• Aucun résultat trouvé

Operator theory and exotic Banach spaces

N/A
N/A
Protected

Academic year: 2022

Partager "Operator theory and exotic Banach spaces"

Copied!
93
0
0

Texte intégral

(1)

Operator theory and exotic Banach spaces

(Banach spaces with small spaces of operators) Bernard Maurey

This set of Notes is a largely expanded version of the mini-course “Banach spaces with small spaces of operators”, given at the Summer School in Spetses, August 1994.

The lectures were based on a forthcoming paper [GM2] with the same title by Tim Gowers and the speaker. A similar series of lectures “Operator theory and exotic Banach spaces”

was given at Paris 6 during the spring of ’95 as a part of a program of three mini-courses organized by the “Equipes d’Analyse” of the Universities of Marne la Vall´ee and Paris 6.

We present in section 10, 11 and 12 several examples of Banach spaces which we call

“exotic”. The first class is the class of Hereditarily Indecomposable Banach spaces (in short H.I. spaces), introduced in [GM1]: a Banach spaceX is called H.I. if no subspace of X is the topological direct sum of two infinite dimensional closed subspaces. One of the main properties of a H.I. Banach space X is the following: every bounded linear operator T from X to itself is of the form λIX +S, where λ C, IX is the identity operator on X and S is strictly singular. It is well known that this implies that the spectrum of T is countable, and it follows easily that a H.I. space is not isomorphic to any proper subspace.

More generally, we present in section 11 a class of examples of Banach spaces having “few”

operators. The general principle is the following: given a relatively small semi-group of operators on the space of scalar sequences (for example, the semi-group generated by the right and left shifts), we construct a Banach space such that every bounded linear operator on this space is (or is almost) a strictly singular perturbation of an element of the algebra generated by the given semi-group. We obtain in this way in section 12 a new prime space, a space isomorphic to its subspaces with finite even codimension but not isomorphic to its hyperplanes, and a space isomorphic to its cube but not to its square.

We have chosen to present a fairly detailed account of all the tools of general interest that are necessary to the analysis, although these appear already in many classical books (but probably not in the same book); we develop elementary Banach algebra theory in section 2, Fredholm theory in section 4, strictly singular operators and strictly singular perturbations of Fredholm operators in section 6, an incursion into theK-theory for Banach algebras in section 9. Ultraproducts and Krivine’s theorem about the finite representability of `p are also presented in sections 7 and 8, with some emphasis on the operator approach to these questions. The actual construction of our class of examples appears in section 11, and the applications to some specific examples in the last section 12.

1. Notation

We denote by X, Y, Z infinite dimensional Banach spaces, real or complex, and by E, F finite dimensional normed spaces, usually subspaces of the preceding. Subspaces are closed vector subspaces. We write X = Y ⊕Z when X is the topological direct sum of two closed subspacesY andZ. The unit ball ofX is denoted byBX. We denote by Kthe field of scalars for the space in question (K=R or K=C).

We denote by L(X, Y) the space of bounded linear operators between two (real or complex) Banach spaces X and Y. When Y = X, we simply write L(X). We denote

(2)

by S, T, U, V bounded linear operators. Usually, S will be a “small” operator; it could be small in operator norm, or compact, or finite rank or strictly singular. . . ByIX we denote the identity operator fromX to X. Aninto isomorphismfrom X toY is a bounded linear operator T from X to Y which is an isomorphism between X and the image T X; this is equivalent to saying that there exists c > 0 such that kT xk ≥ckxk for every x ∈X. Let K(X, Y) denote the closed vector subspace ofL(X, Y) consisting of compact operators (we write K(X) if Y =X).

Anormalizedsequence in a Banach spaceX is a sequence (xn)n≥1 of norm one vectors.

The closed linear span of a sequence (xn)n≥1 is noted [xn]n≥1. A basic sequence is a Schauder basis for its closed linear span [xn]n≥1. This is equivalent to saying that there exists a constant C such that for all integers m≤nand all scalars (ak)nk=1 we have

°°Xm

k=1

akxk

°°≤C°

°Xn

k=1

akxk

°°.

The smallest possible constant C is called thebasis constant of (xn)n≥1.

An unconditional basic sequence is an (infinite) sequence (xn)n≥1 in a Banach space for which there exists a constant C such that for every integer n≥ 1, all scalars (ak)nk=1 and all signs (ηk)nk=1, ηk =±1, we have

°° Xn

k=1

ηkakxk

°°≤C°

° Xn

k=1

akxk

°°.

The smallest possible constant C is called the unconditional basis constant of (xn)n≥1. A question that remained open until ’91 motivated much of the research contained in these Notes: does every Banach space contain an unconditional basic sequence (in short: UBS).

The answer turned out to be negative and lead to the introduction of H.I. spaces.

2. Basic Banach Algebra theory

(For this paragraph, see for example Bourbaki, Th´eories spectrales, or [DS], or many others.) A Banach algebra A is a Banach space (real or complex) which is also an algebra where the product (a, b) →ab is norm continuous from A×A to A. This means that the product and the norm are related in the following way: there exists a constantC such that for all a, b∈A, we have

kabk ≤Ckak kbk.

It is then possible to define an equivalent norm on A satisfying the sharper inequality

∀a, b∈A, kabk ≤ kakkbk.

We shall call a norm satisfying this second property a Banach algebra norm. In order to define an equivalent Banach algebra norm from a norm satisfying the first property with a constant C, we may for example consider

|||a|||= sup{kab+λak:kbk ≤1,|λ| ≤1}.

