 [I211 Pichorides, S.K. 482,494 [127] Pietsch, A. 250,269 [122]; 439, 440,444, 448, 451455,458,459,462464,466,469471,475, 476,480,484,493 [102]; 494 [126]; 494 [128]; 494 [1291; 494 [130]; 494 [131]; 494 [132]; 494 [1331; 494 [134]; 494 [135]; 494 [136]; 495 [137]; 495 [138]; 495 [139]; 590, 597 [59]; 929,935 [19]; 943-945,947,950,951,953,954. 957,961,966,969,973,974 [7]; 974 [28]; 974 [291; 974 [301; 974 1311; 974 1321 Pigno, L. 576,577,596 [49] Pinsker, A.G. 87, 106, 107, 120 1751 Pintz, J. 207, 230 [5] Pisier, G. 47, 51, 53, 83 (171; 206, 231 [35]; 231 [36]; 235,237,269 [123]; 269 (1241; 276,277,284-286, 305, 314 [49]; 315 [76]; 315 [83]; 315 [84]; 315 [85]; 320, 338, 346, 365 [136]; 365 [137]; 365 [1381; 365 [139]; 365 [140]; 365 [141]; 466, 468,471,473475,477,479,483,489,494 [I 121; 495 [140]; 495 [141]; 495 [142]; 495 [143]; 495 [144]; 495 [145]; 495 [146]; 495 [147]; 495 [148l; 495 [149]; 495 [150]; 518,523, 531 [106]; 531 [119]; 695, 705 [38]; 710, 726,752, 754,760,767,770-772, 776 [124]; 776 [125]; 777 [1491; 778 [160]; 778 [161]; 778 [162]; 778 [1631; 778 [164]; 793, 804, 805,812, 831 [54]; 834 [1361; 834 [146]; 834 [1471; 834 [1481; 834 [1491; 843, 845, 849,870 [49]; 870 [50]; 870 [511; 879, 882,886, 888,891,893,894,
897 [291; 897 [30]; 897 (311; 897 [32]; 964, 974 1331 Pitt, L.D. 94, 121 [1091; 905, 936 [42] Pittenger, A.O. 257,269 [125] Pitts, D.R. 341, 362 [65]; 362 [66]; 365 [142] Plemmons, R.J. 98, 113, 118 [35] Plichko, A. 304,313 [21 Pliska, S.R. 374, 377, 391 [30] Plotkin, A.I. 901-903,905,910, 937 [861; 937 [871; 938 [105]; 938 [106]; 938 [107]; 938 [log]; 938 [lo91 Polya, G. 907, 938 [110]; 938 [ill]; 956, 974 [I21 Polyrakis, I.A. 116, 118 [141 Pommerenke, Ch. 701, 705 [40] Pompe, W. 573,597 [601 Popescu, G. 341,365 [143] Popovici, I.M. 97, 98, 121 [110]; 121 [ 1111 Poulsen, E.T. 617, 669 [141] Power, S.C. 339. 341,362 [67]; 365 [144] Preiss, D. 37, 42, 84 [39]; 396, 399,413,418, 425, 433 [S];434 [37]; 435 [50]; 435 1621; 793, 798, 820, 821, 829, 830 1191; 833 [107]; 834 [1401; 834 [150]; 834 [1511 F’rivalov, A.A. 578, 597 [61] Privalov, 1.1. 694, 705 [41] Protter, P. 376, 391 [42] Priiss, J. 250,268 [113]; 269 [126]; 269 [I271 Przelawski, K. 793, 830 [13] Przeworska-Rolewicz, D. 445,495 [ 1511 Pukhlikov, A.V. 732, 776 [lo21 Pustylnik, E.I. 87, 120 [79] Quenez, M.C. 382,390 [24] Quinzii, M. 115, 120 [951 Rabinovich, L. 917,938 [I121 Radjavi, H. 99, 101, 119 [48]; 121 [112]; 542,549, 559 [29]; 559 [30] Raghavan, T.E.S. 98, 113, 118 [331 Rainwater, J. 609, 669 [142] Raja, M. 805, 828,834 [152]; 835 [I531 Randrianantoanina, B. 515,531 [120]; 901,905, 936 [49]; 938 [113]; 938 [114] Range, R.M. 675,683, 705 [42] Rao, M. 612,669 [143] Raynaud, Y. 137, 139, 140,158 [431; 159 [87l; 159 [88]; 159 [89]; 519,523,527,528 [8]; 528 [9]; 531 [121]; 531 [1221; 531 11231; 531 [124]; 531 [i251; 531 [i261; 532 [1271;532 11281 Read, C.J. 98, 102,121 [115]; 121 [ 1161; 121 [I 171; 274, 295,297, 299, 301,315 [86]; 549, 550, 555, 556, 559 [31]; 559 [32]; 559 [33]; 559 [34]; 559 [35]; 559 [36]
Author Index Reinov, 0,462,484,495 [152]; 495 [I531 Reisner, S . 527.532 [129]; 759, 775 [79]; 778 11651; 896,897 [12] Retherford, J.R. 953,957,961,964,965, 974 [IS] Revalski. J. 434 1311 Revuz. D. 372,373, 378,379,391 1431 Reznick, B. 916, 917, 938 11151 Rhandi, A. 112.121 11131; 121 [114];433 [ l l l Ribe. M. 829, 835 [154]; 835 11551 Rieffel, M.A. 484.495 11541; 634, 669 11441 Riesz, F. 442,495 [1551; 495 11561 Riesz, M. 244,269 11281 Ringrose, J.R. 339,365 114.51; 439,495 [157] Roberts. J.W. 403,435 1571; 603, 669 [145] Robinson, A. 535,558 1101 Robinson, D.W. 620, 665 [23] Rochberg, R. 703, 705 (431; 706 1441; 706 1451 RodC, G. 641, 669 [1461; 669 11471 Rodin. V.A. 519,532 I1301 Rodriguez-Salinas, B. 518, 527,529 [61]; 530 1621; 530 1631; 530 [641 Rogalski, M. 611,669 [148] Rogers, C.A. 177, 193 1291; 458,492 1441; 631, 668 1941; 720, 735, 774 [54]; 918, 937 1731 Rogers, L.C.G. 376, 383,391 1441; 391 [451 Rohlin, V.A. 616,669 11491 Rolewicz, S . 445, 495 11511; 769, 778 [166] Ropela, S . 575.597 1621 Rerdam, M. 325.362 1781; 363 [791 Rosenoer, S. 340,365 11461 Rosenthal, H.P. 20, 59, 84 1401; 101, 119 1481; 125, 128, 129, 131, 134, 140, 145-148, 150, 151, 153, 154, 157 1111; 157 [21]; 157 1371; 158 [54]; 158 [68]; 159 1901; 159 1911; 1.59 [92]; 159 1931; 204, 231 1371; 255,265 1221; 268 1991; 269 11191; 274, 277, 279, 290, 291, 293, 294, 296, 300, 305, 314 [501; 314 [SI]; 315 [81]; 315 1871; 445, 466, 473,480,492 1401; 494 [lOS]; 495 [1581; 495 [1591; 521,532 11311; 542, 549,559 1291; 559 "1; 576. 581, 595 1111; 596 1331; 612, 614, 636,644, 645.665 1191; 669 [134]; 669 [150]; 669 11511; 813, 821,833 [IOS]; 835 [156]; 835 [1571; 868,870 [52]; 875, 876, 897 1331 Rosinski. J. 905, 938 11161 Ross, K.A. 886,897 1241 Rosset, S . 730, 778 11671 Roth, K. 200. 231 1251 Royden, H.L. 6, 13, 15. 2 0 , 6 2 8 3 1181 Rubio de Francia, J.L. 250,269 [129]; 269 [1301; 269 11311 Rudelson, M. 737,767, 778 11681; 778 11691; 778 I1701 Rudin, W. 19, 37, 83 1191; 147, 159 1941; 197, 205, 206,231 1381; 359,365 11471; 603,669 11.521; 883, 897 1341; 901,902, 938 11171
987
Rudnick, Z. 345,365 11481 Ruelle, D. 354,360 [I]; 619,620,669 11531 Ruiz, C. 527,530 [651; 530 1661 Ruston, A.F. 449,450. 484. 495 [1601; 495 [161]; 495 11621 Rutkowski, M. 374, 391 1411 Rutman, M.A. 99,120 1821 Ruzsa, I. 205,231 1391 Ryan, R. 263,269 11321 Ryll-Nardzewski, Cz. 906, 938 [ l o l l
Saab, P. 95,120 1731 Saakyan, A.A. 575,596 1361