(3)

We say that A is unital if there exists an element e A such that ea= ae = a for every a∈A; we write usually 1A for this elemente. If A is unital, we get an equivalent Banach algebra norm on A using the formula

|||a|||= sup{kabk:b∈A,kbk ≤1}

and for this norm|||1A|||= 1. A Banach algebra norm with this additional property will be called unital Banach algebra norm.

A C-algebra is a complex Banach algebra A with a Banach algebra norm and with an anti-linear involution a→a (i.e. a∗∗=a, (a+b) =a+b, (λa) =λa, (ab) =ba for every a, b∈A and λ C) and such that

∀a∈A, kaak=kak2.

If A is unital, then 1A = 1A and k1Ak = 1. An element x A is Hermitian (or self- adjoint) ifx =x. Everya ∈Acan be written as a=x+iy, where x andyare Hermitian (x= (a+a)/2,y =i(a−a)/2). At some point we will need the self-explanatory notion of a C-norm on a not complete algebra with involution.

The above definition is not satisfactory in the real case. Indeed, ifAis any real unital subalgebra of the complex algebraC(K) of continuous functions on a compact topological space K, and if we define on A the trivial involution f = f for every f A, then all properties of the preceding definition hold (because λ is now restricted to R), but A is not necessarily what we want to call a real C-algebra. In order to obtain a reasonable definition for the real case, we need to add an axiom which is consequence of the others in the complex case, but not in the real case, for example

∀a, b ∈A, kaak ≤ kaa+bbk.

Adding 1

When A has no unit it is possible to embed A in a larger unital Banach algebra, by considering onA+ =A⊕Kthe product (a, λ)(b, µ) = (ab+λb+µa, λµ). Then 1A+ = (0,1) is the unit of A+ and A is a closed two-sided ideal in A+. When A is a C-algebra, it is possible to define on A+ a C-norm.

An important example of unital Banach algebra isL(X) whereX is a (real or complex) Banach space. The operator norm is a unital Banach algebra norm onL(X). The subspace K(X) is a non unital closed subalgebra of L(X), actually a closed two-sided ideal ofL(X).

The algebra (K(X))+ is isomorphic to the subalgebra ofL(X) consisting of all operators of the form T =λIX +K, K compact (recall that X is infinite dimensional). The Calkin algebra C(X) = L(X)/K(X) is another important example. It will play a role for the notion of essential spectrum later in this section.

When H is a Hilbert space, L(H) is a C-algebra. It is a fundamental example since every C-algebra can be ∗-embedded in some L(H). The quotient of a C-algebra by a closed two-sided ideal is a C-algebra for the quotient norm (it is true but not obvious that for any such ideal I, we have x ∈I ⇒x ∈I); in particular the Calkin algebra C(H) of a Hilbert space H is a C-algebra.

(4)

Complexification

Most of spectral theory is done when the field of scalars isC. It is therefore important to be able to pass from the real case to the complex case. IfX is a real Banach space, the complexified space is XC =X⊕X with the rule

i(x, y) = (−y, x).

In this way we have (0, y) = i(y,0) and identifying (x,0) XC with x X, we see that every z ∈XC can be written as z =x+iy with x, y ∈X.

In order to define a complex norm on XC it is useful to think about XC as being C⊗X; it is clear that any tensor norm on C⊗X will give in particular a complex norm on XC (of course we choose the modulus as norm on the 2 dimensional real spaceC). We can consider for example the injective tensor norm CεX,

kx+iyk= sup{|x(x) +ix(y)|:x ∈X,kxk ≤1}.

With this norm C⊗X is isometric to (the real space) L(X,C). A (real) linear functional x on X is extended to a complex linear functional on XC by setting simply x(x+iy) = x(x) +ix(y), and it is easy to see that every complex linear functional on XC can be obtained as x +iy, where x, y X. Given T ∈ L(X, Y), one gets a complex linear operator TC ∈ L(XC, YC) by setting TC =Id⊗T in the tensor product language. In the previous language we can write TC(x+iy) =T x+iT y for all x, y ∈X.

WhenAis a real Banach algebra, it can be checked that if we normACbyAC =C⊗πA, we get a complex Banach algebra norm (if A had one); if A = L(X), X real, then AC identifies with L(XC) and any complex norm on XC gives a Banach algebra norm on AC; given V = T +iU in AC, T, U A = L(X), we associate the operator on XC = X⊕X given by the matrix

V =

µT −U

U T

.

Invertible elements

Let A be a unital Banach algebra over K. We say that a A is invertible (in A) if there exists x ∈A such that ax=xa= 1A.

Lemma 2.1. Let A be a unital Banach algebra (with a Banach algebra norm). If b∈ A and kbk<1 then 1−b is invertible in A.

Proof. (1−b)−1 =P

n=0bn.

Corollary 2.1. Let A be a unital Banach algebra. The set of invertible elements is open in A. Furthermore, if K =C, when a ∈A is invertible, the function f(z) = (a−zb)−1 is analytic in a neighborhood of 0 in C for every b∈A.

Proof. Writea−b=a(1−a−1b) and use Lemma 2.1. Ifkbk<ka−1k−1, thenka−1bk<1, and we obtain the following formula for the inverse

(a−b)−1 =u+ubu+ububu+· · ·,

(5)

where u = a−1. Applying to zb instead of b clearly gives an analytic function of z in a neighborhood of 0 in C,

(a−zb)−1 =u+z ubu+z2ububu+· · ·.

Remark 2.1. The same proof works when a is replaced by T ∈ L(X, Y) invertible, and b is replaced by a small operator S ∈ L(X, Y); the above formulas hold with u = T−1 L(Y, X), provided kSk<kT−1k−1.

Spectrum of an element, resolvent set

Let A be unital over C and let a A. The resolvent set ρ(a) is the set of all λ C such thatλ1A−a is invertible inA. Thespectrumσ(a) is the complementary set C\ρ(a).

This set ρ(a) is open by Corollary 2.1 and is clearly a neighborhood of infinity, henceσ(a) is a closed and bounded subset of C.

Exercise 2.1. Leta, b∈A. Show thatσ(ab)\{0}=σ(ba)\{0}(hint: ifz(λ1A−ab) = 1A, find u∈A such that u(λ1A−ba) =λ1A). Why did we exclude 0?

Real spectrum

If A is a real unital Banach algebra and if a ∈A we may define a real spectrum by σR(a) = R: (a−λ1A) not invertible in A}.

The problem is that this spectrum may be empty (contrary to what happens in the complex case: we shall recall that the spectrum is non empty in the complex case). If we extend the scalars and consider the same a ∈A as an element of AC, we can work with the complex spectrum σAC(a).

Exercise. Let A be a real unital Banach algebra and let a∈A.

1. Show that σAC(a) =σAC(a) 2. Show that σR(a) =R∩σAC(a).

In what follows we shall denote by σK(a) the real spectrum when A is real and the complex spectrum if A is complex.

Changing the ambient algebra

Assume thatA is a closed subalgebra of a unital Banach algebra B, with the induced norm and containing 1B. If x A is invertible in A, it is obviously also invertible in B, but it is possible for x A to be invertible in B but not in A. It is necessary in this case to distinguish the spectrum of x relative to B or relative to A; we denote the spectrum by σAK(x) orσKB(x), and similarly for the resolvent sets. If x∈A, it is clear that σAK(x) σKB(x) (equivalently, ρKA(x) ρKB(x)). We also have ∂σKA(x) ∂σKB(x) from the next Lemma.

Lemma 2.2. Let B be a unital Banach algebra, and let A be a closed subalgebra with 1B A. For every x∈A, the resolvent setρKA(x) is open and closed in ρKB(x). Hence, for every connected componentω of ρKB(x), eitherω ⊂ρKA(x) or ω∩ρKA(x) =∅. It follows that

∂ρKA(x) =∂σAK(x)⊂∂σBK(x) =∂ρKB(x). If ρKB(x) is connected, then σAK(x) =σBK(x).

(6)

Proof. Letxbe fixed inAand write simplyρA, ρB forρKA(x) andρKB(x). We know thatρAis open. Let (λn)⊂ρA,λn →λ∈ρB. There existsb∈Bsuch that (x−λ)b=b(x−λ) = 1B. On the other hand, for everynthere existsan ∈Asuch that (x−λn)an = 1B. Multiplying by b we get

b= (bx−λnb)an= (1B + (λ−λn)b)an;

when n→ ∞ we see that (1B+ (λ−λn)b)−1 tends to 1B, thus an →b, hence b∈A and λ∈ρA.

Let λ ∂ρA. Then λ /∈ ρA since ρA is open in K, and therefore λ /∈ ρB since ρA is closed in ρB; this implies that λ ∈∂ρB.

Remark. When σA(x)6=σB(x), we see that the interior int(σA(x)) must be non empty.

Examples 2.1.

1. It will be shown later (section 6, Proposition 6.1) that the spectrum (inB =L(X)) of a strictly singular operator S on a complex Banach space X is a countable compact subset of C. This implies that the resolvent set ρB(S) is connected, hence the inverse of λIX−S, when it exists, belongs to the closed subalgebraA of L(X) generated by IX and the operator S.

2. Suppose that A is a complex unital C-subalgebra of L(H) and let u A be hermitian. Since the spectrum of uinB =L(H) is contained inRand bounded, it is clear that the resolvent setρB(u) is connected, henceσA(u) =σB(u). This remark implies easily that σA(x) =σB(x) for every x A (it is enough to show that when x A is invertible in B its inverse belongs to A; if x is invertible inB, then the hermitian operators xx and xx are invertible inB, hence inA, thereforex is invertible inA: there existy, z ∈A such that y(xx) = (xx)z = 1A, and yx = xz is the inverse of x in A). More generally, if an isomorphism from H ⊕H onto H is given by a matrix (a, b) with a, b∈A, the inverse operator from H to H⊕H is given by a matrix with two entries in A.

3. We can always embed a Banach algebraAin the space of bounded linear operators on some Banach space. Suppose that A is a unital Banach algebra with a unital Banach algebra norm. We simply embed A into L(A) by mapping each a∈A to the operator Ma

of left multiplication by a, defined by Ma(x) = ax for every x A. It is clear that this gives an isometric embedding fromAinto L(A). We show now that the spectrum ofa ∈A is the same relative to A or to the larger algebraB =L(A). We only need to show that a is invertible inA iffMa is invertible inB. One direction is obvious. In the other direction, assume that Ma is invertible in B. Then Ma is onto and there exists u A such that 1A = Ma(u) = au. We see that MaMu = 1B. Since Ma is invertible in B, this implies MuMa = 1B and ua=au= 1A shows that a is invertible in A.