Saccone. S.F. 682,685,686, 706 1461; 706 1471 Sagher, Y. 505,529 1391 Sahakian, A. 578,596 [44] Saint Raymond, J. 633, 669 11541; 759. 778 11711; 793,831 1381 Saks, S . 444,490 141 Salem, R. 873, 897 1171 Salinas, N. 324, 332, 333, 360 [ I l l ; 362 1621; 365 [1491; 536,539,558 I271 Samuel, C. 286, 305,315 [88]; 316 1891 Samuelson, P.A. 370.391 1461 Saphar, P.D. 285, 310,314 [31]; 461,495 11631; 495 11641; 792. 829,832 [Sll Sarnak, P. 338. 345, 359, 364 [ I 161; 365 11481 Savage, L.J. 616,667 1871 Scalora, F.S. 260,269 11331 Schachermayer, W. 263.267 [82]; 374, 377-386, 389, 390 1101; 390 [ I l l ; 390 1121; 390 [13]; 390 1141; 390 1151; 390 [161; 391 1471; 391 [481; 404, 435 1491; 633,636,669 [1551; 669 [156]; 798,799, 832 [731; 835 [I581 Schaefer, H.H. 87, 88,92,99, 110, 111, 121 11201; 121 [121]; 121 [122]; 121 11231; 121 [124]; 821, 835 [I591 Schaffer, J.J. 805, 835 11601 Schatten, R. 439, 484, 486,495 [165]; 495 (1661; 495 [1671; 495 11681; 495 [I691 Schauder, J. 273,316 [90]; 442,495 [170] Schechtman, G. 47,48, 51.53, 83 1161; 129-132, 134, 140, 148, 150, 153, 154, 156, 157 1211; 158 [SO]; 158 1531; 158 [55];159 1781; 159 1791; 159 1951; 159 [961; 159 1971; 176, 194 [361: 208, 231 1311; 257,268 1961; 278, 314 1461; 336, 338, 346, 347, 363 [981; 364 11241; 365 [ISO]; 471,475, 494 11131; 511,521-523, 530 1701; 530 1711; 530 1721; 531 11081; 576, 581, 595 1111; 710,713, 718,735,737,739-742,745,746,748,749,755, 769, 770, 775 1961; 775 1971; 776 11181; 777 [ISO]; 777 [151]; 777 [152]; 778 11721; 778 11731;
988
Author Index
778 [174]; 778 [175]; 792, 793, 821, 829, 833 [1061; 833 [107]; 840, 844, 845, 847, 848, 851-853, 855,857-859, 862-864,866,867, 869 [211; 869 1221; 869 [29]; 869 [30]; 869 [31]; 869 [321; 869 [331; 869 [34]; 869 [461; 869 [471; 870 [531; 870 [541; 870 [55]; 870 1561; 884, 897 [26]; 906, 925, 936 [45]; 936 [46] Schep, A.R. 108,121 [125] Schipp, F. 576, 597 [63] Schlumprecht, Th. 139, 143, 144, 159 [84]; 159 [85]; 176, 177, 193 [18]; 194 [36]; 813, 820, 821, 833 11171; 834 [1431; 834 [144]; 919, 935 [291 Schmidt, E. 439,441,496 [171]; 716, 778 [176] Schnaubelt, R. 112, I21 [ 1141 Schneider, R. 710,720,726,734, 778 [177]; 778 [178]; 911, 921, 938 [118]; 938 [119]; 938 [ 1201 Schoenberg, I.J. 906, 938 [121] Scholes, M. 369,371,390 [41 Schur, I. 439,442,496 [173]; 496 [I741 Schiitt, C. 140, 158 [63]; 158 [64]; I59 [88]; 471, 496 [172]; 525-527, 530 [84]; 530 [85]; 532 [128]; 532 [132]; 532 [133]; 855, 856, 869 [39]; 933, 934, 937 [67]; 937 [71] Schwartz, J.T. 106, 107, 119 [54]; 439,492 [42] Schwartz, L. 471,496 [175] Schwarz, H.U. 87,121 [126]; 488,496 [176] Schweizer, M. 387-389, 390 [lo]; 390 [26] Sedaev, A.A. 527,532 [134] Seidel, J.J. 916,936 [34] Seifert, C.J. 479,491 [36] Semadeni, Z. 625,669 [157] Semenov, E.M. 74,78, 81,83 [13]; 87, 120 [83]; 519, 523,527,531 [114]; 531 [1161; 532 [1301; 597 [55]; 812, 813, 834 [141] Sersouri, A. 798, 835 [158] Shapiro, J.H. 470, 471, 496 [177]; 496 [178]; 496 [179]; 496 [180] Sharpley, R. 505,515,528 [lo] Shashkin, Yu.A. 640, 669 [158] Shatalova, O.A. 917,937 [92] Shcherbina, M. 349,361 [45]; 364 [1331 Shields, A.L. 322,364 [135]; 559 [281 Shiga, K. 440,490 [2] Shilov, G.E. 902, 908, 920,936 [31] Shiryaev, A.N. 374,391 [49] Shlyakhtenko, D. 360,365 [151] Shreve, S. 374,391 [33] Shultz, F. 627, 665 [3] Shura, T.J. 136, 157 [28]; 276, 313, 313 [18]; 812, 830 [27] Silverstein, J.W. 349, 353, 365 [152] Silverstein, M.L. 266 [44] Simon, B. 439,496 [181]
Simon, P. 576,597 [63] SimoniE, A. 105, I21 11271; 121 [128]; 535, 559 1371 Simons, S. 644,669 [159] Sinai, Ya.G. 349,365 [153] Singer, I.M. 333, 334,363 [99]; 859,869 [35] Sjolin, P. 595,597 1641 Skorohod, A S . 384,391 [37] Smickih, S.V. 99, 122 [ 1461 Smidek, M. 798,833 [97] Smith, B. 576, 577,596 [49] Smith, K.T. 535,558 [4] Smoluchowski, M. 369,391 [SO] Smulyan, V.L. 793,835 [161] Snobar, M.G. 463,493 [81]; 929,936 [48] Sobolev, A.V. 87, 120 [80] Sobolevsky, P.E. 87,120 [79] Soh, H. 250,267 [83]; 267 [84]; 269 [I271 Solin, P. 223,231 [15] Sondemann, D. 387,390 [27] Soshnikov, A. 346,365 [154] Soshnikov, A.B. 349,365 [153] Spalsbury, A. 545,559 [38] Srinivasan, T.P. 702, 706 [48] Starbird, T. 129, 131, 147, 157 [31]; 157 [36]; 822, 830 [15] Stegall, C. 129, 158 [65]; 483, 496 [182]; 798, 835 [162]; 835 [163]; 835 [164] Stein, E.M. 215,231 [401; 244,267 [72]; 269 [134]; 269 [135]; 269 [1361; 505,532 [135]; 894,897 [35] Stephani, I. 452,491 [25]; 958,959,973, 973 [3] Stephenson, K. 905, 938 [122] Stem, R. 412,434 [15]; 434 [16] Sternfeld, Y. 618,625, 668 [122]; 669 [160] Stezenko, V.Ya. 88, 118 [32] Stmner, E. 620, 669 [161]; 669 [162] Stout, E.L. 675, 706 [491 Straszewicz, S. 628,669 [I631 Strichartz, R. 217,231 [41] Stricker, C. 374, 385, 389, 390 [l]; 390 [6]; 390 [7]; 390 [lo]; 391 [Sl] Stromberg, J.O. 575,597 [65] Study, E. 712, 774 [451 Sucheston, L. 260,267 [69] Sudakov, V.N. 756, 778 [179] Sukochev, F. 515,531 [lo71 Sunder, V.S. 106,119 [67]; 327,365 [155] Swart, J. 484,491 [34] Synnatzsckke, J. 92, 106, 121 [129] Szankowski, A. 275,283-285, 305,316 [91]; 316 [92]; 316 [93]; 316 [94]; 489,496 [183]; 735, 766, 776 [115]; 778 [I801 Szarek, S.J. 146, 159 [98]; 273, 295, 299, 301, 302, 316 [95]; 322, 331, 345, 358, 359, 361 [24]; 363 [95]; 365 [1561; 365 [157]; 365 [158];
Author Index 365 [159]; 365 [160]; 365 [161]; 365 [162]; 365 [1631; 365 [164]; 460,496 [184]; 580, 597 [661; 736, 737,749,765-767, 770, 773 [19]: 773 1351;774 1551; 778 [isi]; 778 [1821; 778 [1831; 778 11841; 778 11851; 778 11861; 779 [187l; 859. 868 [16]; 870 [571; 969, 973 [61 Szego, G. 907. 938 [ 11 11 Szlenk, W. 805, 835 [165] Talagrand, M. 96, 121 [130]; 181, 194 [37]; 201, 231 [421; 338, 344, 346, 350, 354, 359,364 [I 121; 365 [165]; 365 [166]; 366 [167]; 366 [168]; 366 [169]; 472,473, 493 [97]: 496 [185]; 523, 530 [871; 633,665 [20]; 737, 740,755,756,766, 769, 776 [109]; 778 [186]; 779 [188]; 779 [189]; 828,835 ~1661; 844,845,848,849,869 r411; 870 lS81; 870 1591; 885,898 [361; 925, 939 [I231 Tamarkin, J. 968. 974 [ 131 Tang, W.K. 793,835 (1671 Tao, T. 222. 223,231 [29]: 231 [30]; 231 [43]; 231 [44] Taylor, A.E. 265 1171 Taylor, P.D. 470,496 [ 1801 Thompson, A.C. 664,666 [46] Thomson, J.E. 341,364 [I281 Thorbj@msen,S. 358,363 [XS]; 363 [891 Tokarev, E.V. 527,531 [ I 161 Tomas, P. 215,231 [451 Tomczak-Jaegermann, N. 43,4547,51,53,60,64, 66. 83 [201; 84 [36]; 266 [55]; 277, 279, 287, 315 ~571; 315 ~721; 315 1731; 315 ~781; 316 ~961; 358,364 [117]; 365 [164]; 459,464,471,473, 496 [186]: 496 [187]; 496 [188]; 710,722,726, 736,743,749-751,756,757,765-767,770, 774 1491; 774 [60]; 776 [I 161; 776 [121]; 777 11551; 777 (1561; 779 [187]; 779 [1901; 779 11911;779 r1921; 779 [1931;779 [1941; 779 [195]; 793, 831 [33]; 835 [168]; 841, 850,853, 854, 858, 864,868 [17]; 869 [251; 870 1601; 930-934, 937 [67]; 937 [68]; 937 [691 Tonge, A. 47. 55.60, 65,66, 72,83 [9]; 455,458,459 464,466,471,474,491 [351; 518,529 [40]; 675, 681,690. 705 1151; 867, 868 [18]; 879, 881,882, 887, 897 [lo] Torrea, J.L. 250. 269 [131]; 505,529 [53] Tomnczyk, H. 799,835 [I691 Tracy, C.A. 349,358,366 [170]; 366 11711 Trautman, D.A. 519,526, 527.528 [ZX]; 528 1291; 529 [49] Triebel, H. 463,495 [138]; 958, 973 [4] Troitsky, V.G. 102, 121 11311; 121 [132]; 556,558, 559 [39]; 559 [40] Troyanski, S.L. 635, 669 [164]; 792, 805, 814, 821, 832 1831;832 ~841; 834 [ m ] ; 834 [mi; 834 [ 1391; 835 11701