4. Suppose that A is a real unital Banach algebra; if a A is invertible in AC, then a is invertible in A.

5. Consider the embedding of L(X) into L(X) given by the adjoint (this is not exactly an algebra morphism since (U T) = TU); the spectrum of T in L(X) is the same as the spectrum of T ∈ L(X) (of course this is totally obvious when X is reflexive).

(7)

Spectral radius

Let a∈A. The spectral radius of a is defined by r(a) = lim

n kank1/n. Exercise. Show that the above limit exists.

Example 2.2. If u is hermitian in a C-algebra, we get ku2nk= kuk2n for every integer n≥1, thus r(u) =kuk.

Notice that the spectral radius is not changed if the norm on A is replaced by an equivalent norm. Also, r(a)≤ kak if the norm is a Banach algebra norm.

Proposition 2.1. (K=Cand 1A 6= 0)The spectrum σ(a) is contained in the closed disc

∆(0, r(a)) inC centered at 0 and of radiusr(a), and it intersects the circle of radius r(a).

In other words,

r(a) = max{|λ|:λ ∈σ(a)}.

In particular, σ(a)is non empty.

Proof. The functiong(z) = (1A−za)−1 is clearly defined and analytic when |z|< r(a)−1, henceσ(a)⊂∆(0, r(a)). Since 1A 6= 0 we may assumek1Ak= 1 for some equivalent unital Banach algebra norm. If a is invertible, it is then easy to see that r(a)r(a−1) 1, hence r(a)>0. If r(a) = 0, we know therefore that a is not invertible, thus 0 ∈σ(a) and finally σ(a) ={0} in this case.

Assume now r(a) > 0. If the inverse (1A −za)−1 exists for every z on the circle

|z|=r(a)−1, we can show that the functiong is analytic in a neighborhood of a closed disc of radius R > r(a)−1. It follows then from Cauchy’s inequalities that for some constant M, we have

∀n≥0, kank ≤ M Rn yielding r(a)≤1/R and contradicting the choice ofR.

Resolvent equation. Spectral projections

Let A be a unital Banach algebra over C and let a A. The resolvent operator of a is defined for z ρ(a) by R(z) = (z1A −a)−1. Note that R(z) and R(z0) commute and commute with a. We have

R(z0)−R(z)

z0−z =−R(z0)R(z).

Letλ∈Cbe isolated inσ(a). Let ∆rbe the closed disc ∆(λ, r) with radiusr centered at λ; let r0 > 0 be such that λ is the unique point of σ(a) contained in ∆r0; let γr be the boundary of ∆r, oriented in the counterclockwise direction. Let f be holomorphic in a neighborhood V of λ; for r > 0 such that ∆r ⊂V and r r0 we know that γr ρ(a) and we set

Φ(f) = 1 2iπ

Z

γr

f(z)R(z)dz ∈A.

(8)

Since f is holomorphic, the result does not depend upon the particular value r (0, r0] such that ∆r V. Also Φ(f) commutes with a and with every R(µ). The mapping f Φ(f) is clearly linear with respect to f. What is more interesting is that

Φ(f g) = Φ(f)Φ(g).

For the proof let 0< r < s≤r0 be such that f and g are holomorphic in a neighborhood of ∆s and write

Φ(f)Φ(g) = 1 2iπ

Z

z∈γr

1 2iπ

Z

z0∈γs

f(z)g(z0)R(z)R(z0)dzdz0;

use then the resolvent equation and Cauchy’s formula. It follows that for every r with 0< r≤r0

p= Φ(1) = 1 2iπ

Z

γr

(z1A−a)−1dz

is an idempotent (p2 =p), commuting with a and with every R(µ); furthermore for every element b= Φ(g)∈A we have

pb=bp=b=pbp since pΦ(g) = Φ(1)◦Φ(g) = Φ(g).

To every idempotent p in A we may associate the Banach algebra Ap = pAp of all elements of the formpap,a ∈A. As a Banach algebra, the norm and the product inAp are those of A, but the unit ofAp isp. For example, let pbe a bounded projection defined on some Banach space X, and let Y =p(X) be the range of p. We see that L(X)p identifies with L(Y) (with an equivalent norm).

Let us come back to our isolated λ ∈σ(a) and let again p= Φ(1). The above remark shows that Φ(f)∈Ap for everyf holomorphic in a neighborhood of λ. Also notice that

Φ(z)−ap= 1 2iπ

Z

γr

(z1A−a)R(z)dz= 0,

so thatap= Φ(z). Suppose thatf(λ)6= 0; theng= 1/f is holomorphic in a neighborhood ofλ, and Φ(f)◦Φ(1/f) = Φ(1) =p. It follows that Φ(f) is invertible inAp whenf(λ)6= 0.

This applies in particular to f(z) =z−µwhen µ6=λ to show that Φ(z−µ) =ap−µp is invertible in Ap. This shows that the spectrum ofpa=papin Ap reduces to {λ}. Letting q = 1A−p, the spectrum of qa in Aq does not contain λ. This follows from

(λ1A−a)³ 1 2iπ

Z

γr

R(z) z−λ dz´

=q.

More generally, when the spectrum σ(a) can be decomposed into two subsets σ1 and σ2 open and closed inσ(a), we can construct similarly a spectral projection pby replacing the above circle γr by a curve γ around σ1, such that σ2 is exterior to γ. We still get pa=ap and σ(pa) =σ1 (in Ap) and similarly in Aq we have that σ(qa) =σ2.

(9)

Exercise. Suppose that ka2−ak< ε < 1/4; show that there exists an idempotent p such that ap = pa and kp−ak < f(ε,kak) (prove that σ(a) is contained in the union of the interiors of two circles γ1 and γ0 with radius 1/2 and centered at 1 and 0, for example by considering (a−t)(a−(1−t)) for|t(1−t)| ≥1/4; give an upper bound fork(z−a)−1kwhen z belongs to γ0 or γ1; let Φ0 and Φ1 be the operators as above associated with the two circles, and consider p = Φ1(1); use b = Φ1(1/z) for proving that (p−ap) = (1A−a)abp is small, and similarly for (1A−p)a).

Essential spectrum

Let X be an infinite dimensional Banach space. The ideal K(X) is then proper and the Calkin algebra C(X) = L(X)/K(X) is not {0}. For every T ∈ L(X), the essential spectrum bσK(T) is the spectrum of the image Tb of T in C(X). We also consider the corresponding resolvent set ρbK(T). A scalar λ K belongs to this essential resolvent set iff T −λIX is invertible modulo compact operators. We shall recall in section 6 that this happens iff T −λIX is a Fredholm operator on X.

Commutative Banach algebras

A (complex) Banach algebraA which is a (skew-)field is isomorphic to C. Indeed, let a A and λ σ(a). Since a−λ1A is not invertible, we must have a−λ1A = 0. Hence every element of A is of the form λ1A for some λ∈C.

Let A be a unital commutative Banach algebra over C. A maximal ideal I is closed and is an hyperplane: first, I is closed because the set of invertible elements is open, hence the closure ofIis still a proper ideal, equal toIsinceIis maximal; second,A/I is a Banach field, therefore isomorphic to C, andI is thus an hyperplane. The linear functional χ such that kerχ =I, normalized by the condition χ(1A) = 1, is called a character. A character χ is a non zero bounded linear functional on A which is also multiplicative. Indeed, if a ∈A and χ(a)6= 0, let g(x) =χ(ax)/χ(a). Theng vanishes on I, so g is proportional to χ, but χ(1A) =g(1A) = 1, thus χ=g and χ(ax) =χ(a)χ(x) for every x∈A.

The set of characters on A is called spectrum of A, and denoted by Sp(A); suppose thatA is equipped with a Banach algebra norm; then Sp(A) is a subset of the unit sphere of the dual space A; to see this, observe that the sequence (an)n≥0 is bounded when kak ≤ 1, so that (χ(an)) = (χ(a)n) is bounded, therefore |χ(a)| ≤ 1, and kχk = 1 since χ(1A) = 1.

An element a A is invertible in A iff χ(a) 6= 0 for every characterχ on A. Indeed, it is clear that χ(a) 6= 0 when a is invertible; if a is not invertible, aA is a proper ideal in A, thus contained in a maximal ideal I. If χ is the character such that I = kerχ then χ(a) = 0. It follows that for everya ∈A,

σ(a) ={χ(a) :χ∈Sp(A)};

since χ(a−χ(a)1A) = 0, we see that χ(a) σ(a) for every χ Sp(A). Conversely, if a−λ1A is not invertible, there exists a character χ such that χ(a−λ1A) = 0.

Furthermore Sp(A) isw-closed in the unit sphere; this allows to map A to the space C(Sp(A)) of continuous functions on the compact space Sp(A); for every a A, let j(a)

(10)

be the continuous function on Sp(A) defined by j(a)(χ) = χ(a); this map j need not be injective in general.

When A is a unital (complex) commutative C-algebra, this embedding is isometric;

we first observe that when u is hermitian, a = eiu is unitary, i.e. aa = aa = 1A; then a = a−1 and kak =ka−1k = 1; this implies that the sequence (an)n∈Z is bounded, thus (χ(an))n∈Z is bounded and therefore 1 = |χ(a)| = |eiχ(u)|, hence χ(u) is real for every hermitian u; this implies

∀a ∈A, χ(a) =χ(a).

If u is hermitian, we have r(u) = kuk by Example 2.2, hence there exists χ Sp(A) such that |χ(u)|=kuk. This implies that for every a∈A, there exists χ such that

|χ(a)|2 =χ(aa) =kaak=kak2,

showing that the mapping j :A→C(Sp(A)) is isometric in this case. The image ofA into C(Sp(A)) is now a subalgebra closed under complex conjugation and obviously separating points of Sp(A), therefore our embedding is onto by Stone-Weierstrass’ theorem.

Suppose that ϕ is a (unital) ∗-homomorphism between two unital commutative C- algebras, thenkϕk ≤1; by the preceding paragraph we may think thatA=C(K) and B= C(L), where K andL are two compact topological spaces. Since ϕ is a∗-homomorphism, it sends every functionf 0 onK toϕ(f)≥0 (introduce the hermitian elementg=

f).

If kfk ≤ 1, then ff 1, so that ϕ(ff) 1 and kϕ(f)k ≤ 1; furthermore if ϕ is injective then ϕ is isometric; this is because the adjoint map ϕ : (C(L)) (C(K)) sends Sp(B) ' L into Sp(A) ' K, and must be onto by the preceding result (otherwise we may find a continuous functionf 6= 0 supported onK\ϕ(Sp(B)), and thenϕ(f) = 0, contradicting the injectivity). Observe that we don’t need B to be commutative in the above argument, because the range ϕ(A) is commutative, so that its closure in B is a commutative unital C-algebra.