989
Tsay, J. 348,360 [I71 Tsirelson, B.S. 136, 140, 159 [991; 276, 316 [97]; 799, 812,835 [ i n ] Tsolomitis, A. 722, 774 [701 Turett, B. 505,528 1311 Tzafriri, L. 7, 10-14, 18.21-27, 30, 33, 50,51, 78-80, 83 [141; 83 [151; 84 [41]; 87, 92, 120 [86]; 120 [87]; 121 [1331; 125, 133-135, 140, 143, 146, 147, 158 [501; 158 [691; 158 [70]; 159 (711; 159 [72]; 159 [IOOI; 257, 258,268 [106]; 268 [107]; 276, 279,281,282,284,289,300,301,305-307,309. 313 [ U I ; 313 r151; 315 ~641;315 [651; 315 ~661; 315 [671; 336, 337,361 [42]; 361 [43]; 361 [44]; 505, 51 I , 514, 515, 518, 523, 529 [51]; 530 [70]; 530 [941; 531 [95]; 531 [961; 531 [97]; 531 [98]; 531 [991; 574,577,580,596 1421; 596 1431; 6x1, 705 [33]; 784, 834 [130]; 853-855, 859, 860, 862-864. 866, 867, 868 [7]; 868 [9]; 868 [lo]; 868 [ I 11; 868 [12]; 868 1131; 869 [25]; 869 [291; 870 [61]; 906,928. 936 [46]; 937 [77]; 965, 974 [23] Uhl, J.J. 35, 36, 3 8 4 0 , 83 [Sl; 106, 107, 119 [51]; 259, 260, 263, 266 [561; 479, 480,483. 484, 491 [37]; 558 1141; 793,831 [501 Vaaler, J.D. 175, 194 [381 van Dulst, D. 515,529 [50] van Mill, J. 603,669 [I651 van Rooij, A.C.M. 91, 121 [IIX]; 121 [I191 Vandenverff, J. 434 [36]; 828,835 [172] Varga, R. 555,559 [41] Vargas, A. 222, 231 [44] Vaserstein, L. 916, 917, 938 [93] Vasin, A.V. 905, 939 [I241 Vega, L. 222,23 I [44] Veksler, A.I. 87, 89, 119 [391; 119 [40] Venakides, S. 349,362 [701 Venni, A. 250, 267 [60]; 267 [61]; 267 [62] Vesely, L. 650, 655-657, 667 1711; 669 [I661 Vilenkin, N.Ya. 919, 936 1321 Villa, R. 716, 772 [I I] Villadsen. J. 627,669 [1671 Villani, C. 350,364 11291 Vincent-Smith, G.F. 612,669 [I681 Virasoro, M.A. 354,364 [123] Vlasov, L.P. 663, 670 [I691 Voiculescu, D. 323, 326, 330-332, 345, 347, 356-360, 360 [91; 360 [lo]; 366 11721; 366 [173]; 366 [174]; 366 ri751; 366 ~1761;366 [ I V I ;366 van Koch, H. 45 1,493 1871 von Neumann, J. 439,447,484,494 [115]; 494 [ I 161; 495 [169]; 670 [17Ol; 906,936 [47l
990
Author Index
Vu, V.H. 346,363 [I051 Vulikh, B.Z. 87, 88.91.92, 106, 107, 120 [75]; 120 1761; 121 [134]; 121 [135]; 121 [I361 Vuza, D.T. 97, 98,121 [ 1101; 121 [ 11 I] Wachter, K.W. 343, 353, 366 [179] Wade, W.R. 576,597 [63] Wagner, G. 769, 779 [196]; 925, 939 [I251 Walsh, B. 95,121 [137] Wang, G. 215,230 [I]; 230 [2]; 259,265 [91; 265 [lo]; 269 [I371 Wang, J.-K. 702, 706 [48] Wassermann, S. 332, 359, 366 [180]; 366 [181] Wegmann, R. 358,366 [I821 Weil, W. 906, 911, 936 [35]; 938 [120]; 939 [126] Weinberger, W.F. 339,362 [68] Weis, L. 445,446,496 [189]; 860, 870 [62] Weiss, B. 619,620,667 (761 Weiss, G. 336,361 [30]; 505,532 [135]; 859,868 [3] Wells, J.H. 906, 939 [127] Wenzel, J. 250. 269 [1221; 269 [138]; 269 [139]; 471, 476,495 [139]; 590,597 [59] Wermer, J. 673475, 704 [ I 1; 706 [50] Werner, D. 143,158 [60]; 821,829,833 [93]; 833 [I 141 Werner, E. 798,835 [158] Werner, J. 116, 118 [24] Werner, W. 829,833 [93] Weyl, H. 326,366 [183]; 439,496 [190]; 948, 974 [34] Wheeler, R.F. 636, 666 [49] White, M.C. 968, 974 [351 Whitfield, J.H.M. 413,434 [37]; 792, 798, 812, 832 [63]; 832 [83]; 832 [84]; 834 [125]; 835 [173] Whitney, H. 484,496 [I911 Wickstead, A.W. 91, 93, 95, 96, 105, 118 [13]; 118 [18]; 118 [19]; 118 [20]; 118 [21]; 121 [138]; 122 [139]; 122 [140]; 122 [141]; 122 [I421 Widder, D.V. 605,670 [171] Widom, H. 349,358,366 [170]; 366 [171] Wielandt, H.W. 328,363 [96] Wigner, E.P. 342, 358,366 [184]; 366 [I851 Williams, D. 376,391 [45] Williams, L.R. 906, 939 [I271 Willinger, W. 374,390 [8] Willis, G.A. 286, 309, 314 [34]; 316 [98] Wils, I.M. 334, 366 [I861 Wo-Sang Young 576,597 [67] Wojtaszczyk, P. 7, 13, 16, 58,66, 83 [21]; 84 [42]; 146, 149, 157 [34]; 159 [82]; 274, 279, 31 1, 313 [16]; 315 [79]; 315 [82]; 459,471,477, 496 [192]; 518,532 [136]; 566,567, 569,570, 573, 575,577,578,580,596 [34]; 597 [681; 597 1691; 597 [701; 597 [71]; 597 [72]; 597 1731: 597 [74];
674, 675, 686, 687, 689, 690, 702, 703, 706 [Sl]; 841,870 [63]; 905, 936 [30] Wolenski, P. 412,434 [16] Wolff, T. 221, 223, 231 [461; 232 [471 Wolfson, H. 767,771, 772, 773 [40]; 777 [153]; 777 [ 1541 Wolniewicz, T. 704, 706 [52] Wolnik, B. 573,596 [28] Wong, Y.C. 87,122 [ 1431 Wood, G.V. 278,314 [55] Woiniakowski, K. 577,578,597 [69] Wulbert, D. 626,668 [ 1231
Xu, Q. 477,495 [ISO]; 700, 705 [32]; 879,898 [37] Yahdi, M. 828, 835 [I741 Yan, J.A. 380,391 [52] Yin, Y.Q. 344, 353, 358, 360 [18]; 360 [19]; 361 [20]; 366 [187] Yood, B. 558 [12]; 559 [42] Yor, M. 372, 373,378, 379,391 [43] Yosida, K. 443,496 [I931 Yost, D. 793, 830 [13]; 835 [I751 Zaanen, A.C. 87, 89,99, 106. 110, 120 [93]; 220 [94]; 122 [144]; 122 [I451 Zabreiko, P.P. 87, 99, 120 [79]; I22 [I461 Zachariades, Th. 139, 157 [12] Zaidenberg, M.G. 901,939 [I301 Zajitek, L. 435 [631; 798, 828,832 [641; 833 [971; 834 [151] Zalgaller, V.A. 710,712, 726, 774 [43] Zanco, C. 662,667 [70] Zastavnyi, V.P. 909, 939 [128]; 939 [129] Zeitouni, 0. 346, 363 [87] Zemanek, J. 959,974 [25] Zhang, G. 177, 194 [39]; 918, 919, 939 [131]; 939 [132]; 939 [133] Zhong, Y. 101, 119 [48] Zhou, X. 349,362 [70] Zhu, K. 471,494 [lo71 Zhu, Q. 418,433 191 Zimmermann, F. 250,269 [140] Zinger, A.A. 905,939 [I341 Zinn, J. 52, 84 [34]; 129, 158 [55]; 176, 194 [36] Zippin, M. 19,84 [43]; 131, 134, 140, 141, 146, 149, 158 [54]; 158 [56]; 158 [57]; 159 [loll; 274, 277, 279, 290, 293,294, 296, 300,314 [SO]; 316 [99]; 581,596 [33]; 829, 833 [lo81 Zizler, V. 6, 7, 13, 20, 33, 34, 36, 41, 42, 57, 63, 83 [6]; 83 [Ill; 83 [12]; 84 [44]; 406,408, 409, 413,415,434 [241; 434 [25]; 434 [371; 476, 491 [331; 644,666 [411; 783, 789, 792,793, 798,
Author Index
99 1
This Page Intentionally Left Blank
Subject Index positive, 286 - unconditional, 291,295 - u n i f o r m (UAP), 60, 305 ff - uniform projection (UPAP), 305 ff arithmetic diameter, 891,893 Asplund space, 410, 411,795 associate space, 512 asymptotic freeness of matrices, 356, 359 asymptotic order, 320 Auerbach lemma, 45 automorphism, 3
A-convex, 404 A-martingale, 402 A~-set, 404 absolutely continuous vector measure, 39 abstract L p space, 22 abstract M space, 22 adjoint -Banach ideal, 457 - ideal norm, 457 admissible class of perturbations, 395,398, 399 admissible cones of perturbations, 401 Aldous theorem, 515 Alexandrov theorem, 420 Alexandrov-Fenchel inequalities, 727 algebra - C*, 627 - Jordan-Banach, 627 allocation, 113 almost isometric embedding, 925 almost sure convergence, 343, 352, 353, 357, 360 Amir-Cambern theorem, 702 analytic continuation, 908, 921 analytic decomposition of unity, 699 analytic distribution, 921 analytic function on Banach space, 675, 806 -m-homogeneous polynomials, 676 analytic map, 806 analytic Radon-Nikod2~m property (ARNP), 236, 262, 638 approximate identity, 882 approximating sequence, 288, 292, 296, 297 approximation numbers, 452, 945 approximation property (AP), 12, 97, 275 ff, 488 - b o u n d e d (BAP), 12, 274 ff - bounded compact, 308 - commuting bounded (CBAP), 12, 291 ff -commuting compact, 310 -commuting metric, 291 -commuting unconditional metric, 295 - compact, 308 - m e t r i c (MAP), 12, 287 if, 488 - m e t r i c uniform, 307
-
( B ( X , Y ) , r), 281,282 B2-sequences, 917 g-differentiable, 406 balayage, 401, 610 Banach couple, 74 Banach ideal, 448 - adjoint, 457 Banach lattice, 21, 60, 89, 125 - p-concave, 27, 504, 855,964 - p-concavity constant, 27 - p-convex, 27, 504, 841,855,964 - p-convexity constant, 27 - absolute value, 21 - functional calculus, 26 - order complete, 23 - order continuous, 23 - order continuous, functional representation, 24 - symmetric, 21, 81 Banach space - B-convex, 52, 474, 894 - C(K), 19 - K convexity constant, 53 - K-convex, 53 - Loc (#), 14 - L p ( l Z ) , 1 ~< p < ec, 13 - L p ( l Z , X), 1 ~< p < oc, 37 - cotype q, 48 - L o r e n t z space, 21 - O r l i c z space, 21 - predual of C ( K ) and L1 (/z), 20, 625,657
993
994
Subject Index
- reflexive, 10, 31,443 - separable conjugate, 10, 38, 635,639 smooth, 30 - Sobolev space, 585 stable, 515, 519 - strictly convex, 30 - superreflexive, 33, 56, 235,237, 479 - t y p e p, 48, 845 -uniformly convex, 31 -uniformly smooth, 31 - w e a k l y sequentially complete, 4 with enough symmetries, 74 Banach-Diedonn6 theorem, 381 Banach-Mazur compactum, 765 Banach-Mazur distance, 3, 43, 858, 862, 924 Banach-Saks property, 444 Bang lemma, 185, 188, 192 barrier, 638 PSH, 638 strong, 638 barycenter, 604 basic sequence, 7 basis, 7, 274, 287 if, 585 K-equivalence, 8 bimonotone, 7 - boundedly complete, 10, 275, 285 constant, 7 equivalence, 8 Faber-Schauder, 9, 564 - Haar, 9, 125 if, 250 ff - Markuschevich, 13 - monotone, 7, 255,256, 277 - problem, 273 - shrinking, 10, 278, 294, 299 - subsymmetric, 11 - symmetric, 11,837, 854-858 - unconditional, 9, 126, 251, 274 if, 301 if, 855, 858 monotone, 9 universal, 11 Berg technique, 331 Bernoulli selectors, 338 Bernstein theorem, 606 Berry-Ess6en theorem, 934 Besicovitch set, 198, 220 Besov space, 575,585, 971 Bessaga-Pelczynski theorem, 648 Bessel inequality, 950, 951 Bessel process, 378, 384 Beurling-Ahlfors transform, 236, 259 biorthogonal functionals, 7 biquasitriangular operators, 332 Bishop-Phelps theorem, 34, 385,395, 396, 641 -
-
-
-
-
-
-
-
-
-
-
-
u
n
c
o
n
d
i
t
i
o
n
a
l
l
y
Black-Scholes model, 371 Blaschke-Santal6 inequality, 728 block basis, 7, 131,134 BMO, 389, 591,695 Bochner integrable functions, 36, 969 Bochner integral, 36, 402 Bochner theorem, 906, 911 Bochner-Riesz multiplier, 222 Borell lemma, 717 Borwein-Preiss theorem, 396 boundary, 639 - Choquet, 640 640, 656 boundedly complete finite dimensional decompositions, 299 Bourgain algebra, 687 Bourgain projections, 699 Bourgain theorem, 687, 689 Bourgain-Stegall minimization principle, 414 Boyd index, 257, 258, 514, 527 Brascamp-Lieb inequality, 164 ff Brenier map, 713 Br6ndsted and Rockafellar theorem, 395 Brown-Douglas-Fillmore theorem, 320, 323, 331 Brownian motion, 369 Brunn-Minkowski inequality, 178, 711 Brunn-Minkowski theorem, 921 bump function, 400, 794 Busemann-Petty problem, 901,918, 919 -
m
i
n
i
m
a
l
,
C1,285 C 2, 295, 301 Cp, 279, 280 C 1 function, 37 C n function, 37 C a function, 37 Calder6n couple, 80 Calder6n-Zygmund singular integral operator, 245, 894 Calkin algebra, 323, 334 Calkin theorem, 447 Cameron-Martin-Girsanov theorem, 373 canonical embedding, 4, 21 Carath6odory theorem, 602, 913-915, 927 central limit theorem, 850 characteristic function, 6 Choquet, Bishop-de Leeuw theorem, 610 Choquet, integral representation, 607 ff Clarkson inequality, 479 closed hull of an ideal, 441 ideal, 441 codomain, 440
-
-
Subject Index
coefficient problem, 163, 183, 189 Cohen idempotent theorem, 874 commodity space, 113 commutant, 539, 540 commutator, 557 compact operator, see operator, compact compactification - Bohr, 645, 876 - Stone-Czech, 645 complementary function, 512 complemented subspaces, 4, 129, 143, 155, 837, 839, 863-868 complemented subspaces theorem, 928, 965,967 complex convexity, 637 complex interpolation spaces, 964 component of a vector in a lattice, 23 component of an operator ideal, 441 composition inequality, for p-summing operators, 65 composition operator, 470 concentration of measure on the sphere, 737 concentration phenomenon, 744 cone generating, 87 constructivity, 330, 332, 358-360 contraction principle, 848, 850 convergence in distribution, 343, 355 convergence in probability, 343,352 convex body, 164, 169, 174, 646 if, 710 ff convolution inequalities, 163 corona problem, 674 coset ring, 874, 875 cotype, 48, 126, 286, 505,526, 882, 892 - Gaussian, 472 - Haar, 477 coupon collector's problem, 847 covering number, 756 critical point theory, 423 crossnorm, 485 general, 486 reasonable, 485 uniform, 486 CSL algebra, 340 cubature formulas, 916 cylindrical measure, 475
-
-
995