Proposition 2.2. Let B be a non necessarily commutative unitalC-algebra. For every unital C-algebra C, every unital ∗-homomorphism ϕ : B C satisfies kϕk ≤ 1. If ϕ is injective, then ϕ is isometric.

Proof. Given an hermitian element u ∈B, we may consider the unital subalgebra A of B generated by u. This is a commutative C-algebra. We have seen that the spectrum of u in A is real, hence σA(u) = σB(u) by Lemma 2.2. Suppose that ϕ is a ∗-homomorphism from B to some C-algebra; restricting ϕ to A, it follows from the preceding remark that kϕ(u)k ≤ kuk, and this is true for every hermitian u ∈B. For a general b∈B, we write

kϕ(b)k2 =k(ϕ(b))ϕ(b)k=kϕ(bb)k ≤ kbbk=kbk2.

Suppose further that ϕ is injective; then ϕ is isometric; indeed, the preceding remarks show that kϕ(u)k=kuk when u is hermitian; for a general b∈B, we write

kϕ(b)k2 =k(ϕ(b))ϕ(b)k=kϕ(bb)k=kbbk=kbk2.

(11)

Wiener’s algebra

The Wiener algebra W is the algebra of continuous (complex) functions on T with absolutely summable Fourier coefficients; the product is the pointwise product and the norm is the `1-norm of Fourier coefficients. This algebra W is clearly isometric to `1(Z) with its usual norm and the convolution as product.

A function f in W is invertible in W iff f does not vanish on T. This amounts to showing that the characters on W reduce to evaluation at points of T. Let χ be a character on W and let λ = χ(e). Since the family (einθ)n∈Z is bounded in W, it follows that the sequence (λn =χ(einθ))n∈Z is bounded, hence |λ|= 1. For every function f(θ) =P

n∈Zaneinθ in W,

χ(f) = X

n∈Z

anλn =f(λ).

Iff does not vanish on T, we see thatχ(f)6= 0 for every characterχ, hence f is invertible in W.

C-algebra summary

(see first pages of Pedersen’s book [Pd])

Let A be a unital (complex) C-algebra. For every hermitian element b∈A, we may consider the unital subalgebra B generated by b; it is a commutative C-algebra. We know that jB : B C(Sp(B)) is an onto isomorphism. For example, when b = b and σA(b) [0,∞), then σB(b) = σA(b) by Lemma 2.2; the function jB(b) is a non-negative continuous function on Sp(B), therefore there exists c B which is the “square root” of b, i.e. c=c, b=c2 andσ(c)⊂[0,∞). Consider

C ={a ∈A:a=a and σ(a)⊂[0,∞)}.

If b = b and a = b2, then a C; this follows from the commutative theory, but also simply from the fact that σ(b) R which implies that for every t > 0, b2 + t1A = (b+i√

t1A)(b−i√

t1A) is invertible.

If a = a and kt1A−ak ≤ t for some t 0, then a C. This is clear when t = 0;

when t > 0, we have k1A −a/tk ≤ 1 and we may define b = p

1A(1A−a/t) = b from its Taylor series. Then a = (

t b)2 C. Conversely, if a C and t = kak, then kt1A−ak ≤t, so that we got a characterization ofC; this last fact is because we know for every hermitian element a that r(a) =kak=t, and

kt1A−ak=r(t1A−a) = max{|t−λ|:λ∈σ(a)⊂[0, t]} ≤t.

It follows immediately from this characterization that a1, a2 ∈C implies a1+a2 ∈C, so that C is a closed convex cone in A, with C∩(−C) ={0}.

The next essential step is to prove thataa∈C for every a ∈A. This is easy once we know that we may embed A as a ∗-subalgebra of some L(H), but it is not obvious from the abstract definition, and it is a main ingredient for the proof of the representation of an abstract C-algebra as a subalgebra of some L(H).

(12)

Observe first that for every b = r+is∈ A, r, s hermitian, we have that bb+bb = 2(r2+s2)∈C. Leta∈A. The commutative theory implies that we can writeaa =u−v, with u, v C and uv = vu = 0. Let b= a√

v; then bb = −v2 ∈ −C, and bb ∈C since bb+bb C; by Exercise 2.1, this implies that σ(bb) = {0}, hence r(bb) = kbbk = kbk2 = 0, thus v2 = 0, and finally v = 0, aa =u∈C.

For every u 6= 0, we know that −uu /∈ C. We may therefore find by Hahn-Banach a linear functional ξ on A such that −ξ(uu)≤ infξ(C); this yields that ξ 0 on C. we define a scalar product on A by

ha, bi=ξ(ab).

To every a∈A we associate the operator Ta(b) =ab. Let a∈A with kak ≤1. We have hTa(b), Ta(b)i=ξ(baab);

Since kak ≤ 1, we know that 1A−aa =c2 ∈C and bb−baab = bc2b ∈C, therefore ξ(baab)≤ξ(bb) =hb, bi. This shows that

∀a∈A, kTakL(H)≤ kak,

where H is the Hilbert space obtained from the above scalar product. It is easy to check that (Ta) =Ta, so that we have a∗-homomorphism from A to L(H). We may improve the argument and obtainkTuk=kukfor the givenuand then find an isometric embedding of A into some L(H) using a direct sum of such embeddings (if kuk= 1, observe that the closed convex cone uu+C is disjoint from the open unit ball, and use the separation theorem as before to obtain ξ such that ξ(uu) =kξk= 1).