density, 840 -change of, 837 ff change of, 840, 849 Maurey change of, 842 Pisier change of, 843 density hypothesis, 225 dentable, 35, 397, 634 derivation, 340 differentiability - Fr6chet (F), 37, 405,788 (G), 37, 408, 788 of a convex function, 41,409 of a vector measure, 39 of Lipschitz functions, 42 differential games, 426 differential subordination, 253, 258 dilation, 610 dimension conjecture, 703 direct sum, 4 direct sum, infinite, 5 Dirichlet problem, 622 Dirichlet series, 198 discrepancy theory, 927 disk algebra, 673, 874, 879 distance maximal, 837, 859 ff distribution function, 5 domination problem, 94 Doob-Meyer theorem, 383 Doubling strategy, 376, 377 dual ideal, 441 Duhamel integral formula, 218 Dunford theorem, 107 Dunford-Pettis property (DP), 61,444, 687 Dvoretzky theorem, 47, 475,735,740, 844, 915 Dvoretzky-Rogers factorization, 737 Dvoretzky-Rogers lemma, 47, 720 dyadic - martingale, 476 - representation, 453 dyadic tree, e-separated, 56 -
L
e
w
i
s
G
~
t
e
a
-
-
-
u
x
-
-
-
-
A 2 condition, 513 A 0 condition, 513 Davie theorem, 673 decomposition method, 14, 125, 129, 135, 149, 151,866 decomposition, ~p, 12 decomposition, monotone, 12 decoupling inequalities, 338
e-entropy, 891 eigenfunctions, 213 eigenvalues, 213,943 ff Ekeland theorem, 395, 396 ellipsoid of maximal volume, 46, 164 if, 719 ff entropy numbers, 202, 946, 958 equiangular lines, 917, 931 equiangular vectors, 931 equimeasurability theorem, 902, 903, 912 evolution case, 426 exchange economy, 113 expectation, 5
996
Subject Index
expectation, conditional, 6 exposed point, 35, 628 if, 790 extension of isometries, 902 extension theorem, 902, 904 extrapolation principle, 15 extremal vectors, 545, 546, 602 ff f (X) = f (X, X), 280
9r(X, Y), 280 qg-function, 514, 527 E and M. Riesz theorem, 878 face, 602 - closed, 614 factorization of operators, 96 factorization property, 443 Fatou norm, 89 filter, 55 filtration, 236, 253, 254 finite decomposition, 837, 865-868 finite dimensional decomposition (FDD), 11, 140 if, 296 ff finite dimensional expansion of the identity, 300 finite geometries, 359 finite nuclear norm, 966 finite rank operator, 106, 441,880 finite variation, measures of, 39 finitely representable, 136, 138 first order Hamilton-Jacobi equations, 426 first order smooth minimization principle, 406, 408 first order sub- and super-differentials, 415 fixed point property, 526 Fock space, 356 F611mer-Schweizer decomposition, 389 Fourier transform, 481,901 -restriction to surfaces, 198, 216 fractional Brownian motion, 383 fractional derivative, 920 Frrchet-differentiable norm, 408, 418 Fredholm resolvent, 441 free central limit theorem, 355 free Poisson distribution, 353 free probability, 354-359 freeness, 355 function - affine, 607 affine continuous, 612 - almost periodic, 644 -completely monotonic, 605 - concave, 251,608 - convex, 608 first Baire class, 611 - Haar, 476 divisible, 606
- plurisubharmonic, 397, 637 positive definite, 606 - Rademacher, 460 envelope of, 608 upper semicontinuous, 608 upper semicontinuous envelope of, 612 function spaces on compact smooth manifolds, 583-588 of function spaces, 587 decomposition of the manifold, 586 spaces on subsets, 585 spaces with boundary conditions, 586 function, biconcave, 236, 251 function, biconvex, 235, 237 function, maximal, 237 functional - w*-support, 628 -support, 628, 641 fundamental theorem of asset pricing, 374, 379 -
- u p p e r
-
-
- d e c o m p o s i t i o n
-
-
-
G-viscosity superdifferential, 419 Yl, 858
Y2,841 Garsia conjecture, 209 G~teaux differentiable norm, 408, 788 Gauss measure, 475, 905 Gaussian correlation problem, 179 Gaussian processes, 338, 351,848 Gaussian variables, 5, 16, 68, 839, 850, 853 Gelfand numbers, 454, 945, 954 general equilibrium, 113 general perturbed minimization principle, 397, 398 generalized Hankel operator, 681 generating cone, 87 gl constant, 837, 858 Glicksberg problem, 680 Gordon-Lewis local unconditional structure (G-L 1.u.st.), 59, 278-280, 302, 466, 680, 834, 872 ff Gorelik principle, 826 Grothendieck constant K G , 67, 467, 860 Grothendieck inequality, 67, 190, 842, 860 Grothendieck space, 505 Grothendieck theorem, 338, 467, 688, 872 ff
-
-
- i n f i n i t e l y
Haar
- function, 476 - polynomial, 476 Haar null set, 42 Hadamard lacunary sequence, 875, 883, 886 Hamilton-Jacobi equation, 419 ff Hardy inequality, 500, 887, 894, 956
Subject Index Hardy spaces, 874 Hausdorff dimension, 220 Hausdorff metric, 924 Hausdorff-Young inequality, 77, 504, 970 Hedging problem, 116 Hilbert spaces - characterization of, 965 Hilbert transform, 235,244, 481 Hilbert-Schmidt norm, 328 Hille-Tamarkin kernel, 469 Hoeffding inequality, 519 H61der continuous, 425 homogeneous function, 910 homogeneous polynomials, 915 hyper-reflexive, 339 hypercontractivity, 50, 350 hyperplane problem, 722, 924 ideal, 21, 90, 440 p-Banach, 448 Banach, 448 - closed, 441 dual, 441 idempotent, 444 injective, 446 quasi-Banach, 448 - regular, 446 Neumann, 447 - sequence, 446 - surjective, 446 - symmetric, 441 - ultrapower-stable, 455 ideal p-norm, 448 ideal norm, 448 - adjoint, 457 non-normalized, 448 ideal property, 61 ideal quasi-norm, 448 ideals of operators on Hilbert space, 66 idempotent ideal, 444 idempotent measure, 875 independent, 5 indicator function, 5 information theory, 350 rejection, 444 mjective, 18, 58, 285 hull of an ideal, 445 ideal, 446 - tensor norm, 486 injective tensor product, 882 injective, separably, 18 insurance, 115 integral operator, see operator, integral -
-
997
intermediate space, 75 interpolation pair, 75 Interpolation space, 502, 511,526, 527 interpolation,/C-method, 78 Interpolation, complex method, 76 Intersection body, 918, 919 lnvariant mean, 872, 876, 877 lnvariant subspace, 98 if, 533 ff inverse Blaschke-Santal6 inequality, 759 inverse limit, 624 Isometric embedding, 906 into Lp, 524, 901 into lp, 911 isometry, 3, 515,526 isomorphic symmetrization, 759 isomorphism, 3 isoperimetric inequality, 163, 173, 346, 715 isotropic - constant, 723 - measure, 722 - position, 722 - vectors, 911 -
-
-
-
-
-
-
-
-
-
S
c
h
a
t
t
e
n
-
v
o
n
James theorem, 34, 385,643 Jensen theorems, 421 John position, 718 John representation of the identity, 721 John theorem, 46, 169, 170, 718 joint densities of eigenvalues, 344, 349 K-convexity, 53,483, 845, 894 K-functional, 78, 502, 523 /C-divisibility theorem, 80 k-intersection body, 923 Kadets-Klee property, 515, 527 Kadison-Singer problem, 333, 859 Kahane inequality, 472 Kahane-Khintchine inequality, 50 Kakeya maximal function, 221 Kakutani representation theorem, 22 Kashin decomposition, 359, 360 KB-space, 89 Khintchine inequality, 16, 26, 460, 472, 519, 717, 850, 934 Kislyakov theorem, 679, 691 Kn6the map, 712 Kolmogorov number, 454 Kolmogorov rearrangement problem, 208 Krein-Milman property (KMP), 633 Krein-Milman theorem, 602, 928 Krein-Rutman theorem, 99 Krein-Smulian theorem, 87, 381
998
Subject Index
Krivine theorem, 48 Kwapieri-Schtitt inequality, 855 Ky Fan norms, 328 L2(ep), 523 L~0, 511, 527 L~0(0, 1), 518 Lp(Lq), 523,527 Lp,cx~, 500, 518 Lp,q, 500, 505, 523 Lw,q, 524, 526, 527 L-functions, 198 e-position, 751 e~, 518 e p, ~ , 500, 505, 519 g.p,q , 500 g-w,q, 525 /21 -space, 302 /~p-space, 57, 126, 146, 279, 287 /2p,Z-space, 57, 129 Ap-set, 197 if, 854, 872 ff
A~o,w(I), 527 Laplacian on the torus, 213 large deviations for random matrices, 344, 346, 348 lattice norm, 89, 820 lattice of measurable functions, 694 BMO-regular lattice, 696 lattice order, 613 left approximate identity, 286 Legendre polynomials, 926 Leontief model, 113 Levi norm, 89 L6vy families, 744 L6vy processes, 370 Lewis lemma, 45 Liapunov theorem, 602 Lidskii trace formula, 451,463 lifting property, 17 limit order of an ideal, 469 linear perturbation principle, 403 Liouville theorem, 905 Lipschitz isomorphic Banach spaces, 826 Littlewood-Paley decomposition, 879 local theory, v, 43,321,455, 710 local basis property, 302, 303 local martingale, 378, 380 local reflexivity principle, 53 locally bounded, 375 locally uniformly convex norm, 784 locally uniformly lower semicontinuous, 418 logarithmic Sobolev inequality, 350 long James space, 411 Lorentz function space Lw,q (I), 524 -
Lorentz sequence space, 519 if, 955, 957, 971 Lorentz spaces, isometries of, 526 low M*-estimate, 749 lower p-estimate, 504 lower q-estimate, 504, 514 Luxemburg norm, 511 M-ellipsoid, 759 M-ideal, 310 majorizing measures, 181,338 Marchenko-Pastur distribution, 353 Marcinkiewicz interpolation theorem, 502, 503, 505, 510 Marcinkiewicz set, 879, 880, 887 marketed space, 115 martingale, 6, 235, 253, 380, 401,476, 630 - analytic, 263 -difference sequence, 6, 236, 476 - dyadic, 242 inequality, 852 - simple, 239 - square function, 256, 257 tangent, 264 -transform, 235, 237, 262, 880 matrix splitting, 859 Matuszewska-Orlicz indices, 514 Maurey-Khintchine inequalities, 510 Maurey-Nikishin-Rosenthal factorization theorem, 872, 883, 884 Maurey-Pisier theorem, 51 maximal quasi-Banach ideal, 456 maximum principle, 426 mean value theorems, 422 measurable function, 36 measure Banach space valued, 39 -concentration of, 321,338, 346, 349, 735 ff - ergodic, 609, 615 -Gauss, 475, 905 - Haar, 617, 873,933 - Jensen, 637 609, 613 stable, 901 T-invariant, 615 variation of, 39 unique maximal probability, 613 - Wiener, 476 Mergelyan theorem, 673 metric (see also approximation property) - entropy, 331,338 injection, 444 - zr-property, 295, 296, 300 surjection, 445 -
-
-
-
m
a
x
i
a
l
-
-
-
-
-
-
t
o
t
m
a
l
,
Subject Index Milne theorem, 676 Milyutin theorem, 702 minimal mean width position, 725 minimal surface position, 724 Minkowski box theorem, 174 Minkowski functional, 918 Minkowski sum, 711 if, 844, 924 Mityagin-Petczyfiski theorem, 688 mixed discriminant, 731 mixed volumes, 726 modular space, 858 modulus of continuity, 971 - convexity, 31, 413 - convexity of power type p, 413 operator, 90 - smoothness, 789 - uniform convexity, 785 moment method, 344 Montgomery conjectures, 223 mountain pass theorem, 424 Muckenhaupt condition, 389 multiplicity, 943 multiplier, 872 ff -
-
N-function, 511 Nagasawa theorem, 702 nest algebra, 339 Nikishin factorization theorem, 516 Nikolskii inequality, 464 No Arbitrage (NA), 378, 384 No Free Lunch with Vanishing Risk (NFLVR), 379ff non-normalized ideal norm, 448 non-smooth calculus, 414 nonlinear Schr6dinger equation (NLS), 197 normal structure, 515 normed vector lattice, 89 nuclear, see operator, nuclear representation, 449, 461 -
999
p-integral norm, 72 - p-nuclear, 461 - p-summing, 63, 220, 459, 475, 677, 929, 950, 957, 959-961 p-summing norm (rrp(T)), 63 (p, 2)-summing, 959 - (p, q)-summing, 459, 677, 950 - (q, p, X)-summing, 693 or-nuclear, 488 - V-summing, 475 - p-summing, 475 CgA-universal, 443 absolute integral, 106 -absolutely summing, 63,677 - almost commuting, 320 almost integral, 106 approximable, 441 band irreducible, 110 - biquasitriangular, 332 - b l o c k diagonal, 324, 330 - compact, 4, 19, 94, 281,316, 442, 535,538, 542, 658, 943,957 - compact friendly, 103 continuous, 442, 686 - composition, 470 - convolution, 973 creation/annihilation, 356 diagonal, 469 dominated, 93 -essentially normal, 323 - factoring through, 14 of, 14, 96 finite nuclear, 966 finite rank, 441 Fourier type p, 481 Fredholm operator, index of, 63 - Gaussian cotype q, 472 - Gaussian type p, 472 - Haar cotype q, 477 - Haar type p, 477 - Hilbert-Schmidt, 439, 470, 949, 950 Hilbertian, 465 - Hille-Tamarkin, 969 -integral, 106, 457, 462, 475,969 lattice homomorphism, 21 isomorphism, 21 lattice-factorable, 466 of, 17 - nearly commuting, 320 - n e a r l y dominated, 687 - nuclear, 45,286, 449, 461, 881,929, 959 ff - order bounded, 90 - P a l e y operator, 678 -
-
-
-
-
-
-
-
-
-
c
o
m
p
l
e
-
-
p
-
c
o
n
c
a
v
i
t
y
-
p
-
c
o
n
v
e
x
i
t
y
-
e
l
y
-
-
-
- f a c t o r i z a t i o n
-
-
-
-
-
operator, 3 1-integral, 457 - 1-nuclear, 45,449 - k-normal, 331 - L p-factorable, 465 - lp-singular, 445 - p-concave, 27 constant, 27 - p-convex, 26 constant, 27 p-integral, 71,462, 488
t
-
- l a t t i c e
-
- l i f t i n g
Subject Index
1000
21, 88 - power-compact, 943, 953,959, 969 quasinilpotent, 90 quasi-p-nuclear, 878 - quasitriangular, 332 - Rademacher cotype q, 472 Rademacher type p, 471 regular, 90 related, 944 -Riesz, 943, 958, 959 - singular integral, 235,244, 249 -strictly L p-factorable, 465 -strictly p-integral, 462, 677 - strictly cosingular, 445 strictly singular, 62, 445 - super weakly compact, 479 transitive, 549 ff translation invariant, 875 - UMD, 476 - uniformly p-smooth, 478 -uniformly q-convex, 478 -uniformly convex, 478 -uniformly convexifiable, 479 - uniformly smooth, 478 -universal, 443, 483 - weakly compact, 4, 95, 442 - w e a k l y singular, 970 operator algebras, 543 operator ideal, 440 p-Banach, 448 Banach, 448 quasi-Banach, 448 operator spaces, 354, 357 operators commuting with translations, 873, 879, 880, 883 optimal control or differential games, 419 optimal control theory, 426 optimal portfolio, 115 option, 369 order continuity, 89 order continuous norm, 89 ordered vector space, 87 ordinal index, 139, 154 Orlicz class, 511 Orlicz function, 855-857 Orlicz norm, 512 Orlicz property, 465 Orlicz sequence space, 140, 512, 518, 522, 658, 855 Orlicz space, 511,523, 527, 855, 910 Orlicz spaces, isometries of, 515 Orlicz-Lorentz space, 527 orthonormal system, 201 -