Suppose now that I is a closed two-sided ideal in A, and let a I. Let |a| = aa;

then|a| ∈I; indeed, there exists a sequence (Pn) of polynomials with real coefficients such that Pn(0) = 0 and such that (Pn(t)) converges to

t uniformly on any compact interval [0, T], so that for every x C ∩I, we have

x I. For every ε > 0, ε21A +aa is invertible, thus there exists uε ∈A such that

a=p

ε21A+aa uε. We obtain from the commutative theory

kuεk2 =k(ε21A+aa)−1/2aa(ε21A+aa)−1/2k ≤1.

When ε 0 we obtain that a= limε→0|a|uε hence a = lim

ε uε|a| ∈I.

Finally, let us say some words about the real case. Let A be a real Banach algebra with involution and with a Banach algebra norm such that

∀a, b∈A, kak2 ≤ kaa+bbk.

(13)

This implies as in the complex case that every hermitian elementx ∈A as a real spectrum inAC (we don’t claim so far thatAC can be equipped with aC-norm). Indeed,a= costx andb= sintxare hermitian for every realt, and a2+b2 = 1A, therefore by our hypothesis kcostxk ≤1 and ksintxk ≤1 for every real t, which implies that eitx is bounded in AC, from which it follows that the spectrum of x is real. With this information, it is possible to reproduce the arguments from the beginning of this paragraph and to ∗-embed A in L(H), for some real Hilbert space H.

Exercise. Complete the details. Show first that if B is the (real) subalgebra generated by an hermitian element a∈A, and if x, y ∈B,

kx+iyk=kx2+y2k1/2A

is a C-norm on BC. If v A is anti-hermitian, i.e. v = −v, observe that evev = 1A and kevk= 1. It follows that σ(v)⊂iR, and v2−t21A is invertible for t real, t 6= 0, thus σ(−v2)[0,+∞).

3. Some operator theory: finitely singular operators See for example [LT1], section 2.c.

Lemma 3.1. Let X andY be real or complex Banach spaces.

1. Trivial principle: ifT is an isomorphism fromX into Y, a small norm perturbation T +S of T is still an into isomorphism.

2. Fundamental principle: if T is an isomorphism from X onto Y, a small norm perturbation T +S is still an isomorphism from X onto Y (this is Remark 2.1).

Lemma 3.2. Let T ∈ L(X, Y) be an into isomorphism and k≥0 an integer.

1. Suppose that codimT X k. There exists c > 0 such that codim(T +S)X k whenever kSk< c.

2. Suppose that codimT X = k. There exists d > 0 such that codim(T +S)X = k whenever kSk< d.

Proof. If codimT X ≥k we can find a subspaceF ⊂Y such that dimF =kandT X∩F = {0}. Let πF denote the quotient map Y →Y /F. Then πF ◦T is an isomorphism from X into Y /F. Ifc >0 is small enough and kSk< c, πF (T +S) is also an into isomorphism by Lemma 3.1. This implies that F (T +S)X ={0}, hence codim(T +S)X ≥k.

In the second case the proof is similar, but we can now select a subspaceF such that Y =T X ⊕F, dimF = k. Then πF ◦T is an onto isomorphism, hence there is d >0 (we may choose d < c), such thatπF (T +S) is an onto isomorphism when kSk< d. By the above argument we already know that F∩(T +S)X ={0}; furthermore, for everyy ∈Y, there exists x ∈X such that πF(y) = πF((T +S)x), so y−(T +S)x F, showing that Y =F + (T +S)X. Finally Y =F (T +S)X and codim(T +S)X =k.

Proposition 3.1. Let T ∈ L(X, Y) be an into isomorphism. Then T +S is an into isomorphism andcodim(T+S)X = codimT X(finite or+∞) for everySin a neighborhood of 0 in L(X, Y).

Proof. Let c >0 be such that T +S is an into isomorphism whenever S ∈Bc ={U ∈ L(X, Y) :kUk< c}.

(14)

Observe that Dk = {U Bc : codim(T +U)X = k} is open and closed in Bc for every integer k 0. Indeed, the set {W Bc : codim(T +W)X k + 1} is open by Lemma 3.2, part 1, while each Dj, j = 0,1, . . . , k is open by part 2 of the same Lemma. Since Bc

is connected, each Dk is empty or equal to Bc. The result follows.

Boundary of spectrum lemma

Lemma 3.3. If U ∈ L(X) is an into isomorphism but is not invertible in L(X), then 0 belongs to the interior int(σK(U)) (relative toK) of σK(U).

Proof. SinceU is an into isomorphism but not invertible,U is not onto and codimU X 1.

This remains true under small perturbation by Lemma 3.2, part 1: there exists ε0 > 0 such that U −εIX is not onto, therefore not invertible, for every ε∈K such that|ε|< ε0. This shows that B(0, ε0)K⊂σK(U).

Corollary 3.1. Let T ∈ L(X). If λ∈ ∂σK(T) (of course this boundary is relative to K), there exists a normalized sequence (xn) in X such that (T −λIX)xn 0 (this sequence is possibly constant).

Proof. What we want to prove is equivalent to saying that U = T −λIX is not an into isomorphism. By our assumption, we have thatU is not invertible, but 0 is not interior to σK(U), henceU is not an into isomorphism by Lemma 3.3.

Exercise. If X is real, T ∈ L(X) and if λ = r(cosθ +isinθ), rsinθ 6= 0, belongs to the boundary of σ(TC), then there exist two sequences (xn) and (yn) in X such that kxnk+kynk= 1, T xn−r(cosθ xnsinθ yn)0 and T yn−r(sinθ xn+ cosθ yn)0.