p
-
-
-
-
-
-
-
-
-
-
-
o
s
i
t
i
v
e
,
p-Banach space, 403 p-concavification, 30 p-convex lattice, 504 p-convexification, 30 p-integral operator, see operator, p-integral p-nuclear, see operator, p-nuclear p-stable random variable, 861 p-stable random vector, 906 p-subadditive, 402 p-summing operator, see operator, p-summing (p, 2)-bounded, 126, 147, 149, 150 PSH p-martingales, 403 p(X), 284 Jrl, 858 Zrp, 840, 861 Jr-property, 295-301,307 zrz-property, 295, 307, 310, 312 Palais-Smale around F, at altitude c, 423 Paley inequality, 678, 679, 894 parabolic Hamilton-Jacobi equations, 428 Pareto optimal allocation, 114 Parseval equality, 460, 921 path of complemented subspaces, 866 paving problem, 334 payoff operator, 116 Petczyfiski property, 685 perfectly homogeneous, 134 periodic boundary conditions, 218 Perron-Frobenius theorem, 98 perturbed minimization principle, 395, 397 Pettis integral, 518 Pietsch factorization theorem, 64, 459, 840, 877, 950 plank problem, 182, 183 plurisubharmonic functions, 397, 637 plurisubharmonic perturbed minimization principle, 403 Poincar6 inequality, 350 point - w*-exposed, 628, 640 - w*-support, 628 denting, 634 -exposed, 601,628, 656 -extreme, 601,602, 605, 640 - farthest, 663 - nearest, 662 PSH-denting, 638 - smooth, 30, 640 -strongly exposed, 628 -support, 601,628 point of continuity property (PCP), 636 polar decomposition, 66 -
-
Subject Index polar decomposition theorem, 945 polydisk algebra, 874 polynomial map on a Banach space, 806 polytope, 650, 924 - or, 659 -/~, 659 tangent, 659 portfolio, 116 positive cone, 87 positive curvature, 923 positive definite distribution, 919, 923 positive definite function, 901,906, 909, 911 predictable or-algebra, 375 preference relation, 113 price space, 113 primary, 133, 865 principal ideal, 90 probabilistic method, 358 product of ideals, 444 projection, 4 - contractive, 255,256 partial sum, 7 partial sum for a decomposition, 11 Rademacher, 52 Riesz, 9 projection constant, 71,902, 931,933 absolute, 928 - relative, 928, 965 projective tensor norm, 485 projective tensor product, 285, 882 property T of Kazhdan, 332, 359, 619 proximal subgradients, 412 pure state, 333 put option, 116 -
~
-
a
p
p
r
o
x
i
m
a
t
i
n
g
-
-
-
-
-
quotient of ideals, 444 quotient of subspace theorem, 752 quotient space, 3 pX(T), 31 Rademacher - cotype q, 472 functions, 16, 125, 126, 460, 848, 850, 853,934 482, 845 theorem, 42 - type p, 471 Radon-Nikod~m property (RNP), 35, 38, 154, 236, 259, 260, 402, 405,414, 483,601,629 Ramanujan graphs, 338, 359 random matrices, 319, 321, 341,969 - edge of the spectrum, 345, 348 - global regime, 345 - local regime, 345 - spectral gaps, 345,358 random matrix ensembles, 343, 347 random orthogonal factorizations, 766 random variable r-stable, 6, 17 5, 16, 68 standard, 5 - symmetric, 6 random walk on the free group, 358 rank-one operator, 106 real interpolation, 502 real variable Hardy spaces, 879 rearrangement invariant space, 21,257 reasonable crossnorm, 485 reflexive algebra, 338 regular hull of an ideal, 445 ideal, 446 - norm, 91 relative Dixmier property, 335 replicate, 372 representable, )~-, 53 representable, finitely, 53 representation - 1-nuclear, 449 p-nuclear, 461 - dyadic, 453 finite, 441 Schmidt, 446 representing matrix, 625 reproducible, 132 restricted invertibility, 837-839, 854, 859-862 retraction continuous affine, 623 reverse Brascamp-Lieb inequality, 171 -
- p r o j e c t i o n ,
-
-
- G a u s s i a n ,
- G a u s s i a n ,
-
q(X), 284 quadratic perturbations, 396 quantum limit, 214 quartercircle law, 343 quasi-p-nuclear operator, 878 quasi-Banach ideal maximal, 456 ultraproduct-stable, 456 quasi-Banach space, 402 quasi-Cohen set, 875, 889, 890 quasi-convex bodies, 769 quasi-interior point, 104 quasi-Marcinkiewicz set, 879 ff quasi-norm, 402 quasidiagonality, 320, 324, 330 quasiidempotent measure, 875 quasireflexive space, 646 Quermassintegral, 727 -
-
1001
-
-
-
-
-
1002
Subject Index
H61der condition, 389 isoperimetric inequality, 163, 169 - metric approximation property, 293 Riemann ~" function, 345 Riesz-Kantorovich formulas, 90 Riesz-Thorin interpolation theorem, 75 risk-neutral, 373 Rochberg theorem, 703 Rodin-Semenov theorem, 519 Rosenthal ~ ! theorem, 18 Rosenthal inequality, 128, 149, 521 ff Rosenthal property, 445 rough norm, 796 Runge theorem, 673 -
-
SS(X, Y), 63 s-number ideals, 948 Saccone theorem, 686 Schatten-von Neumann classes, 447, 895 Schauder bases, 7 - bases with vector coefficients, 588-590, 594, 595 X-basis constant, 589 - equivalent X-bases, 594 unconditional X-basis constant, 589 results, 580 existence results, 581-583 in L p, 569-579 Schauder decomposition, 11,304 Schauder-Tichonoff theorem, 603 Schmidt representation, 446 Schoenberg problem, 901,906 Schr6dinger group, 197 Schur property, 9, 443 second order Hamilton-Jacobi equation, 430, 433 second order smooth minimization principle, 406, 424 second order subdifferential, 419 second order superdifferential, 420 security, 115 selection continuous affine, 622 semi-concave function, 420 semi-martingale, 375, 376, 382 semicircular distribution, 343 separable complementation property, 311 separating polynomial, 412 separation theorem, 601 sequence ideal, 446 sequence basic, 7 normalized, 7 seminormalized, 7 unconditionally basic, 10 - weakly Cauchy, 4
-
-
-
-
-
- n e g a t i v e
- n o n - e x p l i c i t
- u n c o n d i t i o n a l i t y
-
-
-
-