Remark 3.1. Let A be a unital Banach algebra. If λ belongs to the boundary of σK(a), there exists a normalized sequence (bn) in A such that abn −λbn tends to 0 in A. This follows from section 2, example 2.1,3 (we could also get a normalized sequence (cn) such that cna−λcn goes to 0).

Definition 3.1. Let T ∈ L(X, Y); we say that T is finitely singular if there exists c > 0 and a (closed) finite codimensional subspace X0 ⊂X such that

kT xk ≥ckxk

for every x∈X0. In other words the restriction of T to X0 is an into isomorphism.

Let c(T) denote the supremum of all c > 0 for which the above property holds, and set c(T) = 0 if T is not finitely singular.

Our terminology is not classical and perhaps a little strange, since we call a non singular operator, for instance an onto isomorphism, “finitely singular”; it would be better to say “at most finitely singular”, but this is definitely too long.

Remark 3.2. Suppose thatT ∈ L(X, Y) is finitely singular. It is clear that a small norm perturbation ofT is still finitely singular (precisely,T+S is finitely singular ifkSk< c(T);

actually it is enough thatkS|X1k< c(T) for some finite codimensional subspaceX1 of X).

It is also clear that the restriction of T to any infinite dimensional subspace Z of X is finitely singular.

(15)

If T ∈ L(X, Y) and if U T is finitely singular for some U ∈ L(Y, Z), then T is finitely singular.

Exercise 3.1. Suppose that T ∈ L(X, Y) is finitely singular. Show that 1. kerT is finite dimensional.

2. For every (closed) subspace Z of X, T(Z) is a closed subspace of Y.

3. If (xn) is a bounded sequence in X and if (T xn) converges in Y, then there exists a norm-converging subsequence (xnk); in other words, the restriction ofT to any bounded subset of X is a proper map.

4. One can choose X0 in Definition 3.1 in such a way that X = kerT ⊕X0. Hence, if kerT ={0}, then T is an isomorphism from X into Y.

5. Let T ∈ L(X, Y). Show that T is finitely singular if and only if T X is closed and dim kerT <∞.

6. If T is finitely singular from X to Y and K = R, the complexified operator TC is finitely singular from XC to YC.

Lemma 3.4. Let T ∈ L(X, Y). Then T is finitely singular from Y to X if and only if T X is closed and finite codimensional in Y.

Proof. Suppose that T is finitely singular. Since kerT = (T X) is finite dimensional, we know that T X is finite codimensional. It is enough to show that for some c > 0 and for every y T X, there exists x X with ky−T xk ≤ kyk/2 and kxk ≤ kyk/c (the end of the proof is by iteration: one constructs a convergent series x0 =P

xn in X such that y =T x0). If the preceding claim is not true, we can find for every integer n≥ 1 a vector yn ∈T X such that kynk= 1 and

yn ∈/ nT(BX) + 1 2BY.

By Hahn-Banach there existsyn ∈Ysuch thatyn(yn) = 1 andkynk ≤2,kT(yn)k ≤1/n.

Since T is finitely singular, we know from Exercise 3.1,3 that there exists a subsequence (ynk) converging to somey; it follows then thatTy = 0, thusy kerT, which implies that y(yn) = 0 for every n, contradicting yn(yn) = 1 andynk →y.

Conversely, assume T X closed and finite codimensional in Y. By the open mapping theorem, T induces an isomorphism from X/kerT onto T X. Let Y0 = T X and let c > 0 be such that for everyy0 ∈Y0, there existsx∈X such that y0 =T x andky0k ≥ckxk. Let Y =Y0 ⊕F and let Q be the projection from Y0⊕F onto Y0. Given y F, kyk= 1, there exists y = y0+f ∈Y0⊕F such that ky0 +fk ≤1 and y(y0+f) =y(y0) >1/2.

Then, since ky0k=kQyk ≤ kQk there existsx ∈X such that y0 =T x and kxk ≤ kQk/c, hence

kQk

c kTyk ≥T(y)(x) =y(y0)>1/2,

showing that kTyk ≥ ckyk/(2kQk) for every y in the finite codimensional subspace F of Y, hence T is finitely singular.

We say that T is infinitely singular if it is not finitely singular.

Proposition 3.2. Let T ∈ L(X, Y). Then T is infinitely singular if and only if for every ε > 0, there exists an infinite dimensional subspace Z X such that kT|Zk < ε.

Références

Documents relatifs

The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.. L’archive ouverte pluridisciplinaire HAL, est

In particular, we want the Fourier transform to be a complex valued function defined on some explicit ‘frequency space’ that may be endowed with a structure of a locally compact

This hierarchy also provides us with a theorem of the alternative (or duality theorem) which uses a non linear polynomial, a discrete analogue of the celebrated Farkas Lemma in

We will say that S (resp. However, despite the extensive literature related to Toeplitz operators, it seems that it is not known if every Toeplitz operator is induced by a multiplier

Keywords: Fourier transform, Heisenberg group, frequency space, Hermite functions.. 2010 Mathematics Subject Classification:

Analyticity of relative fundamental solutions and projections for left invariant operators on the Heisenberg group.. Annales scientifiques

Important subgroups { Φ � 2s+1 } are introduced and results are proved for these groups (Lemma 5.3 and Theorem 5.4) that play a crucial role in finding the dimension and coordinates

For any connected, topological space X, XK shall denote the space of continuous maps of K into X equipped with the compact- open topology.. The example in the