set - G~, 608 - w-compact convex, 635 1-norming, 655 antiproximinal, 664 - Chebyshev, 662 -closed convex bounded (CCB), 601 - compact convex, 602 - compact convex metrizable, 612 -convex stable, 601,660 - dentable, 634 - independent, 206 interpolation, 644 - Korovkin, 640 - locally compact, 629 - norming, 646 - proximinal, 601,662 - strongly antiproximinal, 664 -thin, 646, 647 - universally measurable, 631 -valued map, 622 - w e a k * compact convex, 611 with dense extreme points, 661 Sidak Lemma, 176, 179 Sidelnikov inequality, 932 Sidon constant, 885 Sidon problems, 200 Sidon sets, 205, 644, 872, 885-888, 891-893 sigma-martingale, 379, 380 simple function, 36 simple growth process, 112 simple predictable, 375 simplex, 613 - Bauer, 615, 616 - compact, 601, 613 - compact prime, 620 finite-dimensional, 613 - Poulsen, 618, 619 simplexoid, 626 singular numbers, 446, 945 skipped block sequence, 141,144 Slepian lemma, 229, 849 Slepian-Gordon lemma, 350 slice, 35, 397, 634 slicing problem, see hyperplane problem small isomorphism, 156 small perturbations, principle, 8 Smirnov domain, 701 smooth minimization principle, 406, 417 smooth point, 30, 640 smoothness, 515 Sobolev embedding operator, 469 Sobolev space, 970 -
-
-
-
-
Subject Index Sobolev-Besov imbedding theorems, 972 social endowment, 113 space - f-convex, 235, 237 - ARNP, 236, 262, 263 - A U M D , 236, 264 - Gurarii, 626 - Hilbert transform (HT), 235, 245 James tree, 636 -locally convex, 603 normed incomplete, 641 of securities, 115 - polyhedral, 650 quotient, 651 - rearrangement invariant, 21,257, 500 if, 574 - RNP, 236, 259, 260, 263,483,601,629 - U M D , 244, 250, 253, 264, 590, 894 spaces of vector-valued functions, 588-595 with values in a UMD space, 590 spaces with mixed norm, 107 spectral distance, 320, 327 spectral radius formula, 959 spectral theorem, 945 spherical design tight, 917 spherical harmonics, 926 formula, 926 spherical isoperimetric inequality, 715 spherical Radon transform, 921 spin glass theory, 354 splitting of atoms, 846, 847, 850 spreading models, 125, 136 square function, 27, 863,932 squares sets of, 206 stable, 137 - embedding, 524 -uniform algebras, 703 star body, 918-920, 923 state space, 626 stationary case, 426 Stein restriction conjecture, 219 Stein theorem, 883 Steiner symmetrization, 712 Steinhaus variables, 934 Stieltjes transform method, 344, 357 stochastic exponential, 383 stochastic interval, 375 stopping time, 375 Strichartz inequality, 197, 217 strict A~-functions, 404 strict .A~-set, 404 strictly - convex norm, 784 -
-
-
-
-
-
a
d
d
i
t
i
o
n
1003
positive functional, 22 striking price, 115 strong minimum, 398, 406 strong minimum at x0,420 strongly exposed point, 35, 628, 790 subdifferentiable function, 795 subdifferential calculus, 414 submartingale, 401 subspace - rich, 687 - smooth, 655 tight, 682 subspaces of Lp infinite dimensional, 140 ff finite dimensional, 837 ff Fourier transform characterization, 906 Sudakov inequality, 756 sufficiently Euclidean, 306 summation operator finite, 479 infinite, 443 super ideal, 455 - property, 56 - weakly compact operator, 479 support of a vector in a lattice, 23 surjection, 445 surjective hull of an ideal, 445 ideal, 446 symbol of an operator, 874 symmetric basis, 11,298, 854 symmetric ideal, 441 systems of functions, 563 -Daubechies wavelets, 567 -Faber-Schauder system, 9, 564 -Franklin system, 564, 574, 575,594 Haar system, 9, 564 ff - Meyer wavelets, 566, 576, 577 - polynomial bases, 577-579 system, 16, 125,460, 563,848 bases, 579 - spline wavelets, 567 systems of analytic functions, 569 tensoring, 568 -trigonometric polynomials, 577, 881,889, 896 -trigonometric system, 13, 564 - Walsh system, 564, 575,576 Szlenk index, 802 -
-
-
-
-
-
-
-
-
-
-
-
-
R
r
a
a
t
d
e
i
o
m
n
a
a
c
h
e
r
l
-
-
T2,279, 305 TP-smooth, 412, 413
Subject Index
1004 tail distribution, 6 Taylor expansion of order p, 412 tensor - norm, 485 153, 484, 916 stability, 959 tiling, 601,661 - bounded, 661 - convex, 661 total variation of a measure, 39 totally incomparable, 63 trace, 450 duality, 44, 351,456 - formula, 451,463,968 matrix, 968 - spectral, 968 transitive algebra, 536-538 translation invariant - space, 873 - subspace, 872 transport of measures, 350 tree, 139, 143, 144, 154 triangular truncation, 319, 329, 360 Tsirelson space, 276, 709 type, 49, 125, 126, 137, 139, 472, 505,526 - Gaussian, 472 - Haar, 477 - Fourier, 481 -
p
r
o
d
u
c
t
,
-
-
-
Up, 279 ultrafilter, 55 ultrapower, 55, 455 ultrapower-stability, 455 ultraproduct, 55,455 ultraproduct of operators, 55 ultraproduct, of Banach lattices, 55 ultraproduct-stability, 456 UMD Banach space, 253,590, 894 unconditional basic sequence, 125, 128, 131 - basis, 9, 126, 251,274, 277-279, 301,302, 304, 855, 858 -constant, 250, 251,256 -constant, complex, 250, 256 - convergence, 9 finite dimensional decomposition, 295, 298, 301, 304 finite dimensional expansion of the identity, 274, 278 structure, local (1.u.st.), 59 uniform - algebra, 673 -convexity, 413, 515, 527, 785 -
-
-
-
- Kadets-Klee property, 526 retract, 876 structure of a Banach space, 876 uniformly - A-dentable, 397 - F-smooth norm, 789 - convergent Fourier series, 895 -G~teaux-smooth norm, 789 integrable, 17, 143 unimodal, 178 uniqueness of complements, 837, 865-868 umqueness theorem for measures, 902 umt in a Banach lattice, 90 unitarily invariant norm, 328 unitary ideal property, 66 unitary orbit, 326, 327 universality conjecture, 345, 346 upper p-estimate, 504, 514 upper semicontinuous function, 608 Urysohn inequality, 728 -
-
-
Vaaler theorem, 175, 176, 179 vacuum state, 356 vacuum vector, 356 valuation, 732 viscosity solutions for Hamilton-Jacobi equations, 427 viscosity subsolutions, 427, 429, 430 viscosity supersolutions, 427, 429, 430 volume ratio, 169, 171,172, 174, 748 w*-H~-sets, 404 Walsh-Paley martingale, 476 wavelet basis, 565 wavelet set on R d, 566 weak Lp, 500, 505,523 weak Fatou norm, 89 weak Hilbert space, 277, 305, 313,968 weak type (1,1), 879-881 weak*-locally uniformly convex, 408 weakly continuous harmonic functions, 404 weakly unconditionally convergent (wuc) series, 686 weight function, 524 weighted norm inequalities, 387 weighted shift, 322, 323,329 Welfare theorems, 114 well-posed, 395 Wermer theorem, 673 Weyl inequality, 943,948, 953,957, 963 Weyl numbers, 945, 954, 956, 957, 970, 972 Weyl-von Neumann-Berg theorem, 324, 328, 330
Subject Index Wiener measure, 476 Wigner semicircle law, 342 Yan theorem, 380 Young inequality, 512
~'-function, 198 zonoids, 768, 844, 902, 911,924 -approximating zonotopes, 925 zonotopes, 768, 902, 924
1005
This Page Intentionally Left Blank